homogeneous nickel and iron complexes for carbon dioxide

288
Homogeneous Nickel and Iron Complexes for Carbon Dioxide and Methanol Transformations: Progress toward Commercial Chemical Production By Elizabeth M. Lane B.A., Dartmouth College 2008 M.A., Brown University 2015 A Dissertation Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the Department of Chemistry at Brown University Providence, Rhode Island May 2018

Upload: others

Post on 20-May-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

Homogeneous Nickel and Iron Complexes for Carbon Dioxide and Methanol

Transformations: Progress toward Commercial Chemical Production

By

Elizabeth M. Lane

B.A., Dartmouth College 2008

M.A., Brown University 2015

A Dissertation Submitted in Partial Fulfillment of the Requirements for the Degree of

Doctor of Philosophy in the Department of Chemistry at Brown University

Providence, Rhode Island

May 2018

Page 2: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

© Copyright 2018 by Elizabeth M. Lane

Page 3: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

iii

This dissertation by Elizabeth M. Lane is accepted in its present form by the Department

of Chemistry as satisfying the dissertation requirement for the degree of Doctor of

Philosophy.

Date: ________________ ____________________________

Wesley H. Bernskoetter, Advisor

Recommend to the Graduate Council:

Date: _________________ ____________________________

Eunsuk Kim, Reader

Date: _________________ ____________________________

Paul Williard, Reader

Approved by the Graduate Council:

Date: _________________ _________________________________________

Andrew G. Campbell, Dean of the Graduate School

Page 4: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

iv

Curriculum Vitae

Elizabeth M. Lane graduated from Mohawk Trail Regional High School in 2004 as

valedictorian of her class and attended Dartmouth College in Hanover, New Hampshire for

her undergraduate studies. At the end of her junior year, she received an Arnold and Mabel

Beckman Research Scholarship and began undergraduate chemistry research in the

laboratory of Professor David Glueck, where she pursued the synthesis and structural

characterization of novel gold-phosphido complexes and completed an honors thesis. She

graduated magna cum laude from Dartmouth in 2008 with a Bachelor of Arts in Chemistry

and Philosophy. She then attended the University of Chicago and worked under the tutelage

of Professor Gregory Hillhouse in the field of organometallic chemistry, and received a

Master of Science in 2009. She spent several years as a high school chemistry teacher and

as an undergraduate laboratory manager, and then entered the doctoral program at Brown

University in the fall of 2013. She joined the laboratory of Professor Wesley Bernskoetter,

where she has studied the activation and transformation of carbon dioxide and methanol

using complexes containing earth-abundant metals like nickel and iron.

Page 5: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

v

Acknowledgements

I feel very fortunate to have completed my graduate studies in the chemistry

department at Brown. The dedication and enthusiasm with which the faculty and staff

approach their duties is apparent, and I will not forget the sense of inclusion that I

experienced when working here. It is definitely an environment that fosters collaboration

between disciplines that is difficult to find elsewhere.

I would first very much like to thank my boss, Wes, for accepting me into his lab

and letting me to pretend like I owned it for five years. His deep understanding of chemistry

is something that I have always admired, and the clarity and patience with which he

conveys that knowledge to others is a rare skill. I cannot fully express how much I value

the time I have spent working for him. His lab has been a home to me, and for that I will

always be grateful.

I have also been gifted a second amazing teacher in our collaborator, Professor

Nilay Hazari, who I would like to thank for his insightful discussions as well as his frequent

encouragement. In addition, I would like to extend my sincere thanks to my committee

members, Professor Eunsuk Kim and Professor Paul Williard, for their years of advice and

their continued support.

I would also like to thank my labmates, a term which I deem to be inclusive for life.

Both the lab group I entered into as well as the one I’m leaving were full of intelligent,

hardworking, fun people who made the hard days bearable and the good days remarkable.

Thank you all, so much, for getting me through this. A special thanks to Alex, Ryan, and

Kevin, who have been like brothers to me, and to Yuanyuan, Ariana, Zoe, Melissa, and

Katie for their friendship and for many days spent commiserating.

Page 6: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

vi

A final but equally important group of people to acknowledge is my family. Thank

you Mom, Dad, and Kate, for everything that you did to help get me here. Your strength,

intelligence, and kindness are what I aspired toward when I was growing up. You supported

me, loved me, and moved me back and forth across the country more times than can be

rationally expected of any one group of people. I absolutely could not have done this

without you.

Page 7: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

vii

Table of Contents

Vita…………………………………………………………………………………iv

Acknowledgements………………………………………………………………...v-vi

List of Figures……………………………………………………………………...x-xi

List of Schemes…………………………………………………………………….xii-xiv

List of Tables……………………………………………………………………….xv

List of Abbreviations……………………………………………………………….xvi

General Introduction: The Significant Role of Homogeneous Catalysts in the Field of Small

Molecule Activation………………………………………………………………...1-6

Chapter 1: The Application of Homogeneous Nickel and Iron Complexes in the

Transformation of CO2 and CO2 Analogues into Valuable Organic Products……...7

1.1: Introduction: Carbon Dioxide as a Carbon Feedstock………………….8-16

1.2: Transition-Metal-Catalyzed Oxidative Coupling of CO2 and CO2 Analogues

with Olefins………………………………………………………………….17

1.2.1: Introduction………………………………………………..….17-27

1.2.A.i.: Nickel-Promoted CO2-Alkene Coupling for the Stereoselective

Production of Substituted Cyclic Anhydrides……………………….28

1.2.A.i.1: Introduction……………………………………….28-33

1.2.A.i.2: Results and Discussion…………………………...34-40

1.2.A.i.3: Summary……………………………………….…41

Page 8: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

viii

1.2.A.ii.: The Use of CO Surrogates in Nickel-Promoted CO2-Alkene

Coupling for the Catalytic Production of Cyclic Anhydrides….…..42

1.2.A.ii.1: Introduction………………………………….….42-43

1.2.A.ii.2: Results and Discussion……………………….…44-47

1.2.A.ii.3: Summary…………………………………….…..48-49

1.2.B.: Nickel-Promoted Coupling of Isocyanates and Alkenes for the

Production of N-Substituted Acrylamides………………………….50

1.2.B.1: Introduction………………………………………..50-55

1.2B.2: Results and Discussion…………………………….55-65

1.2.B.3: Summary…………………………………………..65-66

1.2.C.: Section 1.2 Summary……………………………………….66-67

1.3: Transition-Metal-Catalyzed Hydrogenation of CO2 to Methanol…68

1.3.1: Introduction……………………………………….…..68-79

1.3.2: Results and Discussion……………………………….79-86

1.3.3: Summary……………………………………………...86-87

1.4: Chapter Summary………………………………………………88-89

1.5: References……………………………………………………....90-107

Chapter 2: The Application of an Iron Catalyst in the Transformation of Methanol into

Formamides and Ureas via Acceportless Dehydrogenation Reactions……………...108

Page 9: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

ix

2.1: Introduction: Methanol as a Carbon Feedstock……………………….109-113

2.2: Iron-Catalyzed Amide Formation from the Dehydrogenative Coupling of

Alcohols and Secondary Amines…………………………………………..114

2.2.1: Introduction……………………………………………….…114-119

2.2.2.: Results and Discussion………………………………….......119-127

Catalytic Studies

Mechanistic Investigations

2.2.3: Summary…………………………………………………….127-128

2.3: Iron-Catalyzed Urea Synthesis: Dehydrogenative Coupling of Methanol and

Amines………………………………………………………………..……129

2.3.1: Introduction…………………………………………………129-133

2.3.2: Results and Discussion…………………………………...…133-142

Catalytic Studies

Mechanistic Investigations

2.3.3: Summary…………………………………………………….142-143

2.4: Chapter Summary……………………………………………………..144-145

2.5: References……………………………………………………………..146-155

Appendix 1…………………………………………………………………………156-228

Appendix 2…………………………………………………………………………229-272

Page 10: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

x

List of Figures

Figure 1.1.1: CO2 reduction potentials at pH 7 (E° vs. NHE)……………………….…13

Figure 1.1.2: Metal binding modes of CO2………………………………………….....14

Figure 1.2.1: Depiction of β-H orientation in a nickelalactone complex and corresponding

d-orbital splitting diagram……………………………………………………………....24

Figure 1.2.A.1: Examples of [N,N], [P,N], and [P,P] types of bidentate ligands. tBu = tert-

butyl, Cy = cyclohexyl…………………………………………………………………..32

Figure 1.2.A.2: Structure key for ligand numbering. iPr = isopropyl, tBu = tert-butyl....34

Figure 1.2.A.3: Chiral nickel complexes competent for coupling propylene, CO2, and CO

to give MSA. iPr = isopropyl, tBu = tert-butyl………………………………………….40

Figure 1.2.A.4: CO surrogates and bases used, with abbreviations……………………44

Figure 1.2.B.1: 31P{1H} NMR spectrum in C6D6 of the nickel-promoted coupling products

from the ethylene/ethyl isocyanate reaction (2 atm/10 equiv.). Products C/D do not

incorporate ethylene. 300 MHz spectrometer…………………………………………...58

Figure 1.2.B.2: 31P{1H} NMR spectrum in C6D6 of the nickel-promoted coupling products

from the ethylene/tert-butyl isocyanate reaction (2 atm/10 equiv.) See Figure 1.2.B.1 for

labeling guide. 300 MHz spectrometer………………………………………………….60

Figure 1.2.B.3: 31P{1H} NMR spectrum in C6D6 of the nickel-promoted coupling products

from the ethylene/phenyl isocyanate reaction (1 atm/2 equiv.). See Figure 1.2.B.1 for

labeling guide. 300 MHz spectrometer…………………………………………………..62

Figure 1.2.B.4: Alternative ligands tested in Ligand-Ni(COD)-promoted

ethylene/isocyanate coupling reactions………………………………………………......64

Page 11: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

xi

Figure 1.3.1: Previous (1-5) and current metal catalysts for CO2 hydrogenation to

methanol via CO2 derivatives. OTf = triflate, OAc = acetate…………………………..77

Figure 1.3.2: Structures of the iron-pincer complexes 1, 2, and 1-formate…………...79

Figure 2.2.1: A top-performing ruthenium catalyst (a) and all previous base-metal catalysts

(b) for dehydrogenative amide formation from amines and alcohols…………………. 116

Figure 2.2.2: Efficient ruthenium, manganese, and iron catalysts for the dehydrogenative

coupling of alcohols and 2° amines……………………………………………………..118

Figure 2.2.3: Variable temperature 31P NMR spectra indicate an equilibrium between 1

and 2 in the presence of 6 equiv. of methanol. Sample dissolved in benzene-d6……….126

Figure 2.3.1: Ruthenium and iron catalysts for symmetric and unsymmetric urea formation

from the dehydrogenative coupling of methanol and amines…………………………..132

Page 12: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

xii

List of Schemes

Scheme 1.1.1: Industrial-scale syntheses incorporating CO2 as a carbon source…….….10

Scheme 1.1.2: Select C-O, C-N, C-C, and C-H bond-forming reactions with CO2. Products

in blue are industrially produced……………………………………………………...…..15

Scheme 1.2.1: Generic metal-catalyzed oxidative coupling between CO2 and an

unsaturated hydrocarbon (without a formally metal-bound CO2)……………………….18

Scheme 1.2.2: Commercial acrylic acid production…………………………………….21

Scheme 1.2.3: Proposed catalytic cycle for the transition-metal-catalyzed production of

acrylic acid from ethylene and CO2……………………………………………………..22

Scheme 1.2.4: Examples of nickel- and molybdenum-promoted oxidative coupling

reactions between ethylene and CO2…………………………………………………....23

Scheme 1.2.5: Methods of inducing acrylate removal from nickelalactones using (A) base

or (B) Lewis acids (LA)………………………………………………………………...26

Scheme 1.2.A.1: Proposed mechanism for the nickel-catalyzed production of cyclic

anhydrides from CO2, alkenes, and CO surrogates…………………………………….28

Scheme 1.2.A.2: Most prevalent methods of cyclic anhydride production…………... 30

Scheme 1.2.A.3: (A) Achiral and (B) chiral nickel-promoted methods for the production

of cyclic anhydrides…………………………………………………………………....31

Scheme 1.2.A.4: Methods of nickelalactone production from (A) CO2/alkene coupling and

retrosynthetically from (B) a cyclic anhydride or (C) acrylic acid…………………….35

Scheme 1.2.A.5: Lewis acid enhancement of the 3-,4-, and 5-Ni(COD) coupling reactions

between ethylene and CO2……………………………………………………………..38

Page 13: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

xiii

Scheme 1.2.A.6: Previous application of CO surrogates with nickel species for the

production of cyclic lactams and the current proposed use for releasing cyclic

anhydrides…………………………………………………………………………..…...43

Scheme 1.2.B.1: Large-scale methods of synthesizing acrylamides…………………...51

Scheme 1.2.B.2: Nickel-catalyzed isocyanate/alkene coupling reactions with conditions

for accessing specific acrylamide regioisomers as well as propionamides and

succinimides…………………………………………………………………………….54

Scheme 1.2.B.3: Nickel-catalyzed isocyanate/alkene coupling reactions pursued in this

work…………………………………………………………………………………….55

Scheme 1.3.1: Products of CO2 hydrogenation reaction and the Gibbs free energy for

formic acid and methanol generation…………………………………………………..68

Scheme 1.3.2: Industrial method of methanol production from syngas, depicting (1) the

component equilibria and (2) the overall reaction……………………………………..69

Scheme 1.3.3: General method of H2 storage using reversible metal-catalyzed CO2

hydrogenation/dehydrogenation………………………………………………………..71

Scheme 1.3.4: Methods of indirect hydrogenation of CO2 to methanol……………….73

Scheme 1.3.5: Metal-catalyzed CO2 hydrogenation to methanol via (1) esters or (2)

formamides……………………………………………………………………………...74

Scheme 1.3.6: Flow diagram for the one-pot, two-step reaction setup………………....81

Scheme 2.1.1: (1) Industrial methanol production from syngas and (2) industrial

formaldehyde production from methanol……………………………………………….109

Scheme 2.1.2: Some of the major chemical species derived from methanol and their

uses……………………………………………………………………………………...110

Page 14: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

xiv

Scheme 2.2.1: Metal-free methods for amide production……………………………...114

Scheme 2.2.2: Metal-promoted pathways for dehydrogenative amidation…………….115

Scheme 2.2.3: Iron-catalyzed coupling of benzaldehyde or methyl formate with

morpholine……………………………………………………………………………..125

Scheme 2.3.1: Metal-free pathways for urea formation……………………………….129

Scheme 2.3.2: Metal-catalyzed urea synthesis from non-methanol sources…………..130

Scheme 2.3.3: Metal-promoted pathway for dehydrogenative urea formation from

methanol and amines……………………………………………………………….…..131

Scheme 2.3.4: Potential scrambling pathways in unsymmetric urea formation……….138

Scheme 2.3.5: Proposed mechanistic pathways for metal-catalyzed urea formation starting

from methanol and primary amines……………………………… …………………..141

Scheme 2.3.6: Formamide-dependent synthesis of 1,1,3-tripentylurea……………..…141

Page 15: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

xv

List of Tables

Table 1.2.A.1: Reaction of dcpe-Ni-lactone and dcpe-Ni-lactone’ Species with CO

Surrogates……………………………………………………………………………..46

Table 1.3.1: Select Reaction Optimization Data for CO2 Hydrogenation to Methanol in the

Presence of Morpholine……………………………………………………………….84

Table 2.2.1: Dehydrogenative Amidation of Methanol Catalyzed by 1……………...121

Table 2.2.2: Alcohol Dehydrogenation in the Presence of Morpholine Catalyzed by

1………………………………………………………………………………………..123

Table 2.3.1: Dehydrogenative Symmetric Urea Formation from Methanol and Amines

Catalyzed by 1………………………………………………………………………....134

Table 2.3.2: Dehydrogenative Unsymmetric Urea Formation from Formamides and

Amines Catalyzed by 1………………………………………………………………...137

Page 16: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

xvi

List of Abbreviations

iPr = isopropyl LA = Lewis acid

tBu = tert-butyl SA = succinic anhydride

Cy = cyclohexyl MSA = methylsuccinic anhydride

Ph = phenyl NaBAr4F = sodium tetrakis

Et = ethyl [3,5-bis(trifluoromethyl)phenyl]borate

Me = methyl DBU = 1,8-diazabicyclo[5.4.0]undec-

CO2 = carbon dioxide 7-ene

CO = carbon monoxide NaOtBu = sodium tert-butoxide

H2 = dihydrogen LiOTf = lithium

COD = 1,5-cyclooctadiene trifluoromethanesulfonate

cdt = 1,5,9-cyclododecatriene Complex 1 = (iPrPNP)Fe(H)(CO)

dcpe = 1,2-bis(dicyclohexylphosphino)ethane iPrPNP = N[CH2CH2(PiPr2)]2

TON = turnover number

TOF = turnover frequency

Mt = million metric tons

kt = kilotons

ppm = parts per million

syngas = synthesis gas

[CO] = CO surrogate

[B] = base

NMR = nuclear magnetic resonance

GCMS = gas chromatography-mass spectrometry, FID = flame ionization detector

Page 17: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

1

General Introduction:

The Significant Role of Homogeneous Catalysts in the Field of Small Molecule

Activation

Page 18: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

2

The overwhelming dependence on fossil fuels for supplying the world’s energy

needs is unsustainable, and the rapid depletion of these non-renewable resources has

motivated investigations into alternative means of meeting global demands.1-3 Small

molecules such as dioxygen (O2), dinitrogen (N2), dihydrogen (H2), carbon dioxide (CO2),

and methane (CH4) appear in various biological metabolic processes and large-scale

industrial conversions, exhibiting their participation in the efficient and reversible storage

of energy in the form of chemical bonds.4 This combined with their abundance makes them

attractive targets for the development of alternative chemical feedstocks and fuels.4-7,9,13

Reactions of interest include the conversion of N2 to ammonia or hydrazine,8,9 the

reversible (de)hydrogenation of CO2 as a means of H2 storage for alternative fuels,10 H2

oxidation/O2 reduction coupled fuel cells,11 and CH4 (alkane) functionalization to higher

chemicals.12 However, a major obstacle to realizing these objectives is the thermodynamic

and often also kinetic stability of these materials.1,13 Those compounds that are more

amenable kinetically often suffer from a lack of product control instead (e.g. over-oxidizing

a substrate using O2).13 Addressing these problems is where homogeneous catalysts are

poised to contribute significantly to the advancement of the field of small molecule

activation. In addition to alleviating kinetic barriers, catalysts can enable otherwise

unavailable interactions between the more inert species listed above and highly reactive

molecules to yield a product with a more favorable thermodynamic profile.14-16 They can

also provide a means of controlling reactivity.13 It should be noted that currently the most

active catalysts for enacting these types of transformations are heterogeneous.8,17 Areas

where homogeneous catalysts gain an advantage are in enhanced product selectivity and a

greater mechanistic understanding that allows for the rational tuning of catalytic

Page 19: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

3

behavior.8,18-20 The latter characteristic is especially crucial when developing catalytic

systems for new processes. A host of homogeneous catalysts have therefore been employed

in the task of facilitating small molecule transformations into value added products and

fuels. The application of catalysts containing inexpensive and earth-abundant metals

toward achieving this goal is of particular importance from both a sustainability and an

economic viewpoint.21 The work presented here details several such examples, using

homogeneous nickel and iron complexes for the functionalization of CO2 and methanol,

with the ultimate goal of catalyzing the production of commercially valuable chemicals

from cheap, widely available, and potentially renewable C1 carbon sources.

Page 20: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

4

A note concerning catalyst comparisons: When considering catalytic performances, terms

such as conversion, yield, turnover frequency (TOF), and turnover number (TON) are often

used for comparisons between catalytic species. Conversion usually refers to the

consumption of one of the starting materials as a way of assessing reaction progress, but

this can occasionally be misleading if the selectivity of the reaction is poor. Yield is

determined based on the amount of the desired product in comparison to the theoretical

maximum and therefore avoids selectivity as a factor. While yield is a useful measure of

synthetic practicality, it often fails to express the full activity of a catalyst due to the ceiling

created by the limiting reagent(s). TOF explicitly references the performance of the catalyst

in question, analyzing how many molecules of product are generated per molecule of

catalyst over a given time period. The TOF, like most reaction rates, will change over the

course of catalysis, and is typically reported as the ability of a catalyst to deliver some

number of moles of product in one hour. This can be a valuable method to compare the

nativity of catalysis, but care must be taken to compare catalysts using TOFs determined

over the same reaction time periods. TON is a measurement of the total amount of product

delivered over the whole of the reaction per unit of catalyst. Several factors impact the

maximum TON of a catalyst, including the TOF, reaction selectivity, and catalyst lifetime.

When catalysis is conducted in a large excess of substrate, the maximum TON represents

the greatest number of times a molecule of catalyst can perform the desired transformation

before it becomes inactive. When developing new catalysts, we have selected TON as the

primary point of comparison because it is a value that cannot be easily skewed by changing

experimental methods and reflects several aspects important to catalysis in a single value.

Page 21: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

5

References

1 Olah, G.A.; Goeppert, A.; Prakash, G.K.S. “Chemical Recycling of Carbon Dioxide to Methanol and

Dimethyl Ether: From Greenhouse Gas to Renewable, Environmentally Carbon Neutral Fuels and Synthetic

Hydrocarbons,” J. Org. Chem., 2009, 74, 487-498.

2 Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.;

Hille, R.; Kenis, P.J.A.; Kerfeld, C.A.; Morris, R.H.; Peden, C.H.F.; Portis, A.R.; Ragsdale, S.W.; Rauchfuss,

T.B.; Reek, J.N.H.; Seefeldt, L.C.; Thauer, R.K.; Waldrop, G.L. “Frontiers, Opportunities, and Challenges

in Biochemical and Chemical Catalysis of CO2 Fixation,” Chem. Rev., 2013, 113, 6621-6658.

3 Goeppert, A.; Czaun, M.; Jones, J.-P.; Prakash, G.K.S.; Olah, G. “Recycling of carbon dioxide to methanol

and derived products- closing the loop,” Chem. Soc. Rev., 2014, 43, 7995-8048.

4 Meyer, F.; Tolman, W.B. “Forums on Small-Molecule Activation: from Biological Principles to Energy

Applications,” Inorg. Chem., 2015, 54, 5039-5039.

5 Lehnert, N.; Peters, .J.C. “Preface for Small-Molecule Activation: from Biological Principles to Energy

Applications. Part 2: Small Molecules Related to the Global Nitrogen Cycle,” Inorg. Chem., 2015, 54, 9229-

9233.

6 Darensbourg, M.Y.; Llobet, A. “Preface for Small-Molecule Activation: from Biological Principles to

Energy Applications. Part 3: Small Molecules Related to (Artificial) Photosynthesis,” Inorg. Chem., 2016,

55, 371-377.

7 Fujita, E.; Goldman, A.S. “Preface for Small-Molecule Activation: Carbon-Containing Fuels,” Inorg.

Chem., 2015, 54, 5040-5042.

8 Nishibayashi, Y. “Recent Progress in Transition-Metal-Catalyzed Reduction of Molecular Dinitrogen under

Ambient Reaction Conditions,” Inorg. Chem., 2015, 54, 9234-9247.

9 Bhattacharya, P.; Prokopchuk, D.E.; Mock, M.T. “Exploring the role of pendant amines in transition metal

complexes for the reduction of N2 to hydrazine and ammonia,” Coord. Chem. Rev., 2017, 334, 67-83.

10 Wang, W.-H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. “CO2 Hydrogenation to Formate

and Methanol as an Alternative to Photo- and Electrochemical CO2 Reduction,” Chem. Rev., 2015, 115,

12936-12973.

11 Zhang, W.; Lai, W.; Cao, R. “Energy-Related Small Molecule Activation Reactions: Oxygen Reduction

and Hydrogen and Oxygen Evolution Reactions Catalyzed by Porphyrin- and Corrole-Based Systems,”

Chem. Rev., 2017, 117, 3717-3797.

12 Munz, D.; Strassner, T. “Alkane C-H Functionalization and Oxidation with Molecular Oxygen,” Inorg.

Chem., 2015, 54, 5043-5052.

13 Zimmermann, P.; Limberg, C. “Activation of Small Molecules at Nickel(I) Moieties,” J. Am. Chem. Soc.,

2017, 139, 4233-4242.

14 Huang, K.; Sun, C.-L.; Shi, Z.-J. “Transition-metal-catalyzed C-C bond formation through the fixation of

carbon dioxide,” Chem. Soc. Rev., 2011, 40, 2435-2452.

Page 22: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

6

15 Pearson, R.L.; Eisaman, M.D.; Turner, J.W.G.; Edwards, P.P.; Jiang, Z.; Kuznetsov, V.L.; Littau, K.A.;

Di Marco, L.; Taylor, S.R.G. “Energy Storage via Carbon-Neutral Fuels Made from CO2, Water, and

Renewable Energy,” Proceedings of the IEEE, 2012, 100, 440-460.

16 Boddien, A.; Gärtner, F.; Federsel, C.; Piras, I.; Junge, H.; Jackstell, R.; Beller, M., Catalytic Utilization

of Carbon Dioxide: Actual Status and Perspectives, in Organic Chemistry: Breakthroughs and Perspectives;

Ding, K., Dai, L.-X., Eds.; Wiley-VCH: Weinheim, Germany, 2012, pp. 685-724.

17 Burford, R.J.; Yeo, A.; Fryzuk, M.D. “Dinitrogen activation by group 4 and group 5 metal complexes

supported by phosphine-amido containing ligand manifolds,” Coord. Chem. Rev., 2017, 334, 84-99.

18 Huff, C.A.; Sanford, M.S. “Cascade Catalysis for the Homogeneous Hydrogenation of CO2 to Methanol,”

J. Am. Chem. Soc., 2011, 133, 18122-18125.

19 Wang, W.; Wang, S.; Ma, X.; Gong, J. “Recent advances in catalytic hydrogenation of carbon dioxide,”

Chem. Soc. Rev., 2011, 40, 3703-3727.

20 Gamba, I.; Codolà, Z.; Lloret-Fillol, J.; Costas, M. “Making and breaking of the O-O bond at iron

complexes,” Coord. Chem. Rev., 2017, 334, 2-24.

21 Kärkäs, M.D.; Åkermark, B. “Water oxidation using earth-abundant transition metal catalysts:

opportunities and challenges,” Dalton Trans., 2016, 45, 14421-14461.

Page 23: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

7

Chapter 1:

The Application of Homogeneous Nickel and Iron Complexes in the Transformation of

CO2 and CO2 Analogues into Valuable Organic Products

Page 24: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

8

1.1 Introduction: Carbon Dioxide as a Carbon Feedstock

The most recent Renewables Global Status Report indicates that in 2015, 78.4% of

global energy consumption stemmed from fossil fuels, with renewable sources (such as

wind, solar, geothermal, and biomass) accounting for only 19.3%, and the remaining 2.3%

belonging to nuclear power.1 This worldwide dependence on non-renewable energy

resources has not only had significant deleterious environmental effects,2 but faces an

inevitable future where these finite resources are gone and will not be realized again on a

human timescale.3,4 One consequence of fossil-fuel dependence is that the cost of these

resources will increase exponentially as their availability plummets.4 Another is the ever-

increasing production of carbon dioxide (CO2) due to the burning of fossil fuels, the rising

atmospheric levels of which have been implicated in widespread destructive environmental

alterations, including climate change and ocean acidification.4,5 Atmospheric CO2 content

is dramatically higher than its pre-industrial level (from 250 ppm to 410 ppm in 2017),6,7

is increasing at a rate of ~1% per year,5 and despite efforts to decrease the contribution

from energy consumption,8 the natural forms of CO2 sequestration such as plant

photosynthesis and aqueous sinks5 cannot keep up. There are therefore significant

economic and environmental motivations to transition to renewable methods of meeting

the global fuel demand for energy, transportation, and commodity chemical applications.

One such method for partially relieving CO2 concentrations while decreasing fossil fuel

dependence is to utilize CO2 as a carbon feedstock in the production of alternative fuels

and commercial chemicals.6

CO2 is an attractive C1 carbon source for organic synthesis because it is cheap,

abundant, renewable, and non-toxic (in comparison to other sources like phosgene or

Page 25: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

9

carbon monoxide (CO)).6,9-13 As opposed to techniques aimed only at carbon capture and

storage, carbon recycling methods both remove CO2 from the atmosphere and transform it

into practically useful species while avoiding issues of high transportation costs to storage

sites and dependability of storage time periods.14,15 It is also the only method that generates

a profit from what is otherwise a waste item.14 While the utilization of CO2 for fuels and

commodity chemical production is unlikely to significantly counteract anthropogenic

emissions,10 it is an economically beneficial way of sequestering CO2 for multiple

lifetimes. By replacing organic precursors typically derived from non-renewable sources

with inexpensive CO2, the energy and cost of chemical production is reduced (with a

corresponding price reduction for the consumer) while also decreasing fossil fuel

dependency.10,15-18 If this synthesis of functionalized desirable species from CO2 is

accomplished using energy sources like solar, wind, or geothermal, it serves as a possible

avenue for the storage and transportation of renewable energy in chemical form, without

necessitating a change in the energy infrastructure.15,18 Alternatively, CO2-derived fuels

themselves can be viewed as renewable, carbon-neutral energy sources,3,15,18 so their

synthesis and subsequent use pulls current CO2 from circulation while also diverting from

future emissions by substituting for fossil fuels.3

Evidence for the feasibility of CO2 conversion is found in biological systems, which

over billions of years have evolved enzymatic pathways for processing CO2 (via the

making of C-H/C-C bonds and the breaking of C-O bonds) and thereby effectively storing

and utilizing energy in the form of chemical bonds.4 Despite the aforementioned

motivations for using CO2 in the production of high-value chemicals as well as the plethora

of examples of its application in living systems, the actual large-scale use of CO2 as a

Page 26: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

10

carbon source is quite limited. Currently, only ~1% of CO2 produced is used for chemical

synthesis,5,10 and the few CO2-fixation products manufactured on an industrial scale are

carbonates (inorganic, cyclic, and polymeric), urea, salicylic acid, and (to a lesser extent)

methanol (Scheme 1.1.1).6,10

Carbonates:

Inorganic carbonates have applications in the production of paper, plastics, rubber,

paint, and building materials, and calcium carbonate (CaCO3) alone accounts for 15 Mt

Scheme 1.1.1: Industrial-scale syntheses incorporating CO2 as a carbon source.

Page 27: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

11

(million metric tons) of the overall global market of more than 200 Mt per year (not entirely

derived from industrial CO2 sources).15,19 Therefore, the use of industrial residues rich in

magnesium, calcium, or sodium in conjunction with emitted CO2 to give carbonates has

the potential to be very lucrative. In fact, Skymine has received over $25 million in funding

to apply a patented CO2 mineralization process by which it converts flue gas CO2 into

sodium bicarbonate.15 Cyclic carbonates are made commercially from epoxides and CO2

(Scheme 1.1.1, reaction 1),20 but methods of replacing the pre-synthesized epoxides with

the direct reaction of alkenes and CO2 have been developed.21 Their value is largely based

on their role in the production of linear carbonates and as monomers for polycarbonates,

the combined market for which amounts to approximately 20 Mt per year. Uses include

resins for data storage, organic reagents for alkylation and acylation reactions, solvents

(e.g. for pharmaceuticals preparations), and battery electrolytes.10,14,15,20,22

Urea

The commercial production of urea from ammonia and CO2 via a carbamate

intermediate (Scheme 1.1.1, reaction 2) is known as the Bosch-Meiser urea process and

has been in use since the 1920s.15,19 It is conducted under elevated temperature (~160 °C)

and pressure (~110 atm) conditions, where the exothermic first step drives the endothermic

second step. The >150 Mt per year demand for urea represents the largest-volume industrial

example of converting CO2 to a value-added product, with its primary uses being as a

fertilizer and resin precursor.15 The use of homogeneous catalysts with substituted amines

and CO2 has yielded functionalized ureas but has not yet been realized on a commercial

scale (see Chapter 2, section 2 for more information).

Salicylic Acid

Page 28: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

12

The synthesis of salicylic acid was developed in the late 1800’s and is shown in

Scheme 1.1.1, reaction 3 (o-isomer is shown but the reaction also produces the p-isomer).

It is also known as the Kolbe-Schmitt reaction and is performed at 120-130 °C and ~100

atm CO2.23 Salicylic acid is used as a food preservative and in a variety of pharmaceuticals,

including those for the treatment of dermatological conditions (acne, psoriasis, ringworm,

dandruff) and for pain relief. The latter accounts for its most widespread application, which

is in the production of aspirin on a 20 kt (kilotons) per year scale.19,24

Methanol

Methanol is produced on a more than 80 Mt industrial scale from a synthesis gas

mixture (syngas; H2/CO2/CO) using heterogeneous alumina-supported mixed copper and

zinc oxide catalysts at elevated temperatures and pressures (50-100 atm, 200-300 °C;

Scheme 1.1.1, reaction 4).25-28 While a small percentage of manufactured methanol is used

as a solvent, most serves as a reagent in the synthesis of a variety of valuable end-products

like textiles, plastics, film, paints, adhesives, and coatings.29 This topic is covered in greater

detail in the last section of this chapter as well as in Chapter 2.

These limited examples offer the possibility of a renewable CO2-based carbon

economy, however, significant obstacles to the more widespread utilization of CO2 as a

carbon building block in the synthesis of organic molecules are its kinetic and

thermodynamic stabilities.5,6,18,30 Containing the most oxidized state of carbon,9 CO2 is

highly unreactive,6 making it a poor ligand and difficult to transform via direct reduction.31

Despite these complications, significant effort has been invested in achieving the

electrochemical reduction of CO2 to CO, methane, formic acid, or methanol.44 The appeal

stems not only from the recycling of CO2 and the production of valuable compounds, but

Page 29: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

13

also from the storage of excess electrical energy in chemical form, as their energy density

can be anywhere from 10 to 100 times that of a battery.19,32 A one-electron reduction of

CO2 is highly unfavorable, and as a result proton-coupled multi-electron steps to yield the

products mentioned above are preferred (Figure 1.1.1).32,44 The electrocatalytic reduction

of CO2 has been studied since the 1980’s with either solid metal electrodes33 or

homogeneous transition metal complexes,32,34-37,44 however, there are significant

challenges with respect to low activities and the generation of side products like CO and

H2.32 Ideal catalysts employed in this process would increase overall activity and

selectivity, but would balance these kinetic advantages with a good match for the redox

potential of the electron-transfer reaction.32,44 These catalysts must also be capable of two-

electron transfer processes and be water stable (due to its use as a proton source). During

the electrochemical reduction reaction, several reductive equivalents build up at the

catalyst site, which then uses multi-electron pathways (hydride transfer, atom transfer,

insertion) to transform CO2.32 The most successful catalysts in this respect are iridium-

based,32,38,39 however, they (like other catalysts for this process) suffer from a trade-off

between the coveted traits of high activity, high selectivity, and low overpotential,32 and

usually sacrifice one for the other two. Photochemical reduction systems face similar

Figure 1.1.1: CO2 reduction potentials at pH 7 (E° vs. NHE).

Page 30: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

14

issues, along with the added challenge of devising ways of enacting a multi-electron

reaction from a single photon, single-electron-transfer process.236 This in combination with

the incompatibility of the reductive conditions with many desirable products has motivated

the investigation of alternative means for CO2 transformations.

Although the energetic barriers for CO2 activation are significant, metal-mediated

non-electrochemical strategies for its incorporation into organic syntheses are promising

targets. A weakly Lewis acidic carbon and weakly Lewis basic oxygens due to bond

polarity make it possible for CO2 to bind to a metal center,4,10 and subsequent reduction

and transformation is much more favorable.42 However, not all metals have the basicity

necessary to accomplish this and metal-CO2 complexes are often unstable. Therefore, while

multiple examples of these complexes exist (the various binding modes of which are

depicted in Figure 1.1.2),10,17,40-43 it would be advantageous to use metal systems where

CO2 pre-coordination is not required for reaction but where participation in coupling and

insertion reactions is still possible. Current strategies aimed at CO2 activation combine CO2

with highly reactive starting materials (such as epoxides, dihydrogen, or unsaturated

compounds) to yield target species that are much lower in energy, thereby providing a

thermodynamic driving force and achieving a much more favorable thermodynamic

profile.9,18,44 The addition of a transition metal catalyst to mitigate unfavorable kinetic

Figure 1.1.2: Metal binding modes of CO2.

Page 31: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

15

parameters5 that can also act on CO2 without it being formally metal-bound45 would

address any further resistance to functionalization.

Despite its chemical inertness, the slightly electrophilic C allows for the reaction of

CO2 with strongly nucleophilic reagents to give new C-H, C-O, C-N, and C-C bonds

(Scheme 1.1.2).30,44 Practical examples that rely on this principle in the large-scale

production of organic compounds are the previously mentioned cases of cyclic and

polymeric carbonates, urea, and salicylic acid.30,44 While reaction with energy-rich

substrates addresses some of the thermodynamic concerns surrounding CO2 utilization,

kinetic challenges can still remain.44 In order to alleviate these problems, the harsh reaction

conditions (high temperatures and pressures) demonstrated in the industrial examples

above and/or the use of metal catalysts are required. For example, even though the metal-

free production of cyclic carbonates has been commercialized for over 50 years, recent

investigations into the use of metal porphyrin catalysts containing aluminum, magnesium,

Scheme 1.1.2: Select C-O, C-N, C-C, and C-H bond-forming reactions with CO2.

Products in blue are industrially produced.

Page 32: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

16

cobalt, chromium, iron, and zinc have significantly increased the reaction turnover under

milder conditions.30 The metal-free and metal-catalyzed manufacturing of new C-O and C-

N bonds using CO2 as a carbon feedstock has been extensively covered by several excellent

reviews and a book, and are beyond the scope of this discussion. They similarly cover

metal-free C-C bond formation using organolithium, organozinc, and Grignard reagents as

highly reactive and strong nucleophiles for the activation of CO2, as well as the metal-free

and metal-catalyzed reaction of less nucleophilic organoboron and organohalide

compounds with CO2, all giving carboxylic acids and their derivatives.9,30,44,46,47 While

these are diverse and effective ways of generating new C-C bonds incorporating CO2, they

are often highly intolerant of other functional groups. This chapter will instead focus on

the formation of new C-C and C-H bonds (inherently more difficult that C-N or C-O due

to decreased reagent nucleophilicity)30 using two metal-catalyzed CO2 activation methods:

the oxidative coupling of CO2 with olefins and the hydrogenation of CO2. These methods

for the transformation of CO2 (and CO2 analogues) will be discussed in the context of the

valuable chemicals they can be used to produce: substituted cyclic anhydrides and N-

substituted acrylamides with the former, and methanol with the latter.

Page 33: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

17

1.2: Transition-Metal-Catalyzed Oxidative Coupling of CO2 and CO2 Analogues with

Olefins

1.2.1: Introduction

The first reported case of a metal-coordinated CO2 species by Aresta et al. in 1975

motivated extensive research focusing on the metal-catalyzed activation and

transformation of CO2.48 One such method involves oxidatively coupling CO2 (which is

highly stable) with reactive small molecules like alkenes or alkynes at a transition metal

center, whereby conversion to valuable organic molecules becomes more

thermodynamically favorable.49 The metal-promoted reaction of CO2 with unsaturated

hydrocarbons has been studied since the 1970’s, with pioneering work by Hoberg, Walther,

Inoue, and Yamamoto. Nearly all examples rely on the use of late, electron-rich transition

metals (d8-d10: Fe0, RhI, Ni0, Pd0 or PdII) with low oxidation states because they have

orbitals available for ligand backbonding, they can easily achieve higher oxidation states

(as part of the coupling process), and because of their nucleophilic nature. This allows for

a σ-type bonding interaction between the metal and the slightly electrophilic carbon of

CO2.10 Computational work by Pápai et al. has suggested that CO2 pre-coordination is not

necessary for these reactions and that coupling can occur in a concerted fashion,45 although

experimental evidence in some cases has indicated distinctly bound CO2 and alkene ligands

at the metal center.50 In cases where CO2 does not act as a discreet ligand, metal

nuclophilicity is still important because it facilitates alkene and alkyne binding via metal

to ligand backbonding, thereby activating the unsaturated species for subsequent CO2

coupling to form a metallocycle (Scheme 1.2.1).

Page 34: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

18

Experimental investigation into the functionalization of CO2 using metal-promoted

coupling reactions can be assessed based on the CO2 co-reactant: alkenes, dienes, and

alkynes. Despite the diversity of substrates and over 30 years of research, there are very

few systems in which these transformations are catalytic.10 Some of the earliest work in

this area of CO2 activation was the palladium-catalyzed coupling reaction of CO2 and 1,3-

butadiene presented by Inoue et al. in 197651 and Musco et al. in 1978.52 These reactions,

however, yielded product mixtures where the desired carboxylic acid was a minor

contributor compared to multiple butadiene oligomers. These results are characteristic of

the behavior of dienes in general in this type of reaction since they tend to telomerize at

late transition metals, a trait which has been established for some time (e.g. the synthesis

of 1,5,9-cyclododecatriene, a.k.a. cdt, from butadiene and nickel metal).10,53-55 Some

progress in controlling product distribution has been made since the original publications

(such as by changing the metal starting material or using more sterically demanding

ligands),56,57 but it still remains the most significant challenge to using dienes as CO2

coupling partners. The metals capable of realizing this transformation are also quite

limited, and those that can do so catalytically primarily rely on palladium,51,52,56-60 with

only a few examples containing nickel46,61 and rhodium.62 The early work on the coupling

reaction between CO2 and alkynes faced similar selectivity issues with the formation of

dimers and trimers in addition to the desired pyrone product. Inoue et al. were also the first

Scheme 1.2.1: Generic metal-catalyzed oxidative coupling reaction between CO2 and

an unsaturated hydrocarbon (without a formally metal-bound CO2).

Page 35: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

19

to publish on this topic, reacting a nickel (0) complex [Ni(COD)2, where COD = 1,5-

cyclooctadiene] with CO2 and 1-hexyne to yield a mixture of cyclotrimerization products,

only 10% of which was the desired species.63 Further development of this field using

rhodium64 and nickel catalysts65 eventually gave 100% yield with 96% selectivity for the

pyrone product using a sterically small phosphine ligand (triethylphosphine).66 However,

these results represent a very limited substrate scope and therefore have not been widely

applied. They also conflict with later findings that the selectivity of reactions between CO2

and diynes is improved with sterically larger ligands like N-heterocyclic carbenes.67,68

Overall, methods for controlling product preference in metal-promoted CO2/alkyne

coupling reactions are still poorly understood. In addition, many examples require the use

of aqueous acid or carbon monoxide (CO) to release the final product, obviating

catalysis.68-74 As a result, while coupling CO2 with dienes and alkynes is relevant to

accessing a range of functionalized and desirable products (covered in detail in the cited

reviews),6,9,10 the results presented in this chapter will instead focus on coupling reactions

of CO2 with alkenes.

The earliest work demonstrating CO2/olefin coupling was published by Inoue et al.

in 1979, featuring a palladium complex (Pd[dba]2, dba = dibenzylideneacetone) for the

production of lactones from CO2 and methylenecyclopropanes, which are highly reactive

due to pronounced ring strain. A mixture of two lactone products was observed, and the

identity of the majority species could be controlled by alteration of the ligand environment

on palladium.75 The original work did not address catalytic activity, and while future

investigations by other groups did, issues of side reactions between the product lactones

and starting material still in solution detracted significantly from yield.76 Therefore, while

Page 36: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

20

these palladium results represent initial forays into the concept of metal-catalyzed coupling

of CO2 with a very restricted set of alkenes, it was investigations by the groups of Hoberg,

Walther, and Yamamoto in the 1980’s that established the scope of the field and identified

its most significant challenges.

Initial studies presented independently by Hoberg and Walther centered on the use

of nickel (0) sources containing weakly bound and easily displaced olefins like Ni(COD)2

and Ni(cdt) in the presence of bulky, electron-donating ligands like phosphines (e.g. 1,2-

bis(dicyclohexylphosphino)ethane = dcpe, triethylphosphine = PEt3) or bipyridine.77,82 The

larger ligand sizes discouraged binding multiple units to the metal center, which would

create catalytically inactive tetra-substituted species (NiL2 for bidentate ligands or NiL4 for

monodentate, both 18-electron species). The ligand σ-donating aspect contributed to the

overall nucleophilicity of the metal in order to promote metal-alkene binding and activation

toward further manipulation (via metal to ligand donation into the alkene antibonding

orbitals, weakening the carbon-carbon double bond) as well as to facilitate the necessary

increase in oxidation state upon CO2 coupling. These early works spanned alkenes as

simple as ethylene77-79 to sterically hindered species like norbornene,80 cyclopentene,81 and

dicyclopentadiene,82 as well as substituted alkenes78 and select linear dienes.83,84 Hoberg

et al. showed that under specific pressure conditions, nickel could facilitate the coupling

of multiple equivalents of an alkene (ethylene) with CO2.85 Hoberg’s group also extended

these concepts to include iron (0) complexes as metal promoters, some of which displayed

the unique characteristic of allowing multiple CO2 insertions to give diesters (after

methanolysis workup).79,84 It was demonstrated that while oxidative coupling of CO2 and

the selected alkene could easily proceed to the five-membered metallocycle in the presence

Page 37: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

21

of a metal (known as nickelalactones for M = Ni), spontaneous elimination of the desired

product was not observed. While this permitted the extensive study of diverse metallocycle

structures (sometimes via ligand substitution),86 product liberation was usually attained by

acid hydrolysis or methanolysis (to yield a carboxylic acid or an ester, respectively) or by

addition of CO gas (to yield a cyclic anhydride).77,78,83 These methods, however, prevented

the catalytic application of CO2/olefin coupling toward synthesizing organic compounds

through decomposition of the metal species or conversion to an inactive metal carbonyl

complex (e.g. L2Ni(CO)2).

These pioneering works motivated extensive studies into the oxidative coupling of

CO2 with olefins and alkynes. Of particular interest was the coupling reaction between CO2

and ethylene because of the potential to liberate acrylic acid, and from these investigations

the major challenges associated with enacting these types of CO2 transformations have

been established. The growing market for acrylic acid was over 4.92 Mt in 2014.87 Its worth

is mostly dependent on its role as a precursor to polyacrylates, which are used to make

plastics, adhesives, coatings, detergents, and super absorbent polymers (such as those in

diapers).88 Current acrylic acid production depends on a two-step oxidation over

heterogeneous catalysts and begins with propylene (Scheme 1.2.2).89-91 Not only does this

process require harsh temperatures (250-400 °C)89,91 and go through a highly toxic acrolein

intermediate,89 but propylene is non-renewably sourced from fossil fuels and displays

Scheme 1.2.2: Commercial acrylic acid production.

Page 38: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

22

higher costs as a result, which is highly correlated with the cost of acrylic acid.87 There is

therefore significant appeal in developing a method of acrylic acid production under milder

conditions and from cheaper, potentially renewable starting materials like ethylene and

CO2.10 A proposed catalytic cycle for this process is shown in Scheme 1.2.3. It begins with

binding the olefin to the metal, followed by oxidative coupling with CO2 to give the

metallocycle and then β-H elimination to the bound acrylate species. A final reductive

elimination step releases the product and re-opens a binding site for further catalysis.

Scheme 1.2.3: Proposed catalytic cycle for the transition-metal-catalyzed production

of acrylic acid from ethylene and CO2.

Page 39: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

23

The groups of Hoberg and Carmona independently pursued the goal of metal-

catalyzed acrylic acid/acrylate formation from CO2/ethylene coupling using zerovalent

nickel and group VI metals, respectively (Scheme 1.2.4).77,78,92-95 Experimental and

computational mechanistic studies indicated that these metal complexes likely shared

common intermediates but faced different obstacles with respect to achieving

catalysis.45,50,96-100 In the case of complexes containing group VI metals (e.g. Mo, W), the

oxidative coupling of CO2 and ethylene occurred easily at ambient temperatures and

pressure, and β-H elimination to give the bound acrylate species was similarly facile.50,92-

95,101 However, the oxophilicity of group VI metals prevented product release via reductive

elimination in the absence of harsh bases (like butylithium)93 or strong electrophiles (such

as iodomethane),101 which are incompatible with catalysis.

Scheme 1.2.4: Examples of nickel- and molybdenum-promoted oxidative coupling

reactions between ethylene and CO2.

Page 40: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

24

In contrast, acrylate elimination from nickel should occur more easily due to its

lower oxophilicity compared to metals like molybdenum and tungsten, but it faces greater

challenges in the coupling and β-H elimination stages of the catalytic cycle. Higher

temperatures and pressures are generally required to enact CO2/ethylene coupling using

nickel, which, while attainable, can cause prohibitive side reactions like ligand degradation

or multiple CO2 insertions.78,102,103,113 It is suggested that the ancillary ligand environment

can affect the kinetic barrier for coupling,45,104,113 but using it to control reaction energetics

is still not fully understood. By far the most significant issue with using nickel in

CO2/ethylene coupling, however, is its reticence toward β-H elimination, which is

attributed to two factors. The first is that crystallographic evidence indicates that the C-H

bonds on the lactone at the β position to the metal are oriented away from nickel and would

require significant ring distortion (and corresponding ring strain) to put them within

covalent bonding distance (Figure 1.2.1).10,78,86,105,106,109,111,113,149,151,152 The ease with

which group VI metals accomplish β-H elimination from similar metallocycles indicates

an additional factor, which is the necessity of an available empty metal orbital to accept

Figure 1.2.1: Depiction of β-H orientation in a nickelalactone complex and

corresponding d-orbital splitting diagram.

Page 41: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

25

electron donation from the target C-H bond in order to form an agostic intermediate.50,92-

95,107,113 As shown in Figure 1.2.1, the empty dx2-y2 orbital in d8-square planar complexes

like nickelalactones is physically inaccessible for formation of the new Ni-H bond due to

the presence of the existing ligands.108,114

Solutions to this problem have focused on the use of additives to induce ligand

rearrangements that would re-orient the lactone β-hydrogens as well as open up access to

the necessary orbital. The first example of β-H elimination from a nickelalactone species

was achieved by Fischer et al. in 2006 through the addition of dppm (1,1-

bis(diphenylphosphino)methane) to give a dimeric Ni(II)-acrylate species.109

Unfortunately, this process was not catalytic and required the sacrifice of the dppm ligand,

so while it represented a significant step forward in understanding product release from

nickel, it would be a very expensive and wasteful way to generate acrylate. The groups of

Reiger and Kühn employed methyl iodide (MeI) to weaken the Ni-O bond, imparting more

flexibility into the lactone unit to allow for β-H elimination of methyl acrylate. Yields for

this reaction were low and the use of MeI as a reagent made catalysis impossible through

the formation of inactive Ni-halides species.110-112 Recent studies have focused on using an

external sodium base to deprotonate the lactone β-H directly (without the need for a transfer

to nickel) or incorporating a Lewis acid to facilitate ligand reorientation to allow for metal

participation, both with the subsequent release of sodium or lithium acrylate (Scheme

1.2.5).104 Limbach et al. initially applied strong bases like sodium tert-butoxide for this

purpose, and while good acrylate yields were realized, strong bases are incompatible with

the high CO2 pressures needed to initiate the coupling reaction. As a result, either isolation

of the initial nickelalactone prior to base addition or sequential additions of CO2, ethylene,

Page 42: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

26

and base were needed to achieve even low TONs (~10), both of which are highly

impractical.113 Jin et al. developed this concept with the use of Lewis acids for lactone

activation and ring-opening, allowing for the incorporation of a milder base for

deprotonation to yield acrylate.114,115 This process, however, was still not catalytic.

Simultaneous publications in 2014 from the independent groups of Vogt and Limbach

detailed the first cases of one-step nickel-catalyzed CO2/ethylene coupling to produce

lithium or sodium acrylate, respectively. Vogt et al. focused on extending the applicability

of Lewis acids (e.g. LiI) through combination with a zinc reductant and mild bases (e.g.

Et3N) to release lithium acrylate with TONs <25.116 The base in this case was not for direct

deprotonation of the nickelalactone but for quenching the HI byproduct so that it would

not protonate their intended acrylate to propionate. Limbach’s work achieved higher

activities, with TONs of up to 107 for sodium acrylate production using sodium phenoxide

bases for deprotonation and zinc as a reducing agent, presumably to regenerate the

nickel(0) starting complex.117,118 While this accomplishment marks a milestone in the field

of using CO2 as a carbon feedstock for the generation of valuable functionalized

compounds, there are still significant limitations that must be overcome. For example, this

method requires multiple equivalents of a strong reductant, the turnover has been restricted

Scheme 1.2.5: Methods of inducing acrylate removal from nickelalactones using (A)

bases or (B) Lewis acids (LA).

Page 43: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

27

by a side reaction of base with CO2 to form carbonates, and several grams of additives are

required for each reaction.

Rather than focusing on the development of these systems for inducing product

release, the β-H elimination issues can be circumvented altogether by shifting the focus

away from acrylates and toward other value-added chemicals like substituted cyclic

anhydrides or N-substituted acrylamides, which are the topics of the following sections. In

this way, the established reactivity of nickel with respect to CO2/olefin coupling reactions

can be exploited to generate a wider range of products. These behaviors have also been

extended to include systems involving species that are isoelectronic with CO2.

Page 44: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

28

1.2.A.i.: Nickel-Promoted CO2-Alkene Coupling for the Stereoselective Production of

Substituted Cyclic Anhydrides

1.2.A.i.1: Introduction

Cyclic anhydrides (substituted or otherwise) can be produced from ring expansion

and functionalization of nickelalactone species via insertion of CO and subsequent

reductive elimination of the target (Scheme 1.2.A.1).77,153 In order to avoid the catalyst

deactivation pathway leading to inert L2Ni(CO)2 complexes, CO surrogates (denoted as

[CO]) could be employed as slow-release sources of CO. This subsection will address the

application of nickel complexes containing chiral ligands for the regio- and stereo-

Scheme 1.2.A.1: Proposed mechanism for the nickel-catalyzed production of cyclic

anhydrides from CO2, alkenes, and CO surrogates.

Page 45: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

29

selective production of substituted cyclic anhydrides, while the following subsection

discusses using an achiral nickel complex for screening potential CO surrogates for

catalytic applications.

Cyclic anhydrides (such as succinic anhydride (SA) or methylsuccinic anhydride

(MSA)) are commercially valuable chemicals used in the production of

pharmaceuticals,119,120,134 agrochemicals,121 resins and coatings,121 flavors and

fragrances,122 plasticizers,123,124 and as reagents in organic synthesis.125-128 Those

containing a double bond in the five-membered ring, such as maleic anhydride, can

undergo Diels-Alder reactions with olefins to give alkenylsuccinic anhydrides, which are

in high demand in the paper industry129 and as petroleum additives.130,131 While the market

for cyclic anhydrides is highly product-dependent (e.g. 2.72 Mt in 1998 for phthalic

anhydride, 30,000 tons in 2008 for succinic anhydride)132,133 and individually much smaller

than that of acrylic acid and its derivatives,87 specialty applications of these organic

molecules makes them a desirable target. For example, making a substituted cyclic

anhydride like MSA enantioselectively is appealing from a pharmaceutical perspective,

where often only one enantiomer is biologically active or is effective for the intended

application.134

Cyclic anhydride production mainly relies on the metal-free (large-scale) or metal-

catalyzed (mostly research-based) dehydration of dicarboxylic acids (Scheme 1.2.A.2,

reaction 1)135-138 or the heterogeneously-catalyzed oxidation of benzene and benzene

derivatives (Scheme 1.2.A.2, reaction 2).139,140 Both operate under harsh temperature

conditions, require strong dehydrating or oxidizing agents, and are non-stereospecific.

Smaller-scale production from the metal-catalyzed carbonylation of alkynes,141,142 alkenoic

Page 46: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

30

acids,143 and lactones144,145 has been investigated (Scheme 1.2.A.2, reactions 3-5), but these

methods proceed with low yields, significant side products, or limited substrate scope.

They are also not particularly stereoselective. Metal-catalyzed double carbonylation of

Scheme 1.2.A.2: Most prevalent methods of cyclic anhydride production.

Page 47: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

31

epoxides (Scheme 1.2.A.2, reaction 6) has had more success with respect to

enantioselectivity, but it requires pre-synthesizing the chiral epoxide146 and in some cases

isolating the β-lactone intermediate.147,148 Specifically focusing on developing a direct and

inexpensive route to enantiopure versions of cyclic anhydrides through the use of chiral

ligands would therefore be highly advantageous, especially in their role as synthetic

precursors as current methods require special ordering from independent organic labs with

proportional prices and wait times. Such a method would enantioselectively synthesize

substituted cyclic anhydrides directly from olefins, CO2, and CO using a cheap and

abundant metal like nickel (ΔG ~ -6 kcal/mol if using ethylene), which would also generate

less waste and operate under milder conditions that are more tolerant of other functional

Scheme 1.2.A.3: (A) Achiral and (B) chiral nickel-promoted methods for the

production of cyclic anhydrides.

Page 48: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

32

groups. The diversity of products available through functionalization of nickelalactones,

such as nickelamacrocyles, oligomers, carboxylic acids (including acrylic and propionic

acid), and muconic acid derivatives has been established.74,149-152 However, the nickel-

promoted production of cyclic anhydrides has been relatively unexplored, except for two

examples from coupling reactions77,153 and one using nickelalactones that were not CO2

derived (Scheme 1.2.A.3).154,155 There is currently no published work dealing with

enantioselective cyclic anhydride production from metal-promoted coupling of CO2,

alkenes, and CO. Based on the aforementioned precedent, our lab has developed several

nickel systems with the ability to couple CO2 with alkenes and/or alkynes and that

incorporate a variety of mono- or bi-dentate ligands, the most successful of which are of

the [N,N], [P,N], and [P,P] chelating types (Figure 1.2.A.1).114,115,153 These investigations

have determined a set of guidelines for ligand selection, namely that it must have an

electron-donating ability qualitatively equal to or greater than that of 2,2’-bipyridine to

effectively promote olefin binding and enough steric bulk to prevent significant formation

of the bis-ligand complex (e.g. [P,P]2Ni). The work of Greenburg et al. has indicated that

substituted cyclic anhydrides are feasible targets with such bidentate ligand scaffolds on

nickel,153 and the results presented here extend this concept to the use of chiral ligands

Figure 1.2.A.1: Examples of [N,N], [P,N], and [P,P] types of bidentate ligands. tBu =

tert-butyl, Cy = cyclohexyl.

Page 49: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

33

(Scheme 1.2.A.3). Nine nickel complexes containing bidentate [P,P] chiral ligands were

screened for CO2/alkene/CO coupling ability, and it was found that two were competent

for MSA production, albeit with low enantioselectivities. Despite a disappointing lack of

selectivity, these investigations represent initial forays into steric ligand parameters that

could be key to coupling success and product control.

Page 50: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

34

Figure 1.2.A.2: Structure key for ligand numbering. iPr = isopropyl, tBu = tert-butyl.

Page 51: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

35

1.2.A.i.2: Results and Discussion

The nine chiral bidentate [P,P] ligands shown in Figure 1.2.A2 were selected based

on their electron-donating and sterically bulky groups, the coupling advantages of which

were previously discussed and where sterics can also be used to enforce regioselectivity.

They were combined with Ni(COD)2 as a nickel(0) source in order to form the

corresponding P2Ni(COD) complexes, and the majority of them were first screened for

CO2/alkene coupling ability using ethylene (Scheme 1.2.A.4, R’ = H) due to its simplicity,

as it had the least steric hindrance and avoided any complications due to multiple isomers.

Those complexes that were successful with respect to ethylene coupling were then tested

with propylene, a slightly more sterically hindered alkene (Scheme 1.2.A.4, R’ = CH3).

This added a substituent to the nickelalactone that allowed for both regio- and stereo-

isomers, however, analysis was simplified because subsequent CO insertion and product

elimination resulted in indistinguishable regioisomers. These coupling reactions were all

Scheme 1.2.A.4: Methods of nickelalactone production from (A) CO2/alkene coupling

and retrosynthetically from (B) a cyclic anhydride or (C) acrylic acid.

Page 52: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

36

performed in J. Young (<4 atm combined gas pressure) or high-pressure (<13 atm

combined gas pressure) NMR tubes, and the optimized conditions for nickelalactone

formation are listed for each ligand in Tables A1.2 and A1.3 in Appendix 1.

Ligands 1 and 2 established the higher and lower steric limitations (respectively) of

the proposed reaction. Due to the size of its R groups, ligand 1 did not substantially bind

to nickel, forming very little of the 1-Ni(COD) complex (Figure A1.2) and decomposing

upon heating. Formation of the 2-Ni(COD) complex was successful, although a significant

amount of the bis-ligand species (2-Ni-2) was also formed due to the small size of the

methyl substitutents. It was found that 2-Ni(COD) did not bind significantly to ethylene,

even under harsh conditions (with heating or 11 atm of ethylene) (Figure A1.3). Attempts

to hydrogenate COD off of the nickel center using H2 alone had no effect, while addition

of H2 and ethylene only produced ethane and 2-Ni(COD). The methyl groups on 2 are

therefore not large enough to make the smaller ethylene molecule competitive with the

COD as a nickel ligand, which would force it off of the complex. As a result, no further

coupling attempts were made using ligands 1 or 2.

Although ligands 3, 4, and 5 formed the corresponding Ligand-Ni(COD)

complexes in good yields, initial CO2/ethylene coupling attempts were met with varying

degrees of success. Complex 3-Ni(COD) displayed very slow conversion to the 3-Ni-

ethylene complex and similarly slow, partial conversion to the nickelalactone upon

stepwise addition of ethylene and CO2 (Figure A1.15). These conversions were not

significantly enhanced by heating or higher partial pressures of ethylene and CO2. In the

case of 4-Ni(COD), there was similar partial lactone conversion upon CO2/ethylene

addition (Figure A1.16). Treatment with CO produced a small amount of SA on only one

Page 53: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

37

occasion, indicating minimal coupling. The conversion from 5-Ni(COD) to the 5-Ni-

ethylene complex was kinetically fast but would have required unreasonably high gas

pressure to go to completion (Figure A1.17).

The behavior of these complexes suggested a thermodynamic barrier to coupling

that was overcome through the use of a Lewis acid, sodium tetrakis[3,5-

bis(trifluoromethyl)phenyl]borate (NaBAr4F). This solution was inspired by the previous

work of Jin et al., which indicated that the sodium cation stabilized the lactone through its

interaction with the two oxygen atoms, providing a thermodynamic sink that favored

lactone formation.114 While this was primarily applied as a method of inducing β-H

elimination in nickelalactone species, it was extended in this work to encourage CO2/alkene

coupling. Immediate improvement was observed upon addition of NaBAr4F to the coupling

reactions of 3-, 4-, and 5-Ni(COD) (Scheme 1.2.A.5). Using the sodium salt changed the

slow, partial conversion of 3-Ni(COD) to 3-Ni-lactone into full conversion after 24 hours

at 60 °C (Figure A1.4). In the case of 4-Ni(COD), which displayed almost no coupling

ability without NaBAr4F, nearly full conversion to 4-Ni-lactone was achieved after 12 hours

at 50 °C (Figure A1.6). The 5-Ni(COD) complex did not show as large of an improvement

as its conversion to 5-Ni-lactone was very low even in the presence of the Lewis acid

(Figure A1.8), however, this is still an advantage over the complete lack of coupling that it

showed beforehand. The identities of all three lactones were confirmed by the release of

SA upon CO addition (as seen by 1H NMR) and retrosynthetically by reaction of 3-, 4-, or

5-Ni(COD) with SA (Scheme 1.2.A.4, (B); Figures A1.18-A1.20). In the case of the 3-Ni-

lactone, the 31P{1H} NMR spectrum of the coupling product displayed major and minor

peaks, and the major product was identified as a γ-lactone using a combination of HSQC

Page 54: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

38

and HMBC 2D NMR spectroscopy. The minor product was attributed to formation of the

β-lactone, but these peaks did not appear consistently and were occasionally observed pre-

CO2 addition, suggesting that they may have been impurities and not lactone-derived.

Further attempts to confirm via the reaction of 3-Ni(COD) with AA yielded large product

mixtures (Figure A1.21).

While these results are not sufficient to establish a full pattern with respect to steric

influences on the coupling ability of nickel complexes, the activity of 3-Ni(COD)

compared to 5-Ni(COD) (and to a lesser extent, 4-Ni(COD)) suggests that flexibility in the

ligand backbone may be advantageous to nickelalactone formation in these studies.

Therefore, ligand 6 was tested, as it has a similar two-carbon flexible chain and maintains

the general steric size and electron-donating ability of the phosphorous substituents, while

varying them from the five-membered phosphacycle ring featured in ligands 3-5. Ligands

7-9 introduced a ferrocene moiety (and consequently a larger bite angle) into the ligand

backbone, which in the cases of 7 and 8 represented a more rigid motif but with a variation

Scheme 1.2.A.5: Lewis acid enhancement of the 3-,4-, and 5-Ni(COD) coupling

reactions between ethylene and CO2.

γ-lactone β-lactone

Page 55: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

39

in electronic effects, and in the case of 9 modified the linker distance and electronic

influence while maintaining the phosphacycle ring featured in ligands 3-5.

The 6-Ni(COD) species gave low conversion to 6-Ni-lactone (5 atm each gas, room

temperature, 48 hours), but this was attributed to an absence of sodium salt (Figure A1.10).

Product formation was confirmed up SA release after CO addition. Reactions involving

ligands 7-9 were all tested using the NaBAr4F Lewis acid additive. The 7-Ni(COD)

complex was formed in situ from Ni(COD)2 and 7 due to poor conversion at room

temperature and decomposition upon heating. While it showed evidence of binding

ethylene, it did not functionalize CO2 (even at 4 atm total gas pressure, Figure A1.12),

confirmed by HSQC and HMBC NMR spectroscopy using 13CO2. 9-Ni(COD) had a similar

problem in that it appeared to bind propylene (ethylene was not tested) but would not then

perform the oxidative coupling step with CO2 (Figure A1.14). 8-Ni(COD) could couple

ethylene and CO2 but only in very limited amounts (Figure A1.13), as evidenced by 13C

NMR spectroscopic analysis of the 13C-labeled SA product after CO release. Overall, the

nickel complexes containing ferrocene-based ligands exhibited a general lack of

CO2/alkene coupling ability and further trials with them were therefore not attempted. In

past works, nickelalactones containing these types of ligands were prepared via ligand

substitution from bipyridine-Ni-lactones rather than directly from coupling,86,114 so it is

possible that their capacity for electron-donation is insufficient to promote these types of

transformations.

Due to their CO2/ethylene coupling activities, complexes 3-, 4-, 5-, and 6-Ni(COD)

were employed in CO2/propylene coupling reactions (and all in the presence of Lewis

acid). Of these four compounds, neither 4- nor 5-Ni(COD) were able to form the

Page 56: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

40

corresponding nickelalactones (4- and 5-Ni-lactone’, Figures A1.7 and A1.9). In the case

of 4-Ni(COD), this was due to the highly competitive formation of the CO2-adduct, to the

exclusion of the 4-Ni-lactone’ (even under low CO2 pressures). There was evidence of a

similar species being formed in the 5-Ni(COD) reaction, and 13CO2 studies indicated no

sign of CO2 functionalization or production of MSA upon treatment with CO. It is possible

that the rigidity of the ligand backbones in 4 and 5 prevented the necessary structural

rearrangements for incorporating a slightly more sterically hindered alkene into a ring with

CO2, instead favoring ligand displacement. In contrast, 3- and 6-Ni(COD) both

successfully coupled propylene, CO2, and CO to give MSA, as indicated by 1H and 13C

NMR spectroscopy as well as by chiral GCMS (Figures A1.5 and A1.11). The GCMS

analysis, however, indicated only a 1:1.3 ratio between enantiomers for 3 and a 1:2 ratio

(with slight peak overlap) for 6 (Figure 1.2.A.3). Therefore, while 3- and 6-Ni(COD)

presented superior coupling abilities compared to the other ligands tested, this did not

translate to any advantage with respect to enantiomeric selectivity.

Figure 1.2.A.3: Chiral nickel complexes competent for coupling propylene, CO2, and

CO to give MSA. iPr = isopropyl, tBu = tert-butyl.

Page 57: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

41

1.2.A.i.3: Summary

There are numerous examples of nickel-promoted reactions of CO2 with olefins or

alkynes to give nickelalactones, and the incorporation of CO to release cyclic anhydrides

has been established. However, despite a demand for enantiomerically-pure cyclic

anhydrides, there has been no extension of the existing systems to include chiral species.

This work represents an initial foray into the enantioselective production of substituted

cyclic anhydrides using chiral nickel species for coupling alkenes, CO2, and CO. Of the

nine nickel complexes that were tested, only two (3- and 6-Ni(COD)) were competent for

MSA production from propylene, CO2, and CO, and then only with poor selectivities.

Nevertheless, this investigation uncovered a means of enhancing lactone formation through

the use of Lewis acid additives. In addition, these results suggested some important steric

ligand parameters, as well as a preference for a more flexible linker group between the

ligand binding sites. The greatest restriction to a more thorough examination of influential

ligand properties is the scarcity of commercially-available chiral ligands and the difficulties

associated with their synthesis, which prevents a systematic variation of ligand parameters

to determine what is needed for better enantiomeric control. The use of CO gas also

removes the option of catalysis (through the formation of catalytically inept L2Ni(CO)2),

which will be addressed in the following section.

Page 58: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

42

1.2.A.ii.: The Use of CO Surrogates in Nickel-Promoted CO2-Alkene Coupling for the

Catalytic Production of Cyclic Anhydrides

1.2.A.ii.1: Introduction

The use of CO surrogates in cyclization and cycloaddition reactions to produce a

wide range of C=O functionalized species like ketones, esters, and aldehydes, as well as

heterocycles such as indolinones, lactones, and lactams has been established under both

metal-free156 and metal-catalyzed conditions.157,158,159 The incorporation of CO2 as a CO

surrogate has even been demonstrated, for example, in the palladium-catalyzed production

of aldehydes from aryl halides.160 However, the application of CO surrogates in the

catalytic generation of CO2-derived coupling products, such as cyclic anhydrides, has not

yet been explored. Instead, the few cases of using CO surrogates for making cyclic

anhydrides start from non-CO2 sources, such as carboxylic acids.161 The most closely

related example is the generation of α,β-unsaturated lactams, which was published by

Hoshimoto et al. in 2014 and used a (PCy3)2Ni(COD) catalyst (PCy3 =

tricyclohexylphosphine) for the [2+2+1] carbonylative cycloaddition of alkenes or alkynes

with imines (Scheme 1.2.A.6).162 Although this group had already established the

competency of this catalytic system for imine/alkyne coupling reactions,163-165 this is the

first nickel-catalyzed example of this carbonylative cycloaddition reaction using a CO

surrogate. The catalyst loading, however, was high (10 mol%; best TON achieved = 8) and

the imine substrates all contained a sulfonyl or tosyl moiety that would need to be

substituted out in an additional step to increase product diversity.162 Despite these

limitations, their coupling reaction is highly analogous to the nickel-promoted reaction of

CO2, alkenes, and CO previously discussed. Therefore, CO surrogates (regardless of

Page 59: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

43

whether they are applied in a chiral or achiral setting) provide an opportunity to make an

otherwise stoichiometric reaction catalytic (Scheme 1.2.A.1) through the avoidance of

L2Ni(CO)2 (L = monodentate ligand) as a catalyst deactivation pathway. This section

presents experimental results concerning the viability of using CO surrogates to

catalytically liberate a substituted cyclic anhydride, MSA, from CO2/propylene coupling

using the nickel(0) species dcpe-Ni(COD) (dcpe = 1,2-bis(dicyclohexylphosphino)ethane).

This nickel complex was chosen because of its proven CO2 coupling ability.77,115 Overall,

the production of MSA was very low (a highest yield of 9%), even after testing three

separate CO surrogates with three different bases, with the major issue being the rate of

CO production compared to the CO2/propylene coupling rate.

Scheme 1.2.A.6: Previous application of CO surrogates with nickel species for the

production of cyclic lactams and the current proposed use for releasing cyclic

anhydrides.

Page 60: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

44

1.2.A.ii.2: Results and Discussion

Originally, works incorporating CO surrogates like formates required directing

groups for the surrogate, ruthenium co-catalysts, strong bases, high temperatures and

pressures, large excesses of the surrogate, and often had limited substrate scopes and poor

yields.166-168 The groups of Manabe et al. and Tsuji et al. then discovered that weak organic

or inorganic bases could be used to evolve CO from CO surrogates in the absence of

transition metal reagents,169,170 greatly simplifying their application in catalysis. The CO

surrogates that were screened for the production of substituted cyclic anhydrides are listed

in Figure 1.2.A.4, along with the weak bases used with them in order to promote CO

release. Their selection is based on their multitude of appearances in previous

Figure 1.2.A.4: CO surrogates and bases used, with abbreviations.

Page 61: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

45

works.157,158,162,169-172 Initially, stoichiometric reactions were attempted using multiple

equivalents of the CO surrogate/base pair on a small scale (a J. Young tube or scintillation

vial), with the intention of extending these studies to include catalytic reactions once

successful pairings had been established. Experiments began with phenyl formate (PF) and

triethylamine due to their successful application in Hoshimoto’s work. However, it was

found that CO release from PF was too kinetically fast for the propylene/CO2 coupling

reaction to compete, resulting in full and immediate conversion to the dcpe-Ni(CO2)2

complex without any formation of MSA (Table 1.2.A.1, entry 1). This issue was not

alleviated by the using of a different CO surrogate, N-formylsaccharin (NFS), or the use of

NFS even in the absence of a base promoter (Table A1.7). As a result, all further

investigations were performed starting with the pre-synthesized dcpe-Ni-lactone’ (the

nickelalactone from propylene/CO2 coupling). Reactions were initially still performed with

propylene and CO2 gas in the NMR tube headspace in order to allow for potential turnover.

The first attempt at generating MSA from dcpe-Ni-lactone’ and PF/triethylamine

yielded almost no product (Table 1.2.A.1, entry 2), with the primary problem still attributed

to the fast release of CO from PF. Efforts to control the CO release rate by varying the

polarity of the solvent using solvent mixtures were also unsuccessful (entries 3 and 4).

While the first reaction was performed in benzene (deuterated) due to its identity as the

optimized solvent for the previous work, further reactions in 30:70 and 50:50

benzene:tetrahydrofuran (THF) provided no enhancement, even with variations in CO

surrogate:base ([CO]:[B]) ratios (Table A1.8 and Figure A1.23). Performing the reactions

without any propylene or CO2 gas in case increased reaction pressures were inducing faster

CO release had little effect (Table A1.9). The greatest improvement for PF/triethylamine

Page 62: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

46

Entry Ni

Species

CO

Surrogate

[CO]

Equiv.

[CO]

Base

[B]

Equiv.

[B]

Solvent MSA

Produced

(%)b

1 A PF 3 NEt3 4 THF-d8 0

2 B PF 3 NEt3 4 C6D6 0.05

3 B PF 3 NEt3 4 30:70

C6D6:THF-

d8

0.01

4 B PF 3 NEt3 4 50:50

C6D6:THF-

d8

0.03

5c B PF 9 NEt3 12 C6D6 0.09

6 B PF 9 NEt3 24 C6H6 2.00

7c B TCPF 4 NEt3 20 C6D6 9.00

8 B NFS 4 NEt3 10 C6D6 9.00

9 B PF 9 DBU 12 C6D6 -

10 B TCPF 3 DBU 4 C6D6 0.57

11 B NFS 3 DBU 4 C6D6 0.77

12 B PF 9 DMAP 12 C6H6 -

13 B TCPF 9 DMAP 12 C6H6 4.00

14 B NFS 9 DMAP 12 C6H6 6.30

15 B TCPF 4 - - C6H6 5.20

Table 1.2.A.1: Reaction of dcpe-Ni(COD) and dcpe-Ni-lactone’ Species with CO

Surrogatesa

aConditions: 6-10 mg Ni species, X equiv. [CO], and Y equiv. [B] in listed solvent in a J. Young NMR

tube with 0.5-1.5 atm each of propylene and CO2. Heat at 22-70 °C for 2 hrs-3 days. See Appendix 1 for

details. bMSA = methylsuccinic anhydride. Yield determined by GCMS using a naphthalene standard. cNo gas added.

Page 63: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

47

was actually observed from increasing the [CO]:[B] ratio in 100% benzene solvent from

3:4 to 9:24 (Table 1.2.A.1, entry 2 versus 5 and 6), although this only brought the MSA

yield to 2%. Changing the identity of the CO surrogate gave the best results, with

NFS/triethylamine and trichlorophenyl formate (TCFP)/triethylamine both giving 9% yield

of MSA, albeit with different [CO]:[B] ratios. It was found that these ratios had to be

optimized based on the specific CO surrogate and base pairings (Tables A1.10-A1.12).

Changing the base paired with each surrogate gave no improvements, even though DBU

and DMAP were both tested in the nickel-catalyzed generation of cyclic lactams.162 In the

case of PF with either DBU or DMAP (Table 1.2.A.1, entries 9 and 12), there were

problems with product analysis due to overlapping GCMS peaks that could not be resolved.

For the combinations of TCPF (entries 10 and 13) or NFS (entries 11 and 14) with either

DBU or DMAP, the MSA yields only decreased. This was also the case with entry 15,

which repeated a reaction with a CO surrogate in the absence of base except with dcpe-Ni-

lactone’ as the starting nickel species. There was similarly no improvement, although it

confirms that these CO surrogates can produce CO even in the absence of base, making

control of the release rate difficult. Changing the reagent addition order such that the

surrogate was added last in order to minimize CO release before coupling was possible did

not have beneficial results. No evidence of MSA decomposition under the experimental

conditions (Figure A1.24), of MSA side reactions with byproducts from the CO surrogate

in the presence or absence of base (e.g. phenol, Figures A1.25 and A1.26), or of formation

of a deactivating nickel-CO2 adduct (Figure A1.27) was observed.

Page 64: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

48

1.2.A.ii.3: Summary

Several combinations of organic CO surrogates and bases were used in an attempt

to mediate a CO slow-release mechanism to allow for catalysis of the nickel-promoted

production of substituted cyclic anhydrides from the oxidative coupling of alkenes, CO2,

and CO. Unfortunately, CO generation from the surrogates proved too fast for there to be

a competitive coupling reaction, even in the absence of base, and <10% of MSA was

produced from the dcpe-Ni-lactone’ starting material. One potential solution to this issue

would be to screen other nickel(0) species in order to find one that couples alkenes and

CO2 at a faster rate that could become competitive with the rate of CO production from the

surrogate. It has been established, however, that nickel complexes tend to exhibit greater

difficulty in the CO2/alkene coupling step compared to other metals (e.g. group VI), and

conditions like increased temperatures or gas pressures that might promote faster coupling

would also likely enhance the rate of CO release from a surrogate. Alternatively, ex situ

generation of CO in a simple two chamber system has been developed and could be applied

to control CO exposure.173 In this way, a separate surrogate solution could slowly bubble

CO into the lactone solution. This is a similar concept to that presented in a recent

publication by Hoshimoto et al., which indicated that they were able produce cyclic lactams

under very low pressures of CO gas (0.5 atm) in an autoclave without stirring using nickel

catalysts containing trisubstituted monodentate phosphines like tricyclohexylphosphine,

obtaining TONs of up to 100. They attributed this activity to the re-activation of their

L2Ni(CO)2 complex, but it could also be due to the reduced rate of diffusion of such a small

amount of CO gas in the larger volume of the autoclave.174 Regardless, these methods of

slowing down CO introduction even further are the most attractive options for finally

Page 65: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

49

achieving nickel-based catalysis with respect to the production of cyclic anhydrides from

CO2.

Page 66: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

50

1.2.B.: Nickel-Promoted Coupling of Isocyanates and Alkenes for the Production of N-

Substituted Acrylamides

1.2.B.1: Introduction

Isocyanates (R-N=C=O) are small, highly reactive organic molecules that are

isoelectronic with CO2.175,176 It is therefore feasible for them to participate in similar types

of reactions, except that these CO2 analogues have a much smaller thermodynamic barrier

to activation due to the substitution of a nitrogen (with five valence electrons) for an oxygen

(with six), and are much easier to functionalize as a result. Isocyanates are applied in a host

of organic transformations to give a wide assortment of products. For example, they contain

an electrophilic carbon which, when combined with nucleophiles, can yield carbonyl-

containing compounds (e.g. with alcohols or amines to give carbamates or ureas,

respectively) or cleavage products (e.g. with water to give amines and CO2).176 They are

also reactive intermediates, such as those formed by common rearrangement reactions like

Hofmann, Schmidt, Curtius, and Lossen.176,177 Isocyanate-derived products have a

correspondingly diverse range of applications, from the polyurethane industry178 to the

synthesis of diabetic drugs.179 Their metal-catalyzed reaction with olefins to give N-

substituted acrylamides in a manner parallel to that previously discussed for CO2, however,

is of particular interest.

Acrylamides are mainly used in the synthesis of polyacrylamides, which are water-

soluble thickening agents and flocculators that find applications in wastewater treatment,

paper production, ore processing, oil recovery, and the manufacture of permanent press

fabrics or dyes.180,181 Global demand for acrylamide (the simplest, unsubstituted

derivation) alone was 1.16 Mt per year in 2013,182 which does not even include the market

Page 67: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

51

data for all of the specialty applications of substituted acrylamides, such as for temperature-

sensitive gels and resins.183-185 While this is not nearly the demand acrylic acid has, the

diversity of polymer products accessible through N-substituted acrylamides makes them

desirable targets.

Acrylamides are typically prepared from the reaction of acryloyl chloride with

amines186-188 or of acrylonitrile with water or aqueous sulfuric acid (Scheme 1.2.B.1).89,189

The first reaction does not require a catalyst, but the variety of substituents is limited and

produces chloride salts as waste.190 The second reaction with acid is catalyst-free but the

conditions are harsh, and with water requires either a reduced copper species189,191,192 or

immobilized nitrile hydratase to catalyze the process.193 The former gives poorer yields,

deactivates quickly, and can also catalyze the production of unwanted side products like

polymers or acrylic acid,193-196 so the latter is preferred. In both cases, the starting materials

require multi-step pre-synthesis and isolation, where the acryloyl chloride is derived from

acrylic acid and a chloride source and the acrylonitrile is from propylene, ammonia, and

oxygen. Developing a more direct method from a C1 carbon source to the intended

Scheme 1.2.B.1: Large-scale methods of synthesizing acrylamides.

Page 68: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

52

acrylamides, such as from isocyanates and alkenes, would therefore be economically

beneficial and produce less waste.

The inherent activity of isocyanates permits a multitude of metal-free

transformations with strong nucleophiles, however, their interactions with alkenes and

alkynes are metal-promoted (with the exception of a very few, specific, highly reactive

cases that do not produce acrylamides).197,198 While there are many examples of metal-

catalyzed coupling reactions between isocyanates and unsaturated compounds like

aldehydes,199-201 imines,199, and alkynes202-204 that feature promotors incorporating a variety

of metals (such as osmium, rhodium, iridium, platinum chromium, tungsten, copper, iron,

and nickel),205 those between isocyanates and alkenes are dominated by nickel-containing

complexes.205,206 The earliest and most pivotal work in this area was done by Hoberg

beginning in the 1980’s, in parallel to his studies of CO2/alkene coupling on nickel. The

first publications of Hoberg et al. on this topic established the successful formation of the

corresponding azanickelacyclopentanone from ethylene and phenyl isocyanate followed

by hydrolysis to give propanilide or CO insertion to give N-phenyl succinimide.207,208 They

built upon these preliminary results and were able to release the desired N-phenyl

acrylamide by heating (80 °C), but at this stage the reaction only proceeded

stoichiometrically and had to be performed at very low temperatures (-78 °C).209 Soon

after, they were able to achieve catalysis by using slightly higher ethylene pressures and a

more electron-donating ligand on nickel, but their TONs were very low (<20) and they had

a mixture of acrylamide products from both single and double insertion of ethylene into

the metallocycle.210 The scope of the reaction was extended to include a variety of alkenes,

including linear alkenes of the type R1R2C=CH2 (where R1 = Ph, OEt, SPh, CO2Me, CH3,

Page 69: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

53

and CF3 when R2 = H or R1 = R2 = F),211-214 cyclic alkenes (e.g. cyclopentene),215 dienes

(e.g. 1,3-butadiene),216-218 and bicyclics,219 but in all cases the other reactant was phenyl

isocyanate or its derivatives (i.e. one example also used 4-methylphenyl isocyanate). There

was also no significant increase in TONs. The incorporation of additives to divert the

reaction to other products was also investigated, which includes the aforementioned use of

acid to give a saturated amide207,212,213 or CO to yield a succinimide,207,211,212 as well as the

use of oxidizing agents (e.g. FeCl3 or I2) to produce α,ω-diamides.220

The proposed reaction sequence for nickel-catalyzed isocyanate/alkene coupling is

shown in Scheme 1.2.B.2, where (under the right experimental conditions) acrylamide

release occurs through thermally-induced β-H elimination from the nickel metallocycle.209

This represents a drastic change from nickelalactone behavior, suggesting that replacing a

Ni-O bond with a Ni-N bond allows for greater structural flexibility in the metal ring. In

this way, acrylamide formation via nickel-induced coupling reactions bypasses the issues

of CO2/alkene coupling while still accessing valuable functionalized products.

For a substituted alkene like propylene, two potential coupling products could be

formed based on the position of the alkene R group on the nickelacycle (α- or β- to the

metal center, Scheme 1.2.B.2). Location at the α-carbon gives disubstituted α,β-unsaturated

amides upon β-H elimination, while being at the β-position yields a 1,1-disubstituted

acrylamide.206 In Hoberg’s studies, which incorporated simple trisubstituted alkyl or aryl

phosphine ligands like tricyclohexylphosphine or triphenylphosphine, the major species

contained the R group at the α-position,206,211,212 except in the case of fluoroalkenes.213

Later work by Schleicher and Jamison demonstrated that by using much bigger, sterically

bulky ligands like N-heterocyclic carbenes (NHCs) on nickel (the coupling ability of which

Page 70: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

54

had already been established for isocyanates and alkynes),221 the regioselectivity could be

forced to favor the position of the R group at the β-carbon to exclusively give 1,1-

disubstituted acrylamides. While this is an impressive example of regio-control and they

achieved good yields (up to 86%) and selectivities (up to 100%), their catalyst loading was

high (10 mol%), so their best TON was 9.190

With the exception of the research presented above by Jamison’s group, there has

been very little advancement in the field of nickel-catalyzed isocyanate/alkene coupling

reactions for acrylamide production since Hoberg’s initial work, especially in terms of

increasing catalytic activity.202,205 Newer studies have instead focused on nickel-catalyzed

Scheme 1.2.B.2: Nickel-catalyzed isocyanate/alkene coupling reactions with

conditions for accessing specific acrylamide regioisomers as well as propionamides and

succinimides.

Page 71: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

55

cycloaddition reactions between isocyanates and unsaturated compounds (alkynes, allenes,

dienes, acrylates) to give heterocycles, in generally low yields (TONs <20) and often with

poor selectivities (due to potentially multiple isocyanate insertions).222-229 This section

concerns the application of a well-established nickel complex for CO2/alkene coupling

(dcpe-Ni(COD); dcpe = 1,2-bis(dicyclohexylphosphino)ethane), toward the coupling of

ethylene with ethyl, phenyl, and tert-butyl isocyanate to give N-substituted acrylamides

(Scheme 1.2.B.3). By varying steric and electronic factors on the isocyanate substituents,

it was hoped that a better understanding of reaction promoting parameters would be

obtained. Overall, it was found that while dcpe-Ni(COD) successfully performed the

coupling reactions, the greatest obstacle to catalysis was the highly competitive isocyanate

trimerization pathway to form isocyanurates. This resulted in only trace amounts of the

intended acrylamide or succinimide species and which ultimately could not be overcome.

1.2.B.2: Results and Discussion

Initial experiments were focused on evaluating the coupling ability and product

selectivity of the nickel species, so they were performed as stoichiometric reactions in

either J. Young (total gas pressures <4 atm) or high-pressure (total gas pressures <13 atm)

NMR tubes. All isocyanates were first tested with ethylene because it is sterically

undemanding and avoids the complication of introducing regioisomers. These un-

Scheme 1.2.B.3: Nickel-catalyzed isocyanate/alkene coupling reactions pursued in this

work.

Page 72: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

56

optimized conditions, however, do not preclude the extension of any preliminary results to

encompass catalytic reactions using substituted alkenes after coupling competency had

been established. Even though the coupling activity of isocyanates usually increases with

the electron-withdrawing nature of the group on the nitrogen (due to the displacement of

electron density onto the R group upon nucleophilic attack), the isocyanate screening began

with ethyl isocyanate because of its simplicity and moderate steric bulk. It also has a low

boiling point, which allows for its addition to the reaction solution via vacuum transfer

after ethylene pressurization. As nickel-ethylene complexes are often unstable in the

absence of a headspace of ethylene, this process ensured the formation of the nickel-

ethylene species prior to isocyanate introduction, which could contribute to favoring the

formation of the desired coupling product.

The stepwise addition of 2 atm of ethylene and 1 equivalent of ethyl isocyanate to

10 mg of dcpe-Ni(COD) began with the promising appearance of a new set of doublets in

the 31P{1H} NMR spectrum (Figure A1.28). These would be consistent with the

inequivalent phosphine groups found in the five-membered nickel metallocycle formed

from ethylene/isocyanate coupling. However, the reaction did not proceed to completion

at room temperature, and upon heating to 60 °C and then 70 °C, multiple phosphorous-

containing products began to appear (Figure A1.28). Adding CO at the end to release any

nickel-bound organic compounds resulted in a mixture of products, some of which

contained olefinic protons (Figure A1.29). Suspecting that 1 equivalent of isocyanate was

perhaps insufficient for the reaction to proceed fully, the experiment was repeated in a

similar manner using 10 equivalents of ethyl isocyanate. After several days at room

temperature, the same (multiple) nickel species had returned (even in the absence of heat;

Page 73: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

57

Figure A1.30), and the solvent was removed under vacuum. The resulting mixture was

recrystallized from pentane/toluene to yield a solid and a yellow solution. Once the yellow

solution was dried under vacuum, the two solids were independently dissolved in C6D6 and

treated with CO gas. The solid from the pentane/toluene recrystallization was N-ethyl

succinimide, while that from the supernatant was N-ethyl acrylamide, as confirmed by

comparison of the 1H NMR spectra (Figures A1.31 and A1.32) to authentic compounds

and by GCMS analysis. Reaction of dcpe-Ni(COD) with purchased N-ethyl acrylamide

yielded a nickel complex with a set of two phosphorous peaks that were very close in

chemical shift to one of the coupling products, tentatively assigned as the η2-acrylamide

species (Figure A1.33). These findings, however, only accounted for two of peak sets in

the 31P NMR spectrum.

In order to verify that all of the nickel species appearing in the NMR spectrum

contained ethylene, 13C-labeled ethylene was used in a coupling experiment under similar

conditions (~1 atm 13C-ethylene, 10 equivalents of ethyl isocyanate). It was immediately

observed that only two of the four sets of phosphorous peaks indicated incorporation of the

13C-ethylene through extra splitting due to 13C-31P interactions (Figure A1.34). Through a

combination of 2D (HSQC, HMBC, COSY) and 1D (1H, 31P, 13C) NMR experiments, it

was determined that the other two phosphorous-containing reaction species (accounting for

the major reaction products) did not incorporate any ethylene (Figure A1.35). Assignments

for the metallocycle and η2-acrylamide nickel complexes formed by ethylene/ethyl

isocyanate coupling are shown in Figure 1.2.B.1. The other two species were attributed to

Page 74: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

58

isocyanate-associated nickel species on the pathway to isocyanate dimerization and/or

trimerization, which are well-known reaction behaviors of isocyanates.176,230-235 Peaks

potentially belonging to the isocyanurate cyclotrimerization product were also identified

in the 1H NMR spectrum (Figure A1.36).

Attempts to minimize or suppress the isocyanate cyclotrimerization pathway

included increasing the ethylene:isocyanate ratio (both in an NMR tube and using a Parr

reactor), performing sequential additions of isocyanate as a method of decreasing the

average isocyanate concentration, and using base additives to try to promote the production

Figure 1.2.B.1: 31P{1H} NMR spectrum in C6D6 of the nickel-promoted coupling

products from the ethylene/ethyl isocyanate reaction (2 atm/10 equiv.). Products C/D

do not incorporate ethylene. 300 MHz spectrometer.

E E F F

C D C D

A

B

Page 75: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

59

of free acrylamide. Using a combination of 8 atm ethylene with 2 equivalents of ethyl

isocyanate in a high-pressure NMR tube reduced the observed products to the nickel η2-

acrylamide and one of the nickel-isocyanate adducts in an approximately 50:50 mixture

(Figure A1.37). As this was nearing the pressure limitations of the NMR tube, further

reactions were performed in a Parr reactor (Table A1.14). It was found that pressures of

~1,000 psi ethylene successfully stopped the cyclotrimerization process, however, the

reactions yielded no acrylamide or any other recognizable nickel species along the coupling

pathway (entries 9-11, Figures A1.38 and A1.39). A possible reason for this would be

multiple ethylene insertions into the five-membered nickelacycle. At pressures above 600

psi ethylene, no acrylamide or acrylamide-containing nickel species were observed (entries

6-11), but at pressures between 300-600 psi ethylene, the cyclotrimerization pathway still

remained active (entries 1, 3-5, Figures A1.38 and A1.39) unless the equivalents of

isocyanate were so low that no reaction occurred (entries 2 and 9). Adjustments in the

number of isocyanate equivalents yielded no significant improvements in selectivity

(entries 2 vs. 3, 6 vs. 7, and 10 vs. 11). Suspecting that product inhibition might be

contributing to the generation of the isocyanurate as the major organic species, a sodium

phenoxide base additive (sodium 3-fluorophenoxide) was used to try to encourage

acrylamide release from the nickel η2-acrylamide complex. Unfortunately, this did not

result in greater acrylamide yields (Table A1.13, entry 4). Attempts to induce free

acrylamide from the nickel η2-acrylamide bound species using other types of bases (e.g.

sodium tert-butoxide, 1,8-diazabicyclo[5.4.0]undec-7-ene = DBU) were also unsuccessful.

One coupling attempt was made using 6-Ni(COD) (see Figure 1.2.B.4 for structure) to

Page 76: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

60

ascertain if a change in ligand would enhance selectivity, but this yielded no improvements.

Further experiments with ethyl isocyanate were therefore not pursued.

In an effort to discourage the cyclotrimerization reaction, tert-butyl isocyanate was

employed in the ethylene coupling reaction due to its sterically more hindered R group.

The first coupling experiment gave two major phosphorous-containing species in the

31P{1H} NMR spectrum after gentle heating (60 °C) over several days (Figure 1.2.B.2),

without any sign of an isocyanurate in the 1H NMR spectrum. Addition of CO gas gave the

free N-(tert-butyl) acrylamide in very small yield, the identity of which was verified by

both 1H NMR spectroscopy (Figure A1.40) and GCMS analysis. Confirmation of the

concurrent production of N-(tert-butyl) succinimide was obstructed by the NMR peaks

corresponding to the ligand. However, the position of one of the phosphorous peak sets is

consistent with the formation of the tert-butyl isocyanate/ethylene derived nickel

metallocycle if compared to the analogous species prepared from ethyl isocyanate and

Figure 1.2.B.2: 31P{1H} NMR spectrum in C6D6 of the nickel-promoted coupling

products from the ethylene/tert-butyl isocyanate reaction (2 atm/10 equiv.) See Figure

1.2.B.1 for labeling guide. 300 MHz spectrometer.

F F E E

B

Page 77: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

61

ethylene. Despite this promising start, future coupling experiments using tert-butyl

isocyanate were unfortunately unable to reproduce these results outside of one occasion, in

which the yield of acrylamide was similarly limited. In all other cases, coupling did not

occur and the reaction mixture sat at the dcpe-Ni-ethylene complex, even with low ethylene

pressures and high isocyanate equivalencies at elevated temperatures. An experiment

featuring stepwise reagent addition beginning with tert-butyl isocyanate indicated the

formation of a nickel-bound isocyanate species. However, the isocyanate ligand was

immediately displaced upon ethylene addition (Figure A1.41). If the mechanism for the

coupling pathway requires initial binding of the isocyanate before it can proceed, the ease

with which ethylene displaces tert-butyl isocyanate may inhibit the intended reaction. In

addition, rapid decomposition and polymerization of tert-butyl isocyanate was observed

under preparative and storage conditions, so it was hypothesized that this instability may

account for its limited capacity to engage in coupling interactions with ethylene to a

significant degree.

The final reactive ethylene coupling partner to be applied to this transformation was

phenyl isocyanate, with which Hoberg found great success in both the production of

nickelacycles and the thermally-induced elimination of N-phenyl acrylamide. Initial

studies indicated full conversion from the dcpe-Ni-ethylene complex to one major nickel

species (and giving only a trace amount of a second) after only 1 hour at room temperature

with only 1 equivalent of phenyl isocyanate (Figure 1.2.B.3). The 31P spectral peaks of this

new compound appeared as singlets, which is consistent with those of the nickelacycles

formed from the coupling reactions between ethylene and the other two isocyanates.

Conversely, the nickel η2-acrylamide species have all displayed large enough coupling

Page 78: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

62

constants for their phosphorous NMR peaks to appear as two distinct doublets (and which

likely accounts for the identity of the other trace complex appearing in the NMR spectrum).

No clear vinylic peaks appeared in the 1H NMR spectrum either before or after addition of

CO gas, and no acrylamide was detected by GCMS. However, these analysis methods did

indicate the presence of N-phenyl succinimide (Figure A1.42). The results suggested that

while the coupling reaction between ethylene and phenyl isocyanate was comparatively

facile, elimination of the acrylamide was more difficult than for the other isocyanates. This

barrier could not be overcome by simple heating, as was the case with Hoberg’s

experiments.209

In a manner analogous to the previously mentioned methods aimed at inducing β-

H elimination in nickelalactones for the formation of acrylates, a Lewis acid and several

bases of varying strengths were employed to try to force acrylamide generation from the

nickelacycle. In the case of running the coupling reaction in the presence of a Lewis acid,

Figure 1.2.B.3: 31P{1H} NMR spectrum in C6D6 of the nickel-promoted coupling

products from the ethylene/phenyl isocyanate reaction (1 atm/2 equiv.). See Figure

1.2.B.1 for labeling guide. 300 MHz spectrometer.

F F

Page 79: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

63

NaBAr4F, there was no evidence for production of the nickel η2-acrylamide species or free

acrylamide by NMR spectroscopy or GCMS (Figures A1.43 and A1.44). It therefore

appeared as though this Lewis acid was not capable of significantly disrupting the Ni-N

bond in the nickelacycle to allow for ligand re-orientation and subsequent β-H elimination.

The addition of triethylamine as a weak base similarly had no effect (Figure A1.45).

Reasoning that perhaps triethylamine was simply not basic enough to deprotonate the β-

hydrogen of the nickelacycle, a slightly stronger base was utilized. The reaction revealed

some promising olefinic peaks when performed in the presence of DBU, (Figures A1.46

and A1.47), however, a 2D HSQC NMR spectrum indicated that there were no vinylic

methylene peaks in the product generated after treatment with CO, and no acrylamide was

identified by GCMS. This would suggest that while DBU induces a breakdown of the

nickelacyle, it also causes further reaction either at the nickel center or of the desired

product. Adding hydrochloric acid gas to the compound generated by DBU in order to

assess its reactivity gave no other insight into its identity, although its associated olefinic

peaks did disappear from the NMR spectrum. While this verifies the presence of an

unsaturated species, it was not the targeted acrylamide, so the effect of using progressively

stronger bases was then tested. It was found that sodium 3-fluorophenoxide was unable to

initiate the desired elimination (Figures A1.48 and A1.49). With NaOtBu as an additive,

however, a new set of olefinic peaks appeared in the proton spectrum, potentially as part

of the vinylic set of N-phenyl acrylamide. A combination of HSQC and COSY NMR data

obtained using the reaction solution (Figure A1.50) indicated that these peaks correlated to

a –CH=CH2 functional group consistent with an acrylamide. Encouraged by these results,

enhancement of the coupling reaction and hence the acrylamide yield was attempted

Page 80: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

64

through the use of a Parr reactor in order to accommodate higher ethylene pressures and

equivalents of phenyl isocyanate. Under these conditions (300-500 psi ethylene, 20-50

equivalents phenyl isocyanate, 20 equivalents NaOtBu, 60 °C, 16 hours), no apparent

acrylamide was generated. Consequent NMR-scale experiments revealed the reaction of

phenyl isocyanate with NaOtBu, the rate of which is likely accelerated when combined

with the excesses of base and increased pressure conditions of the Parr environment.

Therefore, while NaOtBu is a competent base for inducing β-H elimination and product

release from more reticent nickelacycles derived from alkenes and isocyanates, it is

incompatible with the conditions necessary for exploring catalysis and would only be

suitable for generating acrylamides on sub-stoichiometric scales.

Due to Hoberg’s achievement of β-H elimination from an ethylene/phenyl

isocyanate-derived nickelacycle,209 the final attempt to devise an alternative way of forcing

off the acrylamide focused on the use of other ligands (depicted in Figure 1.2.B.4.). Four

Figure 1.2.B.4: Alternative ligands tested in Ligand-Ni(COD)-promoted

ethylene/isocyanate coupling reactions.

Page 81: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

65

other Ligand-Ni(COD) complexes were tested (including 6-Ni(COD) from section

1.2.A.i.) for coupling ability with respect to ethylene and phenyl isocyanate and their

enforcement of β-H elimination from the metallocycle intermediate. They were chosen

based on their previously-demonstrated CO2/alkene coupling capabilities, and in the case

of BenzP*-Ni(COD), catalytic activity. For this application, however, none of these

complexes yielded acrylamide, instead giving results that ranged from unreactive (BenzP*)

to partial activity (6, giving the metallocycle but no acrylamide) and decomposition (MAP,

P2N2). A more systematic ligand variation would likely be required to fully understand the

factors promoting or inhibiting the sought-after elimination.

1.2.B.3: Summary

A series of experiments aimed at leveraging the inherent reactivity of isocyanates

in nickel-promoted alkene transformations was undertaken, with the ultimate goal of

significantly enhancing the catalytic production of N-substituted acrylamides in this

respect. It was instead found that diversifying the pool of isocyanate starting materials

elicited a range of challenges, which perhaps influence the prevalence of certain

isocyanates in alkene coupling reactions in the literature (e.g. phenyl isocyanate). In the

case of isocyanates possessing a small, electron-donating R group like ethyl, a

cyclotrimerization pathway that was highly competitive with the intended coupling

transformation was activated, and conditions that suppressed this side reaction were

similarly detrimental to acrylamide formation. Employing a larger R group, like tert-butyl,

effectively discouraged cyclotrimerization but then provided too great an obstacle to

coupling. Phenyl isocyanate was not inclined to cyclotrimerize and it easily underwent an

oxidative coupling reaction with ethylene at a nickel center, fully converting to the

Page 82: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

66

metallocycle in short time periods under mild conditions. However, it was then very

difficult to induce product elimination, even with strong base additives that are also

unsuitable for catalysis. Thus each isocyanate faced a unique challenge at a different stage

in the coupling reaction with alkenes. Future work should approach these problems from a

ligand perspective. By instead starting with a nickel complex that is known for spontaneous

or thermally-induced acrylamide release (where the ligand is reminiscent of those

established by previous work, such as NHC’s or a PR3 group) and then assessing any issues

that occur through variation of the identity of the isocyanate, factors that control the

intended behavior can be more easily identified and controlled. Alternatively, a practical

experimental solution could involve enacting a slow addition method for introducing the

isocyanate into the reaction mixture continuously but gradually over time. This may help

discourage polymerization pathways while still benefitting from the ease of coupling and

product elimination advantages that were displayed by ethyl isocyanate.

1.2.C.: Section 1.2 Summary

Nickel has an illustrious history with respect to coupling CO2 (and CO2 analogues)

with olefins and alkynes, with the consequent production of a vast array of commercially

valuable chemicals from an inexpensive, widely available carbon feedstock. The use of a

cheaper, more abundant metal like nickel also represents an economically beneficial

transition from precious metals like palladium or ruthenium. The work presented herein

sought to extend previously established methods for the transformation of CO2 to include

specialty targets like enantiopure substituted cyclic anhydrides and N-substituted

acrylamides. While ultimately these goals were not achieved, many of the limiting factors

were identified, which provides a basis for developing ways to overcome these obstacles.

Page 83: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

67

The outlook for devising and improving upon the nickel-catalyzed production of these

compounds is therefore promising.

Page 84: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

68

1.3: Transition-Metal-Catalyzed Hydrogenation of CO2 to Methanol

1.3.1: Introduction

In addition to the previously discussed metal-catalyzed coupling reactions between

CO2 and unsaturated organic molecules for the generation of new C-C bonds, a second

means of CO2 activation and transformation is through its reduction using dihydrogen (H2).

Depending on how many equivalents of H2 are incorporated, CO2 hydrogenations can yield

a variety of C1 products, including CO, formic acid, methanol, and methane (Scheme

1.3.1).236,237 Two that have been of particular interest as target molecules are formic acid

and methanol, due to the market for the chemicals themselves, their potential as hydrogen

storage materials, and the feasibility of their reaction thermodynamics236,238-240,266 The

following section will focus on the transition-metal-catalyzed production of methanol from

CO2 and H2.

The global market demand for methanol has seen significant recent growth (49 Mt

in 2010 to 80 Mt in 2016)27,28,241, which has been attributed to its incorporation into newer

end-uses (such as its conversion to olefins for the production of plastics) and the increasing

Scheme 1.3.1: Products of CO2 hydrogenation reaction and the Gibbs free energy for

formic acid and methanol generation.

ΔG° = -2.4 kcal/mole

ΔG° = 7.8 kcal/mole

Page 85: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

69

popularity of methanol or methanol-derived compounds (like dimethyl ether) for fuel

blending or fuel substitutes as a way of extending gas resources.238,241 The methanol-to-

olefin process alone has gone from accounting for only a trace amount of methanol

production to being the second largest application of methanol by end-use (an 18% increase

over a 5 year timeframe).27,241 This trend reflects the fact that despite its widespread

recognition as an organic solvent, >95% of the methanol produced annually is actually

converted into other valuable chemicals.27 Generating methanol from CO2 can therefore be

viewed as a means of accessing a host of other functionalized species from a carbon source

that might otherwise be unavailable to them by direct reaction (see Chapter 2 for a more

detailed discussion of this topic). For example, formaldehyde and acetic acid account for

over 35% of the global methanol demand (formaldehyde far exceeding acetic acid at 27%

and 9%, respectively),27 and they are subsequently converted into other high-demand

products like paints, adhesives, textiles, films, resins, plastics, and pharmaceuticals.242-244

The economic benefits of sourcing methanol from an inexpensive, non-toxic, and abundant

carbon feedstock like CO2 would therefore be realized in a diverse array of industries.

As mentioned briefly in the introduction to this chapter, industrial-scale methanol

production currently relies on the use of an alumina-supported mixed copper/zinc oxide

Scheme 1.3.2: Industrial method of methanol production from syngas, depicting (1) the

component equilibria and (2) the overall reaction.

Page 86: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

70

heterogeneous catalyst for the conversion of syngas (a mixture of CO, CO2, and H2)

through the series of reactions shown in Scheme 1.3.2.245,246 Unfortunately, the co-

production of water necessitates a time and energy-intensive post-synthetic distillation

process for the isolation of the desired methanol.247 While highly efficient, the harsh

temperature and pressure conditions of the methanol production process are incompatible

with many other functional groups, prohibiting the use of a one-pot method for

streamlining further transformations. Accessing methanol using metal-catalyzed CO2

hydrogenation under milder conditions would make this goal much more feasible.239,240

Additionally, syngas is mostly derived from non-renewable coal and natural gas

resources,242 with only a small percentage available via low-yield248,249 biomass

fermentation.250-252 The incorporation of CO2 as a methanol C1 carbon source therefore

represents the opportunity for a much-needed transition from fossil-fuel reliance to a large-

scale renewable resource, imparting environmental as well as economic benefits.

Developing this methanol production method to integrate H2 derived from solar,

geothermal, or hydro powered electrolysis would maximize its sustainability, a process

which has in fact been recently realized on an industrial scale with the opening of the first

renewable CO2-to-methanol plant in Iceland in 2012. Its production capacity in 2015 was

over 5 million liters per year and it recycles an estimated 5.5 thousand metric tons of CO2

annually.253

In addition to its role as an intermediate on the pathway to other commercial

chemicals, methanol is an attractive potential hydrogen storage material. Methanol is 12.6

weight percent hydrogen, making its hydrogen storage capacity about three times that of

the other popular CO2 hydrogenation product, formic acid (4.4 weight percent).239 A major

Page 87: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

71

challenge to the practical application of fuel cells as alternative energy sources is the

necessity of transporting a highly explosive compressed gas, which is both dangerous and

incompatible with the current infrastructure designed for storing, transporting, and

distributing liquid fuels.238 By employing a catalyst for CO2 hydrogenation to methanol,

H2 can be safely stored as a stable, easy to transport liquid until it is needed, at which time

methanol dehydrogenation is initiated to release the H2 (and CO2 enters back into the

storage cycle; Scheme 1.3.3). This has the added benefit of providing energy with no net

CO2 production. The concept of storing and transporting energy in the form of chemical

bonds through a reversible cycle of bond-making and bond-breaking is also relevant to

alternative energy sources like solar or wind, which produce power intermittently. In order

to meet a continuous demand, the excess electrical energy generated by these processes at

a given time could be used in chemical synthesis, thereby storing it as chemical energy to

be released when their production levels are low or nonexistent.238

Extensive studies into metal-free and metal-catalyzed CO2 hydrogenation to

methanol have been reported.239,240 Despite the use of frustrated Lewis pairs237,254,255 and

N-heterocyclic carbenes256,257 as catalysts for this transformation, the majority of studies

concern the use of transition metal catalysts due to generally superior activities.238,239,240

Scheme 1.3.3: General method of H2 storage using reversible metal-catalyzed CO2

hydrogenation/dehydrogenation.

H2

CO2

HCO2H

CH3OH

Hydrogenation Dehydrogenation H2

Potentially

renewable

source

Fuel Cell

Commodity chemicals

[M] [M]

Page 88: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

72

While a multitude of reductants (with or without metal catalyst) have been successfully

applied to this reaction (including boranes,258,259 silanes,257,260 and 1,2-dihydropyridine261),

H2 represents the less expensive option with the least waste generation.239,262 Finally,

heterogeneous CO2 hydrogenations require harsh conditions (>250 °C) and often also

display low activities and poor product selectivity (such as partial CO2 reduction to yield

CO, which is a poison for fuel cells).240,263 Therefore, only homogeneous transition-metal

catalysts for CO2 reduction to methanol using H2 will be presented.

The overall conversion from CO2 to methanol is (slightly) thermodynamically

favorable, with a ΔG = -2.4 kcal/mole.236 However, there is a significant kinetic activation

barrier to the direct hydrogenation of CO2 to methanol that would require the use of very

harsh conditions (e.g. highly elevated temperatures and pressures) to overcome, even in the

presence of a catalyst. These conditions also provide access to undesirable side products

like CO or methane.239 For example, the first case of CO2 hydrogenation to methanol was

reported in 1993 by Tominaga et al. using a [Ru3(CO)12] catalyst with KI at 240 °C and 80

bar (1200 psi) total pressure to give a TON of only 95. The selectivity was also very low,

producing a mixture of CO, methane, and ethane in addition to methanol.264 Future efforts

using other metal carbonyl catalysts (containing M = Ir, Rh, W, Mo, Co, and Fe) did not

overcome these limitations, with Ru providing the best results.265 Consequently, most work

in the field of CO2 hydrogenations takes advantage of CO2 derivatives (e.g. formic acid,

carbonates, esters, amides, carbamates, and ureas) with lower activation barriers to

hydrogenation in order to carry the reaction forward stepwise under much milder

conditions and with greater selectivities (Scheme 1.3.4).236,239 Many of these compounds

are formed upon CO2 sequestration,240 so their incorporation into CO2 hydrogenation

Page 89: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

73

pathways provides an opportunity for a smooth transition from CO2 capture to

transformation and utilization. Hydrogenation of CO2 to formic acid is thermodynamically

uphill by ~7 kcal/mole,266 which can be challenging for many systems. Consequently,

nearly all examples of generating methanol by this pathway do so starting from formic acid

rather than CO2,239,267,268 with the exception of only one recently developed iridium

catalyst.269 There are similar hydrogenation examples of generating methanol from

formate, carbonate, carbamate, and urea starting materials270-272 (covered in detail in the

cited reviews).239,240 While these cases establish proof of concept in the feasibility of using

CO2 derivatives to facilitate CO2 hydrogenation reactions, the necessity of independently

Scheme 1.3.4: Methods of indirect hydrogenation of CO2 to methanol.

Page 90: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

74

synthesizing and isolating each species detracts from the appeal of this method on the basis

of cost and efficiency. In addition, the challenges with transitioning a two-pot method (CO2

to derivative, derivative to methanol) into a one-pot sequence are non-trivial. It is instead

preferable to leverage the kinetic advantages of these derivatives in situ by using additives

(e.g. amines or alcohols) to generate the CO2 derivative as an intermediate in a one-pot

reaction. With only a few exceptions, previous applications of this strategy of shuttling

CO2 to methanol can be divided based on whether they proceed via an ester or a formamide

intermediate (Scheme 1.3.5).

Independent investigations by the groups of Sanford and Leitner established the

field for this manner of achieving a more level kinetic profile for CO2 hydrogenation, which

encompasses a rather limited number of overall examples. Sanford et al. first presented a

cascade catalysis method for achieving CO2 hydrogenation to methanol employing three

different catalysts in a one-pot, three-step transformation (Figure 1.3.1, (1)).262 CO2 was

first hydrogenated to formic acid using ruthenium catalyst A, followed by conversion to

methyl formate (an ester) using scandium catalyst B, and finally hydrogenation of the ester

to methanol using ruthenium catalyst C. Generation of the intermediate ester in these types

Scheme 1.3.5: Metal-catalyzed CO2 hydrogenation to methanol via (1) esters or (2)

formamides.

Page 91: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

75

of reactions is achieved by the addition of an alcohol, often methanol, which should be re-

generated at the end along with a new molecule of methanol. Initial activity (TONs < 3)

was limited by the deactivation of catalyst C due to high CO2 pressures and by the scandium

triflate catalyst for step two. By devising a two-vessel setup where the first two steps

occurred in the inner vessel, followed by volatilization of the ester to an outer vessel

containing catalyst C in solution, they were able to achieve TONs of up to 21 for the

selective production of methanol at 75-135 °C and 40 bar (600 psi) total gas pressure.262

Overall, these TONs were quite low, and using three incompatible catalysts in a two-vessel

setup would be highly impractical. However, these results represented a significant step

forward in terms of increasing product selectivity and decreasing temperature and pressure

requirements.

One year later, Klankermayer and Leitner reported the one-pot hydrogenation of

CO2 to methanol (selectively), achieving TONs of up to 221 using a single ruthenium-

triphos catalyst (Figure 1.3.1, (2)) with a co-catalytic amount of an acid

(bis(trifluoromethane)sulfonamide) at 140 °C and 80 bar (1200 psi) total gas pressure. The

in situ intermediate was again an ester (ethyl formate), formed from the addition of ethanol

to the reaction mixture.273 This catalytic system also proved competent for direct CO2

reduction in the absence of ethanol, which proceeded via a homogeneous ruthenium-bound

formate species where formate coordination proved key to avoiding the thermodynamic

barrier for generation of a true formic acid intermediate. This adjustment yielded a

maximum TON of 442.274 While this was an encouraging improvement with respect to

TON and reaction simplicity, the system operated under acidic conditions that would be

incompatible with the bases often used in CO2 capture.

Page 92: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

76

Rather than using alcohols additives to generate intermediate esters, CO2

hydrogenation can also proceed through formamide intermediates when it is performed in

the presence of amines. In this reaction, the CO2 reacts with the amine to form a carbamate,

after which metal-catalyzed dehydration gives the formamide. Final metal-catalyzed

hydrogenation yields methanol and regenerates the amine. Sanford et al. were the first to

incorporate this strategy, using a ruthenium-pincer catalyst (Figure 1.3.1, (3)) with

dimethylamine and potassium phosphate additives to generate methanol from CO2 (2.5 bar)

and H2 (50 bar) with TONs of up to 550.275 However, in order to maximize catalytic

activity, the authors had to devise a temperature-ramping method (95 °C to 155 °C) in order

to by-pass the dimethylcarbamate (DMC) pathway in the initial step and instead activate

the formation of their dimethylformamide (DMF) intermediate from dimethylformic acid

(DMFA). This resulted in a product mixture containing methanol, DMF, and DMFA, of

which the latter two were the major contributors (together accounting for more than twice

the amount of methanol present).275 The comparatively high yields combined with needing

only very low pressures of CO2 was incredibly promising, however, selectivity

improvements would be required before more widespread applications would be possible.

Further developments in the incorporation of amines for facilitating CO2

hydrogenation to methanol continued to center on ruthenium catalysts. Prakash et al.

applied polyamines to the exact reaction parameters established by Sanford and were able

to attain a top TON of 1060, which is the highest performing catalytic system to date.276

They have also advanced techniques for CO2 capture from air and amine recycling using

polyamines and biphasic solvents.277 Milstein and co-workers established a unique

approach that operates under very low CO2 pressures (< 3 bar) through the use of

Page 93: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

77

Figure 1.3.1: Previous (1-5) and current metal catalysts for CO2 hydrogenation to

methanol via CO2 derivatives. OTf = triflate, OAc = acetate.

Page 94: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

78

aminoethanes to capture the CO2 as a heterocycle in situ, followed by ruthenium-catalyzed

hydrogenation (albeit with TONs < 25).278 The hydrogenation pressures, however, are still

around 60 bar H2.

The literature concerning metal-catalyzed hydrogenation of CO2 to methanol is

dominated by precious-metal ruthenium catalysts, whereas the incorporation of base metals

is far more limited. In fact, there are only two examples of base-metal catalysts for this

transformation, which include one manganese-pincer complex279 and one cobalt-triphos

complex (Figure 1.3.1., (4) and (5)).280 While their structures are highly reminiscent of

their ruthenium counterparts, their performances are not, yielding maximum TONs of 36

and 78 for the production of methanol using manganese and cobalt, respectively (the

former proceeds via the amide pathway while the latter uses the ester option). There is

therefore room for significant improvement for precious-metal and base-metal catalysts

alike. For comparison, top-performing catalysts for CO2 hydrogenation to formic acid or

formate include Nozaki’s precious-metal iridium-pincer catalyst281 with TONs of over 3.5

million and a base-metal iron-pincer catalyst with TONs of greater than 46,000 (when

paired with a Lewis acid co-catalyst).282

Previous work in our laboratories and others has established the reactivity of iron-

pincer complexes of the type (RPNP)Fe(H)(CO) and (RPNMeP)Fe(H)(CO)(BH4) (RPNP =

N[CH2CH2(PR2)]2, RPNMeP = (CH3)N[CH2CH2(PR2)]2; R = iPr, Cy) in a multitude of

hydrogenative and dehydrogenative transformations, including CO2 hydrogenation to

formate (the example which was mentioned above),282 formic acid dehydrogenation,283 and

methanol dehydrogenation.284 In all cases the catalytic capabilities of these complexes far

exceed the performance of any other base-metal catalyst for their respective processes,

Page 95: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

79

occasionally also surpassing the best precious-metal catalysts. It is therefore reasonable to

apply this class of compounds toward the problem of CO2 hydrogenation to methanol. In

addition, these catalysts have been successfully utilized by our laboratory in the N-

formylation of amines285 and the hydrogenation of formamides to amines and alcohols,286

so their catalytic aptitude with respect to the two separate steps of CO2 hydrogenation to

methanol in the presence of amines have also been established. This section summarizes

the ongoing investigation into the catalytic performance of the five-coordinate iron (II)

complex (iPrPNP)Fe(H)(CO) (1; Figure 1.3.2) with respect to the one-pot, two-step

synthesis of methanol from CO2 in the presence of morpholine (pathway (2) in Scheme

1.3.5). Even without being fully optimized, TONs of up to 360 have been attained for

methanol production from this reaction, representing an improvement of nearly half an

order of magnitude over other reported base-metal catalysts for CO2 hydrogenation.

1.3.2: Results and Discussion:

Initial studies in our laboratory with respect to 1-catalyzed CO2 hydrogenation to

methanol were conducted using dimethylamine287 due to its previous successful application

in this type of reaction.275 The production of methanol was not observed in the initial trials,

Figure 1.3.2: Structures of the iron-pincer complexes 1, 2, and 1-formate.

Page 96: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

80

which contained 21 mmol dimethylamine in 6 mL dioxane with 5 mol% 1, 100 psi CO2,

and 900 psi H2 in a Parr reactor at 120 °C for 16 hours. However, a large amount of DMF

was generated and the competency of 1 with respect to DMF hydrogenation had already

been established, suggesting that the reaction should be able to proceed.287 Upon further

investigation, two major obstacles for obtaining methanol were identified. The first was

that the water released during the conversion of CO2 to the intermediate formamide was

significant enough to accelerate catalyst decomposition. The second was the inhibition of

the formamide hydrogenation step due to formation of (iPrPNHP)Fe(H)(CO)(COOH), or 1-

formate (Figure 1.3.2; confirmed using NMR experiments). In the presence of CO2 and

H2, complex 1 undergoes a 1,2-addition of H2 across the Fe-N bond, followed quickly by

CO2 insertion into the Fe-H bond to give 1-formate,282 which is a poor catalyst for amide

hydrogenation.287 The formate can be released from the iron complex using a combination

of a Lewis acid (lithium trifluoromethanesulfonate, or LiOTf) and a base (DBU), but if a

large concentration of CO2 is still present then 1 will also catalyze the competing reaction

of CO2 hydrogenation to formate (for which it can achieve several thousand TONs).282 In

order to solve these problems, a one-pot, two-step reaction sequence was adopted (Scheme

1.3.6). The first step (step A) involved N-formylation of the amine in the presence of 3 Å

molecular sieves to absorb the water generated in situ (the efficacy of which was confirmed

by NMR experiments). After this reaction was completed, the Parr reactor was partially

vented, brought into the glovebox, and opened. Then, the reaction solution was filtered to

remove the sieve dust as well as any solid alkylammonium carbamate. In the absence of a

CO2 headspace, this carbamate salt decomposes back to CO2 and amine, which would act

as an undesirable source of CO2 in the second step. After filtration, a known amount of

Page 97: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

81

formamide (based on analysis of the step 1 solution) was loaded back into the Parr reactor

with more catalyst, along with LiOTf and DBU to regenerate 1 from any 1-formate formed

due to CO2 still dissolved in solution. This was followed by hydrogenation of the

formamide to give methanol and the free amine in the second step (step B).287

Despite these efforts, methanol was unable to be obtained with the use of

dimethylamine in the CO2 hydrogenation reaction.287 However, switching from

dimethylamine to morpholine immediately gave methanol TONs of greater than 250 using

the two-step procedure detailed above.287 Morpholine was recommended by its established

success as a substrate in both the N-formylation of amines and the hydrogenation of

amides.285,286 It was an attractive alternative to dimethylamine from a purification and

handling perspective due to the fact that it is a liquid at room temperature rather than a gas,

and its ring-constrained R groups had the potential to provide a steric advantage in

Scheme 1.3.6: Flow diagram for the one-pot, two-step reaction setup.

Page 98: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

82

accessing hindered reaction intermediates. The morpholine-derived ammonium carbamate

salt was also a solid under ambient conditions rather than a liquid (like that of

dimethylamine), affording greater separation from the formamide before step B. Efforts to

re-integrate the two reaction steps described above and to optimize the reaction conditions

using morpholine have been the current focus of this work.

Initial experiments sought to determine whether it was possible to re-establish a

one-step hydrogenation reaction using a more cooperative amine (Tables A1.15 – A1.18).

The absence of molecular sieves in step A improved the amount of available 4-

formylmorpholine after step 1 (Table A1.15, entry 2), which was unsurprising as some

absorption of the liquid formamide by the sieves would be expected. The methanol yield,

however, was significantly reduced, from approximately 65% to 11% (Table A1.15, entry

3). Despite observed consumption of the 4-formylmorpholine starting material, no

methanol was detected when sieves were present during step B, which was attributed to

their complete absorption of the product (Table A1.15, entry 4). This indicated that

molecular sieves must be present for step A but removed before step B. Considering the

possibility of using a different drying agent, LiOTf and DBU were also screened for their

compatibility between steps. Results indicated that while LiOTf was integral to the

performance of the second step, it detrimentally affected the yield of the first (Table

A1.16). The effect of DBU was found to be concentration-dependent (Table A1.17). The

presence of DBU alone did not harm either step (entry 3), however, if the overall amount

of DBU in solution was doubled, the yield of methanol began to decrease (entry 4). DBU

also caused practical issues with GC-FID analysis of the step A solution due to its affinity

for the column stationary phase, causing its signal to bleed into other parts of the spectrum

Page 99: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

83

and giving the appearance of reaction yield enhancement by its presence. Finally,

morphylammonium morphylcarbamate was independently synthesized (confirmed by

crystal structure, see Figure A1.52), isolated, and added to step B to see if the reaction

could proceed favorably in its presence (Table A1.18). It was found to have a significant

detrimental effect on methanol production. The combination of these results confirmed the

necessity of maintaining the two-step procedure, despite the amenability of morpholine

toward methanol production.

Two-step reaction optimization studies began with screening solvents (Table

A1.19), which had very little effect on methanol production, except for in the case of ethyl

acetate where an unidentified side product was observed by GC-FID. However, it was

indicated that tetrahydrofuran (THF) would be an acceptable substitute for 1,4-dioxane

(Table 1.3.1, entry 1 versus entry 2), which is an easier solvent to handle in terms of drying

procedures. It was therefore the solvent of choice in all future experiments. The methanol

yields from the solvent screen also suggested that a higher loading of 4-formylmorpholine

would be required in step B in order to truly asses the catalytic limitations of 1, since step

B essentially went to completion in all cases except ethyl acetate. Due to the fact that some

of the product is lost to the molecular sieves as well as during the filtration after step A, a

means of increasing that yield had to be devised. Increasing the ratio of CO2 in the gas

mixture (Table A1.20) afforded a moderate gain in product yield. Rather than continue to

increase CO2 levels, which is disadvantageous for step B, the overall reaction pressure was

increased from 1000 psi to 1400 psi and the formamide production was more than doubled

as a result (Table 1.3.1, entry 2A versus entry 3A). The formamide loading in step B could

then be increased from 1.25 mmol to 4 mmol, providing a larger window with which to

Page 100: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

84

fine-tune improvements. The amounts of the LiOTf and DBU additives were varied to

confirm that their concentrations were sufficient to handle the effects of a higher substrate

loading in step B, and it was found that no changes were needed (Table A1.21). It was also

confirmed that loading fresh catalyst solution in step B was necessary to acquire any

detectable methanol.

Entryd Substrate

(mmol)

CO2/H2

(psi)

Catalyst Solvent Temp.

(°C)

FM

(mmol)

CH3OH

(mmol)

TONc

1A Morph.

(21)

100/900 1 1,4-

dioxane

120 4.30 - 286

1B FM

(1.25)

0/900 1 1,4-

dioxane

120 0.03 1.14 114

2A Morph.

(21)

100/900 1 THF 120 4.42 - 294

2B FM

(1.25)

0/900 1 THF 120 0.01 1.13 113

3A Morph.

(21)

200/1200 1 THF 120 9.24 - 616

3B FM

(4.0)b

0/1200 1 THF 120 0.05 3.40 340

4A Morph.

(21)

200/1200 2 THF 120 3.20 - 213

4B FM

(4.0)b

0/1200 - - - - - -

5A Morph.

(21)

200/1200 1 THF 100 16.76 - 1117

5B FM

(4.0)b

0/1200 1 THF 100 0.01 3.60 360

Table 1.3.1: Select Reaction Optimization Data for CO2 Hydrogenation to Methanol in

the Presence of Morpholine.a

aReaction conditions: Step A: 21 mmol morpholine (morph.), 2.02 g molecular sieves, and 15 µmol

catalyst 1 or 2 in 8 mL solvent with X psi CO2 and Y psi H2 at Z °C for 16 hours. Step B: 1.25 mmol

formyl morpholine (FM) from step A, 1 mmol LiOTf, 2.5 mmol DBU in a total volume of 5 mL with Y

psi H2 at Z °C for 16 hours. b4.0 mmol FM in step B. cTON = turnover number for FM in step A and

CH3OH in step B as determined by GC-FID analysis. dAverage of two trials.

Page 101: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

85

Due to the necessity of a two-step reaction process, there was an opportunity for

employing different catalysts in steps A and B in order to maximize the productivity at

each stage. As mentioned briefly in the introduction, previous work in our laboratory has

used the iron pincer complexes 1 and 2 (Figure 1.3.2) for the N-formylation of amines,

where the performance of 2 exceeded that of 1 (attributed in part to its greater stability in

the presence of water). As water sensitivity has been identified as a limiting factor for

complex 1, complex 2 was utilized for N-formylation in first step, with the intention of

applying complex 1 to amide hydrogenation in the second step. However, it was found that

under these reaction conditions, the formamide yield was only a third of what it was when

complex 1 was used as the catalyst (Table 1.3.1, entry 3 versus entry 4). Further

investigations indicated that the presence of the molecular sieves interferes with the

catalytic ability of 2 (Table A1.22), which was tentatively ascribed to their slight acidity.

It is possible that this is also a contributing factor in the behavior of complex 1, as there is

a similar significant increase in formamide yield for step A in the absence of sieves (Table

A1.15) that was originally explained as a matter of product absorption by the additive.

More experiments will be required before a definitive understanding of this phenomenon

can be reached.

The final set of optimizations with respect to temperature found that a 20 °C

decrease nearly doubled the production of 4-formylmorpholine in step A and had no

detrimental effect on step B (Table 1.3.1, entry 3 versus entry 5). Lowering the temperature

further (80 °C) resulted in a slight decrease in yield (Table A1.23). This trend is attributed

to a balance between having a high enough temperature to favor the reaction kinetically

while avoiding thermally-induced catalyst degradation. The TON of 360 achieved for

Page 102: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

86

methanol production in step B under this milder set of temperature conditions represents

the highest activity of any base-metal catalyst for CO2 hydrogenation to methanol by nearly

half an order of magnitude.

1.3.4: Summary

In this section, the iron-pincer complex (iPrPNP)Fe(H)(CO) (complex 1) was

introduced, which belongs to a class of iron catalysts that display remarkable activity with

respect to hydrogenation and dehydrogenation reactions, especially those incorporating

CO2 (a particularly inactive substrate). The initial studies presented here revealed its

proficiency for the hydrogenation of CO2 to methanol in the presence of amines, and it far

outperformed any other base-metal catalyst for this transformation. Its best TON of 360

was also highly competitive with the reported activities of its precious-metal counterparts.

The promising catalytic capability of 1 therefore represents significant advancement

toward the goal of the more widespread utilization of CO2 as a carbon feedstock, since by

using CO2 as a direct source for methanol it also indirectly sources all of the materials that

can be made from methanol. The incorporation of abundant, less expensive base metals in

the catalysts for these processes also assists in maintaining the economic benefit and long-

term sustainability of transitioning to such a C1 carbon source. However, further

experiments will need to be conducted before the upper catalytic limit of complex 1 can be

determined. Unfortunately, the water sensitivity of 1 is a restrictive factor, and devising a

means of bypassing this without sacrificing catalytic performance is of the utmost

importance. The use of molecular sieves, while effective, prohibits the application of

structurally similar but more highly active catalytic systems for the first step of the

hydrogenation process, which implies a previously unrecognized influence on chemical

Page 103: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

87

reactivity that needs to be further elucidated. The investigation of other drying agents or

making adjustments to the step A processing methods (i.e. add sieves after step A is

complete and filter just before step B) may provide alternative (albeit more costly or

energy-intensive) solutions. Another factor to consider is the necessity of using one of the

highest total pressures reported for this hydrogenation reaction in order to maximize

conversion. Following the ester rather than the amide pathway through the use of alcohol

additives in place of amines may be more kinetically accessible and help to alleviate the

need for such forcing conditions. This would also avoid the formation of an ammonium

carbamate, which can serve as a masked CO2 source that is detrimental to step B catalysis.

The impressive performance of 1 in the face of these challenges, however, suggests at a

much greater potential once they have been overcome.

Page 104: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

88

1.4: Chapter Summary

In this chapter, the catalytic competency of nickel and iron complexes with respect

to the activation of CO2 and CO2 analogues and their subsequent conversion into valuable

commercial chemicals was explored. These transformations were enacted using either

alkenes or dihydrogen as highly reactive co-reagents in order to make the reactions more

energetically favorable. In the first two sections, it was found that the encouraging

stoichiometric coupling performances of nickel promotors in CO2/alkene/CO and

isocyanate/alkene reactions did not translate to product selectivity, which would need to be

overcome (likely via systematic ligand variation) before they could be extended to catalytic

investigations. The last section of this chapter detailed ongoing studies regarding the

promising ability of the iron pincer complex (iPrPNP)Fe(H)(CO) (complex 1) in the

hydrogenation of CO2 to methanol. Even without final optimization of the reaction

conditions, complex 1 achieves TONs that are nearly five times those of any other base-

metal catalyst and that match or surpass those of most of its precious metal competition for

the production of methanol. This performance therefore denotes a significant improvement

in catalytic capability, the importance of which rests in the opportunity the realization of

this transformation represents for sourcing not only methanol but the multitude of products

it engenders from a cheap, abundant, and renewable carbon source. That being said,

addressing the issues of the water-sensitivity of 1 (necessitating a two-step reaction

sequence) and the use of reaction pressures on the higher end of those reported to achieve

the best TONs will likely be needed before its more widespread application is realized.

Overall, while the results presented in this chapter are modest, the reactions they

address are inherently challenging. For example, nickel-catalyzed production of acrylate

Page 105: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

89

from CO2/ethylene coupling has only recently been attained after over 40 years of study,

with TONs on the order of ~100.117 Understanding the role of the base, Lewis acid, and

zinc additives that made this catalysis possible will likely impact the field as a whole and

help to advance analogous nickel-based reactions, like those discussed herein. With only a

handful of examples that have achieved at best 1000 turnovers, the metal-catalyzed

hydrogenation of CO2 to methanol is a similarly non-trivial transformation. A pivotal issue

is the beneficial impact of high CO2 content on the first half of the reaction but its inhibition

of the final hydrogenation step. Evolving methods of using low CO2 pressures but still

maintaining high overall yields will likely be key to making this reaction more competitive

with other CO2 hydrogenation/dehydrogenation cycles.

Page 106: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

90

1.5 References

1 REN21.2017, Renewables 2017 Global Status Report, Renewable Energy Policy Network for the 21st

Century. http://www.ren21.net/wp-content/uploads/2017/06/17-

8399_GSR_2017_Full_Report_0621_Opt.pdf

2 Houghton, J. Global Warming, 4th ed., Cambridge, U.K.: Cambridge Univ. Press, 2009.

3 Olah, G.A.; Goeppert, A.; Prakash, G.K.S. “Chemical Recycling of Carbon Dioxide to Methanol and

Dimethyl Ether: From Greenhouse Gas to Renewable, Environmentally Carbon Neutral Fuels and Synthetic

Hydrocarbons,” J. Org. Chem., 2009, 74, 487-498.

4 Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.;

Hille, R.; Kenis, P.J.A.; Kerfeld, C.A.; Morris, R.H.; Peden, C.H.F.; Portis, A.R.; Ragsdale, S.W.; Rauchfuss,

T.B.; Reek, J.N.H.; Seefeldt, L.C.; Thauer, R.K.; Waldrop, G.L. “Frontiers, Opportunities, and Challenges

in Biochemical and Chemical Catalysis of CO2 Fixation,” Chem. Rev., 2013, 113, 6621-6658.

5 Xiaoding, X.; Moulijn, J.A. “Mitigation of CO2 by Chemical Conversion: Plausible Chemical Reactions

and Promising Products,” Energy and Fuels, 1996, 10, 305-325.

6 Riduan, S.N.; Zhang, Y. “Recent developments in carbon dioxide utilization under mild conditions,” Dalton

Trans., 2010, 39, 3347-3357.

7 https://www.co2.earth/daily-co2

8 BP Statistical Review of World Energy, June 2017.

https://www.bp.com/content/dam/bp/en/corporate/pdf/energy-economics/statistical-review-2017/bp-

statistical-review-of-world-energy-2017-full-report.pdf

9 Huang, K.; Sun, C.-L.; Shi, Z.-J. “Transition-metal-catalyzed C-C bond formation through the fixation of

carbon dioxide,” Chem. Soc. Rev., 2011, 40, 2435-2452.

10 Cokoja, M.; Bruckmeier, C.; Rieger, B.; Herrmann, W.A.; Kühn, F.E. “Transformation of Carbon Dioxide

with Homogeneous Transition-Metal Catalysis: A Molecular Solution to a Global Challenge?” Angew. Chem.

Int. Ed., 2011, 50, 8510-8537.

11 Sakakura, T.; Kohno, K. “The synthesis of organic carbonates from carbon dioxide,” Chem. Commun.,

2009, 1312-1330.

12 Sakakura, T.; Choi, J.-C.; Yasuda, H. “Transformation of Carbon Dioxide,” Chem. Rev., 2007, 107, 2365-

2387.

13 Arakawa, H.; Aresta, M.; Armor, J.N.; Barteau, M.A.; Beckman, E.J.; Bell, A.T.; Bercaw, J.E.; Creutz, C.;

Dinjus, E.; Dixon, D.A.; et al. “Catalysis Research of Relevance to Carbon Management: Progress,

Challenges, and Opportunities,” Chem. Rev., 2001, 101, 953-996.

14 Aresta, M.; Dibenedetto, A. “Utilisation of CO2 as a chemical feedstock: opportunities and challenges,”

Dalton Trans., 2007, 2975-2992.

Page 107: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

91

15 Quadrelli, E.A.; Centi, G.; Duplan, J.-L.; Perathoner, S. “Carbon Dioxide Recycling: Emerging Large-

Scale Technologies with Industrial Potential,” ChemSusChem, 2011, 4, 1194-1215.

16 Aresta, M., Carbon Dioxide: Utilization Options to Reduce its Accumulation in the Atmosphere. In Carbon

Dioxide as Chemical Feedstock, 1st Ed., Wiley: Hoboken N.J., 2010.

17 Gibson, D.H. “The Organometallic Chemistry of Carbon Dioxide,” Chem. Rev., 1996, 96, 2063-2095.

18 Pearson, R.L.; Eisaman, M.D.; Turner, J.W.G.; Edwards, P.P.; Jiang, Z.; Kuznetsov, V.L.; Littau, K.A.;

Di Marco, L.; Taylor, S.R.G. “Energy Storage via Carbon-Neutral Fuels Made from CO2, Water, and

Renewable Energy,” Proceedings of the IEEE, 2012, 100, 440-460.

19 Aresta, M.; Dibenedetto, A.; Angelini, A. “Catalysis for the Valorization of Exhaust Carbon: from CO2 to

Chemicals, Materials, and Fuels. Technological use of CO2,” Chem. Rev., 2014, 114, 1709-1742.

20 Fukuoka, S.; Kawamura, M.; Komiya, K.;Tojo, M.; Hachiya, H.; Hasegawa, K.; Aminaka, M.; Okamoto,

H.; Fukawa, I.; Konno, S. “A novel non-phosgene polycarbonate production process using by-product CO2

as starting material,” Green Chem., 2003, 5, 497-507.

21 Eghbali, N.; Li, C.-J. “Conversion of carbon dioxide and olefins into cyclic carbonates in water,” Green

Chem., 2007, 9, 213-215.

22 Mazzotti, M.; Abanades, J.C.; Allam, R.; Lackner, K.S.; Meunier, F.; Rubin, E.; Sanchez, J.C.; Yogo, K.;

Zevenhoven, R. “Mineral carbonation and industrial uses of carbon dioxide,” IPCC Special Report on

Carbon dioxide Capture and Storage,” Ch.7, pp. 321-335.

23 Lindsey, A.S.; Jeskey, H. “The Kolbe-Schmitt Reaction,” Chem. Rev., 1957, 57, 583-620.

24 Stuart, M.C.; Kouimtzi, M.; Hill, S.R. WHO Model Formulary 2008; WHO Press, World Health

Organization: Geneva, Switzerland, 2009, 1-628.

25 Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies, F.;

Kurr, P.; Kniep, B.-L.; Tovar, M.; Fischer, R.S.; Nørskov, J.S.; Schlögl, R. “The Active Site of Methanol

Synthesis over Cu/ZnO/Al2O3 Industrial Catalysts,” Science, 2012, 336, 893-897.

26 Cañete, B.; Gigola, C.E.; Brignole, N.B. “Synthesis Gas Processes for Methanol Production via CH4

Reforming with CO2, H2O, and O2,”Ind. Eng. Chem. Res., 2014, 53, 7103-7112.

27 IHS Markit Report, 2015, http://www.methanol.org/wp-content/uploads/2016/07/Marc-Alvarado-Global-

Methanol-February-2016-IMPCA-for-upload-to-website.pdf

28 IHS Chemical Bulletin, 2016, http://www.methanol.org/wp-content/uploads/2016/07/IHS-

ChemicalBulletin-Issue3-Alvarado-Jun16.pdf

29 Olah, G.; Goeppert, A.; Prakash, G.K.S. Beyond Oil and Gas: The Methanol Economy; VCH, Weinheim:

2006; Chapter 10-13, pp. 168-253.

30 Liu, Q.; Wu, L.; Jackstell, R.; Beller, M. “Using carbon dioxide as a building block in organic synthesis,”

Nat. Commun., 2015, 6, 5933, 1-15.

31 Jitaru, M. “ELECTROCHEMICAL CARBON DIOXIDE REDUCTION- FUNDAMENTAL AND

APPLIED TOPICS,” Journal of the University of Chemical Technology and Metallurgy, 2007, 4, 333-344.

32 Kang, P.; Chen, Z.; Brookhart, M.; Meyer, T.J. “Electrocatalytic Reduction of Carbon Dioxide: Let the

Molecules Do the Work,” Top Catal., 2015, 58, 30-45.

Page 108: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

92

33 Vayenas, C.G., White, R.E., Gamboa-Aldeco, M.E., Eds. Modern Aspects of Electrochemistry; Springer:

New York, 2008; Volume 42.

34 Slater, S.; Wagenknecht, J.H. “Electrochemical Reduction of CO2 Catalyzed by Rh(diphos)2Cl,” J. Am.

Chem. Soc., 1984, 106, 5367-5368.

35 Ishida, H.; Tanaka, H.; Tanaka, K.; Tanaka, T. “J. Chem. Soc., Chem. Commun., 1987, 131-132.

36 Rail, M.D.; Berben, L.A. “Directing the Reactivity of [HFe4N(CO)12]- Toward H+ or CO2 Reduction by

Understanding the Electrocatalytic Mechanism,” J. Am. Chem. Soc., 2011, 133, 18577-18579.

37 Arana, C.; Yan, S.; Keshavarz-K, M.; Potts, .K.T.; Abruña, H.D. “Electrocatalytic Reduction of Carbon

Dioxide with Iron, Cobalt, and Nickel Complexes of Terdentate Ligands,” Inorg. Chem., 1992, 31, 3680-

3682.

38 Kang, P.; Cheng, C.; Chen, Z.; Schauer, C.K.; Meyer, T.J.; Brookhart, M. “Selective Electrocatalytic

Reduction of CO2 to Formate by Water-Stable Iridium Dihydride Pincer Complexes,” J. Am. Chem. Soc.,

2012, 134, 5500-5503.

39 Ahn, S.T.; Bielinski, E.A.; Lane, E.M.; Chen, Y.; Bernskoetter, W.H.; Hazari, N.; Palmore, G.T.R.

“Enhanced CO2 electroreduction efficiency through secondary coordination effects on a pincer iridium

catalyst,” Chem. Commun., 2015, 51, 5947-5950.

40 Behr, A. “Carbon Dioxide as an Alternative C1 Synthetic Unit: Activation by Transition-Metal

Complexes,” Angew. Chem. Int. Ed., 1988, 27, 661-678.

41 Leitner, W. “The coordination chemistry of carbon dioxide and its relevance for catalysis: a critical

survey,” Coord. Chem. Rev., 1996, 153, 257-284.

42 Yin, X.; Ross, J.R. “Recent developments in the activation of carbon dioxide by metal complexes,” Coord.

Chem. Rev., 1999, 181, 27-59.

43 Palmer, D.A.; van Eldik, R. “The Chemistry of Metal Carbonato and Carbon Dioxide Complexes,” Chem.

Rev., 1983, 83, 651-731.

44 Boddien, A.; Gärtner, F.; Federsel, C.; Piras, I.; Junge, H.; Jackstell, R.; Beller, M., Catalytic Utilization

of Carbon Dioxide: Actual Status and Perspectives, in Organic Chemistry: Breakthroughs and Perspectives;

Ding, K., Dai, L.-X., Eds.; Wiley-VCH: Weinheim, Germany, 2012, pp. 685-724.

45 Pápai, I.; Schubert, G.; Mayer, I.; Besenyei, G.; Aresta, M. “Mechanistic Details of Nickel(0)-Assisted

Oxidative Coupling of CO2 with C2H4,” Organometallics, 2004, 23, 5252-5259.

46 Takimoto, M.; Nakamura, Y.; Kimura, K.; Mori, M. “Highly Enantioselective Catalytic Carbon Dioxide

Incorporation Reaction: Nickel-Catalyzed Asymmetric Carboxylative Cyclization of Bis-1,3-dienes,” J. Am.

Chem. Soc., 2004, 126, 5956-5957.

47 Yeung, C.S.; Dong, V.M. “Beyond Aresta’s Complex: Ni- and Pd-Catalyzed Organozinc Coupling with

CO2,” J. Am. Chem. Soc., 2008, 130, 7826-7827.

48 Aresta, M.; Nobile, C.F.; Albano, V.G.; Forni, E.; Manassero, M. “New Nickel-Carbon Dioxide Complex:

Synthesis, Properties, and Crystallographic Characterization of (Carbon dioxide)-

bis(tricyclohexylphosphine)nickel,” J. Chem. Soc., Chem. Commun., 1975, 636-637.

49 Tolman, W.B., “Carbon Dioxide Reduction and Uses as a Chemical Feedstock. Activation of Small

Molecules: Organometallic and Bioinorganic Perspectives,” Wiley-VCH: Weinheim, 2006, pp. 1-35.

Page 109: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

93

50 Bernskoetter, W.H.; Tyler, B.T. “Kinetics and Mechanism of Molybdenum-Mediated Acrylate Formation

from Carbon Dioxide and Ethylene,” Organometallics, 2011, 30, 520-527.

51 Sasaki, Y.; Inoue, Y.; Hashimoto, H. “Reaction of Carbon Dioxide with Butadiene Catalysed by Palladium

Complexes. Synthesis of 2-Ethylidenehept-5-en-4-olide,” J. Chem. Soc., Chem. Commun., 1976, 605-606.

52 Musco, A.; Perego, C.; Tartiari, V. “Telomerization reactions of butadiene and CO2 catalyzed by phosphine

Pd(0) complexes: (E)-2-ethylidenehept-6-en-5-olide and octadienyl esters of 2-ethylidenehepta-4,6-dienoic

acid,” Inorg. Chim. Acta, 1978, 28, L147-L148.

53 Braunstein, P.; Matt, D.; Nobel, D. “Reactions of Carbon Dioxide with Carbon-Carbon Bond Formation

Catalyzed by Transition-Metal Complexes,” Chem. Rev., 1988, 88, 747-764.

54 Walther, D. “Homogeneous-catalytic reactions of carbon dioxide with unsaturated substrates, reversible

CO2-carriers and transcarboxylation reactions,” Coord. Chem. Rev., 1987, 79, 135-174.

55 Wilke, G. “Cyclooligomerization of Butadiene and Transition Metal π-Complexes,” Angew. Chem. Int.

Ed., 1963, 2, 105-115.

56 Behr, A.; Juszak, K.-D.; Keim, W. “Synthese von 2-Ethyliden-6-hepten-5-olid,” Synthesis, 1983, 574.

57 Behr, A.; Juszak, K.-D. “Palladium-catalyzed reaction of butadiene and carbon dioxide,” J. Organomet.

Chem., 1983, 255, 263-268.

58 Döhring, A.; Jolly, P.W. “The palladium catalyzed reaction of carbon dioxide with allene,” Tetrahedron

Lett., 1980, 21, 3021-3024.

59 Behr, A.; Becker, M.; Beckmann, T.; Johnen, L.; Leschinski, J.; Reyer, S. “Telomerization: Advances and

Applications of a Versatile Reaction,” Angew. Chem. Int. Ed., 2009, 48, 3598-3614.

60 Braunstein, P.; Matt, D.; Nobel, D. “Carbon dioxide activation and catalytic lactone synthesis by

telomerization of butadiene and carbon dioxide,” J. Am. Chem. Soc., 1988, 110, 3207-3212.

61 Hoberg, H.; Gross, S.; Milchereit, A. “Nickel(0)-Catalyzed Production of a Functionalized

Cyclopentanecarboxylic Acid from 1,3-butadiene and CO2,” Angew. Chem. Int. Ed., 1987, 26, 571-572.

62 Behr, A.; He, R. “Rhodium catalyzed reaction of butadiene and carbon dioxide,” J. Organomet. Chem.,

1984, 276, C69-C72.

63 Inoue, Y.; Itoh, Y.; Hashimoto, H. “Incorporation of carbon dioxide in alkyne oligomerization catalyzed

by nickel (0) complexes. Formation of substituted 2-pyrones,” Chem. Lett., 1977, 6, 855-856.

64 Albano, P.; Aresta, M. “Some catalytic properties of Rh(diphos)(η-BPh4),” J. Organomet. Chem., 1980,

190, 243-246.

65 Inoue, Y.; Itoh, Y.; Hashimoto, H. “Oligomerization of 3-Hexyne by nickel(0) complexes under CO2.

Incorporation of CO2 and novel cyclotrimerization,” Chem. Lett., 1978, 7, 633-634.

66 Walther, D.; Schönberg, H.; Dinjus, E.; Sieler, J. “Aktivierung von Kholendioxid an

Übergangsmetallzentren: Selektiv Cooligomerisation mit Hexin(-3) durch das Katalysatorsystem

Acetonitril/Trialkylphosphan/Nickel(0) und Struktur eines Nickel(0)-Komplexes mit side-on gebundenem

Acetonitril,” J. Organomet. Chem., 1987, 334, 377-388.

67 Louie, J.; Gibby, J.E.; Farnworth, M.V.; Tekavec, T.N. “Efficient Nickel-Catalyzed [2+2+2] Cycloaddition

of CO2 and Diynes,” J. Am. Chem. Soc., 2002, 124, 15188-15189.

Page 110: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

94

68 Dérien, S.; Duñach, E.; Périchon, J. “From Stoichiometry to Catalysis: Electroreductive Coupling of

Alkynes and Carbon-Dioxide with Nickel-Bipyridine Complexes. Magnesium Ions as the Key for Catalysis,”

J. Am. Chem. Soc., 1991, 113, 8447-8454.

69 Hoberg, H.; Schaefer, D. “Oxanickelacyclopenten-derivate, ein neuer typ vielseitig

verwendbarersynthone,” J. Organomet. Chem., 1982, 228, C21-C24.

70 Hoberg, H.; Schaefer, D. “Modellkomplexe des nickels für die [2+2+2’]-cycloaddition von alkinen mit

kohlendioxid,” J. Organomet. Chem., 1982, 238, 383-387.

71 Saito, S.; Nakagawa, S.; Koizumi, T.; Hirayama, K.; Yamamoto, Y. “Nickel-Mediated Regio- and

Chemoselective Carboxylation of Alkynes in the Presence of Carbon Dioxide,” J. Org. Chem., 1999, 64,

3975-3978.

72 Takimoto, M.; Mizuno, T.; Mori, M.; Sato, Y. “Nickel-mediated cyclization of enynes under and

atmosphere of carbon dioxide,” Tetrahedron, 2006, 62, 7589-7597.

73 Burkhart, V.G.; Hoberg, H. “Oxanickelacyclopenten-Derivate aus Nickel(0), Kohlendioxid und Alkinen,”

Angew. Chem., 1982, 94, 75.

74 Langer, J.; Görls, H.; Walther, D. “Nickelacyclic carboxylates derived from 3-hexyne and CO2 and their

applications in the synthesis of a new muconic acid derivative,” Polyhedron, 2012, 32, 60-67.

75 Inoue, Y.; Hibi, T.; Satake, M.; Hashimoto, H. “Reaction of Methylenecyclopropanes with Carbon Dioxide

Catalysed by Palladium(0) Complexes. Synthesis of Five-Membered Lactones,” J. Chem. Soc., Chem.

Commun., 1979, 982.

76 Binger, P.; Weintz, H.-J. “Reaktionen der Methylencyclopropane, VIII. Palladium(0)-katalysierte

Darstellung ungesättigter γ-Lactone aus Methylencyclopropan und Kohlendioxid,” Chem. Ber., 1984, 177,

654-665.

77 Hoberg, H.; Schaefer, D. “Nickel(0)-Induzierte C-C-Verknüpfung Zwischen Kholendioxid und Ethylen

Sowie Mono-Oder Di-Substituierten Alkenen,” J. Organomet. Chem., 1983, 251, C51-C53.

78 Hoberg, H.; Peres, Y.; Krüger, C.; Tsay, Y.-H. “A 1-Oxa-2-nickela-5-cyclopentanone from Ethene and

Carbon Dioxide: Preparation, Structure, and Reactivity,” Angew. Chem. Int. Ed., 1987, 26, 771-773.

79 Hoberg, H; Jenni, K.; Angermund, K.; Krüger, C. “CC-Linkages of Ethene with CO2 on an Iron(0)

Complex- Synthesis and Crystal Structure Analysis of [(PEt3)2Fe(C2H4)2],” Angew. Chem. Int. Ed., 1987, 26,

153-155.

80 Hoberg, H.; Schaefer, D. “Nickel(0)-Induzierte C-C-Verknüpfung Zwischen Alkenen und Kohlendioxid,”

J. Organomet. Chem., 1982, 236, C28-C30.

81 Hoberg, H.; Ballesteros, A.; Sigan, A.; Jegat, C.; Milchereit, A. “Durch (Lig)Ni(0) induzierte Herstellung

von mono- und di-Carbonsäuren aus Cyclopenten und Kohlendioxid,” Synthesis, 1991, 395-398.

82 Walther, D.; Dinjus, E.; Sieler, J.; Anderson, L.; Lindqvist, O. “Aktivierung von Kohlendioxid an

Übergangsmetallzentren: Metallaringschluss mit Dicyclopentadien am Elektronenreichen Nickel(0)-

Komplexrumpf als Topo- und Stereoselektive Reaktion,” J. Organomet. Chem., 1984, 276, 99-107.

83 Hoberg, H.; Oster, B.W. “Nickel(0)-Induzierte C-C-Verknüpfung Zwischen 1,2-Dienen und

Kohlendioxid,” J. Organomet. Chem., 1984, 266, 321-326.

84 Hoberg, H.; Jenni, K.; Krüger, C.; Raabe, E. “CC Coupling of CO2 and Butadiene on Iron(0) Complexes-

A Novel Route to α,ω-Dicarboxylic Acids,” Angew. Chem. Int. Ed., 1986, 25, 810-811.

Page 111: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

95

85 Hoberg, H.; Peres, Y.; Milchereit, A. “C-C-Verknüpfung von Alkenen mit CO2 an Nickel(0); n-

Pentensäuren aus Ethen,” J. Organomet. Chem., 1986, 307, C41-C43.

86 Langer, J.; Fischer, R.; Görls, H.; Walther, D. “Low-Valent Nickel and Palladium Complexes with 1,1’-

Bis(phosphanyl)-ferrocenes: Syntheses and Structures of Acrylic Acid and Ethylene Complexes,” Eur. J.

Inorg. Chem., 2007, 2257-2264.

87 George, M., I.O.C.L., Acrylic Acid Overview in Global Propylene and Derivatives Summit, 2014.

88 Herth, G.; Schornick, G.; Buchholz, F.L. Polyacrylamides and Poly(acrylic) acids. In Ullmann’s

Encyclopedia of Industrial Chemistry; Wiley-VCH: Verlag, Germany, 2002; pp. 1-16.

89 Ohara, T.; Sato, T.; Shimizu, N.; Prescher, G.; Schwind, H.; Weiberg, O.; Marten, K.; Greim, H. Acrylic

Acid and Derivatives. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Verlag, Germany,

2002; pp. 1-18.

90 Herrmann, W.A. In Applied Homogeneous Catalysis with Organometallic Compounds; Cornils, B.,

Herrmann, W.A., Eds.; 2nd ed.; Wiley-VCH: Weinheim, Germany, 2002, pp. 1297-1304.

91 Arntz, D.; Fischer, A.; Höpp, M.; Jacobi, S.; Sauer, J.; Ohara, T.; Sato, T.; Shimizu, N.; Schwind, H.

Acrolein and Methacrolein. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Verlag,

Germany, 2002; pp. 1-18.

92 Alvarez, R.; Carmona, E.; Cole-Hamilton, D.-J.; Galindo, A.; Gutiérrez-Puebla, E.; Monge, A.; Poveda,

M.L.; Ruiz, C. “Formation of Acrylic Acid Derivatives from the Reaction of CO2 and Ethylene Complexes

of Molybdenum and Tungsten,” J. Am. Chem. Soc., 1985, 107, 5529-5531.

93 Alvarez, R.; Carmona, E.; Galindo, A.; Gutiérrez, E.; Marín, J.M.; Monge, A.; Poveda, M.L.; Ruiz, C.;

Savariault, J.M. “Formation of Carboxylate Complexes from the Reaction of CO2 with Ethylene Complexes

of Molybdenum and Tungsten. X-ray and Neutron Diffraction Studies,” Organometallics, 1989, 8, 2430-

2439.

94 Galindo, A.; Pastor, A.; Pérez, P.J.; Carmona, E. “Bis(ethylene) Complexes of Molybdenum and Tungsten

and Their Reactivity toward CO2. New Examples of Acrylate Formation by Coupling of Ethylene and Carbon

Dioxide,” Organometallics, 1993, 12, 4443-4451.

95 Collazo, C.; del Mar Conejo, M.; Pastor, A.; Galindo, A. “Synthesis and reactivity of bis(ethylene)-

phosphite complexes of molybdenum (0),” Inorg. Chim. Acta., 1998, 272, 125-130.

96 Graham, D.C.; Mitchell, C.; Bruce, M.I.; Metha, G.F.; Bowie, J.H.; Buntine, M.A. “Production of Acrylic

Acid through Nickel-Mediated Coupling of Ethylene and Carbon Dioxide-A DFT Study,” Organometallics,

2007, 26, 6784-6792.

97 Dedieu, A.; Ingold, F. “Monohapto vs Dihapto CO2 Coordination in Bis(amine)Ni0 Complexes: A CAS-

SCF Study,” Angew. Chem. Int. Ed., 1989, 28, 12-13.

98 Sakaki, S.; Mine, K.; Taguchi, D.; Arai, T. “Formation of the Oxanickelacyclopentene Complex from

Nickel(0), Carbon Dioxide, and Alkyne. An ab initio MO/SD-CI Study,” Bull. Chem. Soc. Jpn., 1993, 66,

3289-3299.

99 Sakaki, S.; Mine, K.; Hamada, T.; Arai, T. “Formation of the Oxanickelacyclopentene Complex from

Nickel(0), Carbon Dioxide, and Alkyne. An ab initio MO/SD-CI Study. Part II. Reactivity and

Regioselectivity of Hydroxyacetylene,” Bull. Chem. Soc. Jpn., 1995, 68, 1873-1882.

Page 112: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

96

100 Schubert, Ǵ.; Pápai, I. “Acrylate Formation via Metal-Assisted C-C Coupling between CO2 and C2H4:

Reaction Mechanism as Revealed from Density Functional Calculations,” J. Am. Chem. Soc., 2003, 125,

14847-14858.

101 Wolfe, J.M.; Bernskoetter, W.H. “Reductive functionalization of carbon dioxide to methyl acrylate at

zerovalent tungsten,” Dalton Trans., 2012, 41, 10763-10768.

102 Hoberg, H.; Ballesteros, A.; Sigan, A.; Jégat, C.; Bärhausen, D.; Milchereit, A. “Ligandgesteuerte

Ringkontraktion von Nickela-Fünf-in Vierringkomplexe- neuartige Startsysteme für die präparative

Chemie,” J. Organomet. Chem., 1991, 407, C23-C29.

103 Hoberg, H.; Ballesteros, A. “Ni0-induzierte Herstellung cyclischer C8-Carbonsäuren aus Cyclooctenen

und Kohlendioxid,” J. Organomet. Chem., 1991, 411, C11-C18.

104 Plessow, P.N.; Schäfer, A.; Limbach, M.; Hofmann, P. “Acrylate Formation from CO2 and Ethylene

Mediated by Nickel Complexes: A Theoretical Study,” Organometallics, 2014, 33, 3657-3668.

105 Hipler, B.; Döring, M.; Dubs, C.; Görls, H.; Hübler, T.; Uhlig, E. “Bildung ung Strukturen von

Nickelacyclen des Typs (LL’)NiCH2CH2C(O)O,” Z. anorg. allg. Chem., 1998, 624, 1329-1335.

106 Walther, D.; Liesicke, S.; Fischer, R.; Görls, H.; Weston, J. Batista, A. “2,3-Bis(diarylphosphanyl)-1,4-

diazadienes: P,P Coordination of PdII, PtII, and a Nickelacyclopentanone with Subsequent Formation of

Quinoxalines by a Ring-Closure Reaction at the Periphery,” Eur. J. Inorg. Chem., 2003, 4321-4331.

107 Brookhart, M.; Green, M.L.H.; Parkin, G. “Agostic interactions in transition metal compounds,” Proc.

Natl. Acad. Sci. U.S.A., 2007, 104, 6908-6914.

108 Langford, C.H.; Gray, H.B. Ligand Substitution Processes; W.A. Benjamin, Inc.: New York, 1996.

109 Fischer, R.; Langer, J.; Malassa, A.; Walther, D.; Görls, H.; Vaughan, G. “A key step in the formation of

acrylic acid from CO2 and ethylene: the transformation of a nickelalactone into a nickel-acrylate complex,”

Chem. Commun., 2006, 2510-2512.

110 Bruckmeier, C.; Lehenmeier, M.W.; Reichhardt, R.; Vagin, S.; Rieger, B. “Formation of Methyl Acrylate

from CO2 and Ethylene via Methylation of Nickelalactones,” Organometallics, 2010, 29, 2199-2202.

111 Lee, S.Y.T.; Cokoja, M.; Drees, M.; Li, Y.; Mink, J.; Herrmann, W.A.; Kühn, F.E. “Transformation of

Nickelalactones to Methyl Acrylate: On the Way to a Catalytic Conversion of Carbon Dioxide,”

ChemSusChem, 2011, 4, 1275-1279.

112 Plessow, P.N.; Weigel, L.; Lindner, R.; Schäfer, A.; Rominger, F.; Limbach, M.; Hofmann, P.

“Mechanistic Details of the Nickel-Mediated Formation of Acrylates from CO2, Ethylene, and Methyl

Iodide,” Organometallics, 2013, 32, 3327-3338.

113 Lejkowski, M.L.; Lindner, R.; Kageyama, T.; Bódizs, G.E.; Plessow, P.N.; Müller, I.B.; Schäfer, A.;

Rominger, F.; Hofmann, P.; Futter, C.; Schunk, S.A.; Limbach, M. “The First Catalytic Synthesis of an

Acrylate from CO2 and an Alkene- A Rational Approach,” Chem. Eur. J., 2012, 18, 14017-14025.

114 Jin, D.; Schmeier, T.J.; Williard, P.G.; Hazari, N.; Bernskoetter, W.H. “Lewis Acid Induced β-H

Elimination from a Nickelalactone: Efforts toward Acrylate Production from CO2 and Ethylene,”

Organometallics, 2013, 32, 2152-2159.

115 Jin, D.; Williard, P.G.; Hazari, N.; Bernskoetter, W.H. “Effect of Sodium Cation on Metallacycle β-H

Elimination in CO2-Ethylene Coupling to Acrylates,” Chem. Eur. J., 2014, 20, 3205-3211.

Page 113: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

97

116 Hendriksen, C.; Pidko, E.A.; Yang, G.; Schäffner, B.; Vogt, D. “Catalytic Formation of Acrylate from

Carbon Dioxide and Ethene,” Chem. Eur. J., 2014, 20, 12037-12040.

117 Huguet, N.; Jevtovikj, I.; Gordillo, A.; Lejkowski, M.L.; Lindner, R.; Bru, M.; Khalimon, A.Y.; Rominger,

F.; Schunk, S.A.; Hofmann, P.; Limbach, M. “Nickel-Catalyzed Direct Carboxylation of Olefins with CO2:

One-Pot Synthesis of α,β-Unsaturated Carboxylic Acid Salts,” Chem. Eur. J., 2014, 20, 16858-16862.

118 Mazini, S.; Huguet, N.; Trapp, O.; Schaub, T. “Palladium- and Nickel-Catalyzed Synthesis of Sodium

Acrylate from Ethylene, CO2, and Phenolate Bases: Optimization of the Catalytic System for a Potential

Process,” Eur. J. Org. Chem., 2015, 7122-7130.

119 Gu, Y.-F.; Chen, Y.; Hammond, E.G. “Use of cyclic anhydrides to remove cholesterol and other hydroxyl

compounds from fats and oils,” J. Am. Oil Chem. Soc., 1994, 71, 1205-1209.

120 Yamaguchi, T.; Yanagi, T.; Hokari, H.; Mukaiyama, Y.; Kamijo, T.; Yamamoto, I. “Preparation of

Optically Active Succinic Acid Derivatives. II. Efficient and Practical Synthesis of KAD-1229,” Chem.

Pharm. Bull., 1998, 46, 337-340.

121 https://www.epa.gov/sites/production/files/2016-09/documents/maleic-anhydride.pdf

122 Kraft, P.; Swift, K.A.D., Eds. Perspectives in Flavor and Fragrance Research, Wiley-VCH: Weinheim,

Germany and VHCA: Zürich, Switzerland, 2005.

123 https://thechemco.com/chemical/phthalic-anhydride/

124 Lott, S. Phthalate-free Plasticizers in PVC. In Healthy Building Network, 2014.

125 Bailey, D.M.; Johnson, R.E. “Reduction of Cyclic Anhydrides with NaBH4. Versatile Lactone Synthesis,”

J. Org. Chem., 1970, 35, 3574-3576.

126 Manzano, R.; Andrés, J.M.; Muruzábal, M.-D.; Pedrosa, R. “Synthesis of both Enantiomers of Hemiesters

by Enantioselective Methanolysis of Meso Cyclic Anhydrides Catalyzed by α-Amino Acid-Derived Chiral

Thioureas,” J. Org. Chem., 2010, 75, 5417-5420.

127 DiCiccio, A.; Longo, J.M.; Rodríguez-Calero, G.G.; Coates, G.W. “Development of Highly Active and

Regioselective Catalysts for the Copolymerization of Epoxides with Cyclic Anhydrides: An Unanticipated

Effect of Electronic Variation,” J. Am. Chem. Soc., 2016, 138, 7107-7113.

128 Tan, D.Q.; Atherton, A.L.; Smith, A.J.; Soldi, C.; Hurley, K.A.; Fettinger, J.C.; Shaw, J.T. “Synthesis of

a γ-Lactam Library via Formal Cycloaddition of Imines and Substituted Succinic Anhydrides,” ACS Comb.

Sci., 2012, 14, 218-223.

129 http://www.vertellus.com/products/industrial-specialties/alkenyl-succinic-anhydrides

130 https://www.indiamart.com/proddetail/polyisobutylene-succinic-anhydride-pibsa-7871743555.html

131 Rausa, R. “Synthesis of polyisobutenyl succinic anhydrides. Product distribution and proposed reaction

mechanism,” American Chemical Society, Division of Polymer Chemistry, Polymer Preprints, 2007, 48, 227-

228.

132 IHS Markit Report, Process Economics Report 34B: Phthalic Anhydride, 1998.

https://ihsmarkit.com/pdf/RP034B_toc_173528110917062932.pdf

133 IHS Markit Report. Vaswani, S. Bio-Based Succinic Acid. In Process Economics Program, 2010.

https://ihsmarkit.com/pdf/RW2010-14_220240110917062932.pdf

Page 114: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

98

134 Silvers, M.A.; Robertson, G.T.; Taylor, C.M.; Waldrop, G.L. “Design, Synthesis, and Antibacterial

Properties of Dual-Ligand Inhibitors of Acetyl-CoA Carboxylase,” J. Med. Chem., 2014, 57, 8947-8959.

135 Sustmann, R. In Comprehensive Organic Synthesis; Trost, B.M., Fleming, I., Winterfeldt, E., Eds.

Pergamon: Oxford, 1991; Vol. 6, pp. 301-321.

136 Robert, C.; de Montigny, F.; Thomas, C.M. “Facile and Efficient Synthesis of Cyclic Anhydrides from

Dicarboxylic Acids,” ACS Catal., 2014, 4, 3586-3589.

137 Rashed, M.N.; Siddiki, M.A.H.; Ali, M.A.; Moromi, S.K.; Touchy, A.S.; Kon, K.; Toyao, T.; Shimizu,

K. “Heterogeneous catalyst for the cyclization of dicarboxylic acids to cyclic anhtdrides as monomers for

bioplastic production,” Green Chem., 2017, 19, 3238-3242.

138 Sakakura, A.; Ohkubo, T.; Yamashita, R.; Akakura, M.; Ishihara, K. “Brønstead Base-Assisted Boronic

Acid Catalysis for the Dehydrative Intramolecular Condensation of Dicarboxylic Acids,” Org. Lett., 2011,

13, 892-895.

139 “Catalytic Oxidation of Benzene to Maleic Acid,” Journal of Chemical Education, 1925, 1178-1180.

140 Uraz, C.; Atalay, S. “Oxidation of Benzene to Maleic Anhydride in a Fluidized Bed Reactor,” Chem. Eng.

Technol., 2007, 30, 1708-1712.

141 Zargarian, D.; Alper, H. “Palladium Chloride Catalyzed Dicarbonylation of Terminal Alkynes,”

Organometallics, 1991, 10, 2914-2921.

142 Chiusoli, G.P.; Costa, M.; Cucchia, L.; Gabriele, B.; Salerno, G.; Veltri, L. “Carbon dioxide effect on

palladium-catalyzed sequential reactions with carbon monoxide, acetylenic compounds and water,” J. Mol.

Catal. A.: Chem., 2003, 204, 133-142.

143 Osakada, K.; Doh, M.-K.; Ozawa, F.; Yamamoto, A. “Catalytic and Stoichiometric Carbonylation of β,γ-

Unsaturated Carboxylic Acids to Give Cyclic Anhydrides through Intermediate Palladium-Containing Cyclic

Esters,” Organometallics, 1990, 9, 2197-2198.

144 Mori, Y.; Tsuji, J., “Organic Synthesis by Means of Metal Complexes. III. Cobalt Carbonyl-catalyzed

Carbonylation of β-Propiolactone,” Bull. Chem. Soc. Jpn., 1969, 42, 777-779.

145 Getzler, Y.D.Y.L.; Kundnani, V.; Lobkovsky, E.B.; Coates, G.W. “Catalytic Carbonylation of β-Lactones

to Succinic Anhydrides,” J. Am. Chem. Soc., 2004, 126, 6842-6843.

146 Rowley, J.M.; Lobkovsky, E.B.; Coates, G.W. “Catalytic Double Carbonylation of Epoxides to Succinic

Anhydrides: Catalyst Discovery, Reaction Scope, and Mechanism,” J. Am. Chem. Soc., 2007, 129, 4948-

4960.

147 Mahadevan, V.; Getzler, Y.D.Y.L.; Coates, G.W. “[Lewis Acid]+[Co(CO)4]- Complexes: A Versatile

Class of Catalysts for Carbonylative Ring Expansion of Epoxides and Aziridines,” Angew. Chem. Int. Ed.,

2002, 41, 2781-2784.

148 Mulzer, M.; Coates, G.W. “Carbonylation of cis-Disubstituted Epoxides to trans-β-Lactones: Catalysts

Displaying Steric and Contrasteric Regioselectivity,” J. Org. Chem., 2014, 79, 11851-11862.

149 Langer, J.; Görls, H.; Fischer, R.; Walther, D., “Organometallic Nickelamacrocycles of the Type

[(R2R’P)Ni(C2H4COO)]n: Synthesis and Self-Assembly to Form Different Molecular Architectures Tuned

by the Phosphine,” Organometallics, 2005, 24, 272-279.

150 Langer, J.; Walther, D.; Görls, H., “Saturated and unsaturated nickelalactones with N-heterocyclic carbene

ligands: Synthesis and structures,” J. Organomet. Chem., 2006, 691, 4874-4881.

Page 115: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

99

151 Langer, J.; Fischer, R.; Görls, H.; Walther, D., “A new set of nickelacyclic carboxylates

(“nickelalactones”) containing pyridine as a supporting ligand: synthesis, structures, and application in C-C

and C-S linkage reactions,” J. Organomet. Chem., 2004, 689, 2952-2962.

152 Langer, J.; Walther, D.; Malassa, A.; Westerhausen, M.; Görls, H. “Nickelacyclic Carboxylates with

Pyridine-Based Ligand Sets- From Mononuclear Complexes to Supramolecular Architectures by Hydrogen

Bonding,” Eur. J. Inorg. Chem., 2010, 275-281.

153 Greenburg, Z.R.; Jin, D.; Williard, P.; Bernskoetter, W.H. “Nickel promoted functionalization of CO2 to

anhydrides and ketones,” Dalton Trans., 2014, 43, 15990-15996.

154 Yamamoto, T.; Sano, K.; Yamamoto, A., “Effect of Ligand on Ring Contraction of Six-Membered Nickel-

Containing Cyclic Esters, LnNiCH2CH2CH2COO, to Their Five-Membered-Ring Isomers,

LnNiCH(CH3)CH2COO. Kinetic and Thermodynamic Control of Asymmetric Induction by Chiral

Diphosphines in the Ring Contraction,” J. Am. Chem. Soc., 1987, 109, 1092-1100.

155 Sano, K.; Yamamoto, T.; Yamamoto, A. “Preparation of Several New Ni- or Pd-Containing Cyclic Amide

and Esters, (PR3)nNiCH2CH2CH2COZ (Z = NH, O) and PCy3PdCH2CH2CH2COO, and Ring-Contraction of

the Six-Membered Ni-Containing Cyclic Ester to its Five-Membered Isomer,” Chem. Lett., 1982, 695-698.

156 Song, P.; Yu, P.; Lin, J.-S.; Li, Y.; Yang, N.-Y.; Liu, X.-Y. “Transition-Metal-Free β-C-H Bond

Carbonylation of Enamides or Amides with a Trifluoromethyl Group as CO Surrogate for the Synthesis of

1,3-Oxazin-6-ones,” Organic Lett., 2017, 19, 1330-1333.

157 Gautam, P.; Bhanage, B.M. “Recent advances in the transition metal catalyzed carbonylation of alkynes,

arenes, and aryl halides using CO surrogates,” Catal. Sci. Technol., 2015, 5, 4663-4702.

158 Wu, L.; Liu, Q.; Jackstell, R.; Beller, M. “Carbonylations of Alkenes with CO Surrogates,” Angew. Chem.

Int. Ed., 2014, 53, 6310-6320.

159 Wang, H.; Dong, B.; Wang, Y.; Liu, J.; Shi, Y. “A Palladium-Catalyzed Regioselective

Hydroesterification of Alkenylphenols to Lactones with Phenyl Formate as the CO Source,” Organic Lett.,

2014, 16, 186-189.

160 Yu, B.; Zhao, Y.; Zhang, H.; Xu, J.; Hao, L.; Gao, X.; Liu, Z. “Pd/C-catalyzed direct formylation of

aromatic iodides to aryl aldehydes using carbon dioxide as a C1 resource,” Chem. Commun., 2014, 50, 2330-

2333.

161 Konishi, H.; Nagasi, H.; Manabe, K. “Concise synthesis of cyclic carbonyl compounds from haloarenes

using phenyl formate as the CO source,” Chem. Commun., 2015, 51, 1854-1857.

162 Hoshimoto, Y.; Ohata, T.; Sasaoka, Y.; Ohashi, M.; Ogoshi, S. “Nickel(0)-Catalyzed [2+2+1]

Carbonylative Cycloaddition of Imines and Alkynes or Norbornene Leading to γ-Lactams,” J. Am. Chem.

Soc., 2014, 136, 15877-15880.

163 Ogoshi, S.; Ikeda, H.; Kurosawa, H. “Formation of an Aza-nickelacycle by Reaction of an Imine and an

Alkyne with Nickel(0): Oxidative Cyclization, Insertion, and Reductive Elimination,” Angew. Chem. Int.

Ed., 2007, 46, 4930-4932.

164 Ogoshi, S.; Ikeda, H.; Kurosawa, H. “Nickel-catalyzed [2+2+2] cycloaddition of two alkynes and an

imine,” Pure Appl. Chem., 2008, 80, 1115-1125.

165 Hoshimoto, Y.; Ohata, T.; Ohashi, M.; Ogoshi, S. “Nickel-Catalyzed Synthesis of N-Aryl-1,2-

dihydropyridines by [2+2+2] Cycloadditions of Imines with Alkynes through T-Shaped 14-Electron Aza-

Nickelacycle Key Intermediates,” Chem. Eur. J., 2014, 20, 4105-4110.

Page 116: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

100

166 Schareina, T.; Zapf, A.; Cotté, A.; Gotta, M.; Beller, M. “An Improved Protocol for Palladium-Catalyzed

Alkoxycarbonylations of Aryl Chlorides with Alkyl Formates,” Adv. Synth. Catal., 2010, 352, 1205-1209.

167 Ko, S.; Lee, C.; Choi, M.-G.; Na, Y.; Chang, S. “Chelation-Accelerated Sequential Decarbonylation of

Formate and Alkoxycarbonylation of Aryl Halides Using a Combined Ru and Pd Catalyst,” J. Org. Chem.,

2003, 68, 1607-1610.

168 Carpentier, J.-F.; Castanet, Y.; Brocard, J.; Mortreux, A.; Petit, F. “A novel and convenient method for

palladium-catalysed alkoxycarbonylation of aryl and vinyl halides using HCO2R/NaOR system,”

Tetrahedron Lett., 1991, 32, 4705-4708.

169 Ueda, T.; Konishi, H.; Manabe, K. “Palladium-Catalyzed Carbonylation of Aryl, Alkenyl, and Allyl

Halides with Phenyl Formate,” Organic Lett., 2012, 14, 3100-3103.

170 Fujihara, T.; Hosoki, T.; Katafuchi, Y.; Iwai, T.; Terao, J.; Tsuji, Y. “Palladium-catalyzed esterification

of aryl halides using aryl formates without the use of external carbon monoxide,” Chem. Commun., 2012, 48,

8012-8014.

171 Ueda, T.; Konishi, H.; Manabe, K. “Palladium-Catalyzed Reductive Carbonylation of Aryl Halides with

N-Formylsaccharin as a CO Source,” Angew. Chem. Int. Ed., 2013, 52, 8611-8615.

172 Konishi, H.; Manabe, K. “Formic Acid Derivatives as Practical Carbon Monoxide Surrogates for Metal-

Catalyzed Carbonylation Reactions,” Synlett., 2014, 25, 1971-1986.

173 Hermange, P.; Lindhardt, A.T.; Taaning, R.H.; Bjerglund, K.; Lupp, D.; Skrydstrup, T. “Ex situ

Generation of Stoichiometric and Substoichiometric 12CO and 13CO and Its Efficient Incorporation into

Palladium Catalyzed Aminocarbonylations,” J. Am. Chem. Soc., 2011, 133, 6061-6071.

174 Hoshimoto, Y.; Ashida, K.; Sasaoka, Y.; Kumar, R.; Kamikawa, K.; Verdaguer, X.; Riera, A.; Ohashi,

M.; Ogoshi, S. “Efficient Synthesis of Polycyclic γ-Lactams by Catalytic Carbonylation of Ene-Imines via

Nickelacycle Intermediates,” Angew. Chem. Int. Ed., 2017, 129, 8318-8322.

175 Müller, T.E.; Leitner, W. “CO2 Chemistry,” Beilstein J. Org. Chem., 2015, 11, 675-677.

176 Six, C.; Richter, F. Isocyanates, Organic. In Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-

VCH: Verlag, Germany, 2002; Vol. 20, pp. 63-82.

177 Shioiri, T. Heteroatom Manipulation. In Comprehensive Organic Synthesis; Trost, B.M., Fleming, I., Eds.;

Elsevier Science Ltd.: Toronto, Canada, 1991; Vol. 6, pp.795-828.

178 Sharmin, E.; Zafar, F. Polyurethane: An Introduction. InTech publication, 2012.

http://cdn.intechopen.com/pdfs/38589/InTech-Polyurethane_an_introduction.pdf

179 Richter, R.H.; Priester, R.D. Isocyanates, Organic. In Kirk-Othmer Encyclopedia of Chemical Technology

1; Wiley Blackwell: Hoboken, New Jersey, USA, 2007; Vol. 14, pp. 902-938.

180 https://www.futuremarketinsights.com/reports/acrylamide-market

181 https://ihsmarkit.com/products/acrylamide-chemical-economics-handbook.html

182 http://chemplan.biz/chemplan_demo/sample_reports/Acrylamide_Profile.pdf

183 Ayaz, E.A.; Durkan, R.; Bagis, B. “The effect of acrylamide incorporation on the thermal and physical

properties of denture resins,” J. Adv. Prosthodont., 2013, 5, 110-117.

Page 117: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

101

184 Feng, X.; Liu, J.; Rieke, P.C.; Fryxell, G.E. “Reversible Surface Properties of Glass Plate and Capillary

Tube Grafted by Photopolymerization of N-Isopropylacrylamide,” Macromolecules, 1998, 31, 7845-7850.

185 Maeda, Y.; Nakamura, T.; Ikeda, I. “Changes in the Hydration State of Poly(N-acrylamide)s During Their

Phase Transitions in Water Observed by FTIR Spectroscopy,” Macromolecules, 2001, 34, 1391-1399.

186 Sokolava, T.A.; Ovsyannikova, L.A. “Synthesis of N-Substituted Acrylamides,” Izvestiya Akademii Nauk,

1963, 9, 1658-1659.

187 Allen, C.E.; Curran, P.R.; Brearly, A.S.; Boissel, V.; Sviridenko, L.; Press, N.J.; Stonehouse, J.P.;

Armstrong, A. “Efficient and Facile Synthesis of Acrylamide Libraries for Protein-Guided Tethering,” Org.

Lett., 2015, 17, 458-460.

188 Agüero, L.; Guerrero-Ramírez, L.G.; Katime, I. “New Family of Functionalized Monomers Based on

Amines: A Novel Synthesis that Exploits the Nucleophilic Substitution Reaction,” Materials Science and

Applications, 2010, 1, 103-108.

189 Onuoha, N.I.; Wainwright, M.S. “Kinetics of the Hydrolysis of Acrylonitrile to Acrylamide over Raney

Copper,” Chem. Eng. Commun., 1984, 29, 1-12.

190 Schleicher, K.D.; Jamison, T.F. “Nickel-Catalyzed Synthesis of Acrylamides from α-Olefins and

Isocyanates,” Org. Lett., 2007, 9, 875-878.

191 Hirai, H.; Wakabayashi, H.; Komiyama, M. “Polymer-Protected Copper Colloids as Catalysts for

Selective Hydration of Acrylonitrile,” Chem. Lett., 1983, 1047-1050.

192 Hirai, H.; Wakabayashi, H.; Komiyama, M. “Catalytic Hydration of Unsaturated Nitriles to Unsaturated

Amides Using Colloidal Copper Dispersions,” Bull. Chem. Soc. Jpn., 1986, 59, 545-550.

193 Yamada, H.; Kobayashi, M. “Nitrile Hydratase and Its Application to Industrial Production of

Acrylamide,”Biosci. Biotech. Biochem., 1996, 60, 1391-400.

194 Onuoha, N.I.; Wainwright, M.S. “Catalyst Deactivation in the Hydrolysis of Acrylonitrile to Acrylamide

Over Raney Copper,” Chem. Eng. Commun., 1984, 29, 13-25.

195 Lee, J.C.; Trimm, D.L.; Wainwright, M.S.; Kohler, M.A.; Onuoha, N.I.; Cant, N.W. “Copper Catalysts

for the Hydrolysis of Acrylonitrile to Acrylamide: Deactivation and Cure,” Studies in Surface Science and

Catalysis, 1987, 34, 235-243.

196 Kohler, M.A.; Lee, J.C.; Wainwright, M.S.; Trimm, D.L.; Cant, N.W. “Hydrolysis of acrylonitrile to

acrylamide over an ion-exchanged copper on silica catalyst,” App. Catal., 1987, 35, 237-248.

197 Seo, J.H.; Ko, H.M. “Transition-metal-free synthesis of aromatic amines via the reaction of benzyne with

isocyanates,” Tetrahedron Lett., 2018, 59, 671-674.

198 Ivanovich, R.A.; Clavette, C.; Vincent-Rocan, J.-F.; Roveda, J.-G.; Gorelsky, S.I.; Beauchemin, A.M.

“Intramolecular Alkene Aminocarbonylation Using Concerted Cycloadditions of Amino-Isocyanates,”

Chem. Eur. J., 2016, 22, 7906-7916.

199 Hoberg, H. “Nickela-Heterocyclen als Intermediate der präparativen Isocyanatchemie,” J. Organomet.

Chem., 1988, 358, 507-517.

200 Hoberg, H.; Sümmermann, K. “Nickel(0)-induzierte kupplung von benzaldehyd mit isocyanaten zu

nickela-heterocyclen,” J. Organomet. Chem., 1984, 264, 379-385.

Page 118: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

102

201 Hoberg, H.; Sümmermann, K. “Nickel(0)-catalyzed synthesis of imines from isocyanates and aldehydes,”

Z Naturforsch B, 1984, 39, 1032-1036.

202 Thakur, A.; Louie, J. “Advances in Nickel-Catalyzed Cycloaddition Reactions To Construct Carbocycles

and Heterocycles,” Acc. Chem. Res., 2015, 48, 2354-2365.

203 Domínguez, G.; Pérez-Castells, J. “Recent advances in [2+2+2] cycloaddition reactions,” Chem. Soc. Rev.,

2011, 40, 3430-3444.

204 Ohshiro, Y.; Kinugasa, K.; Minami, T.; Agawa, T. “Chemistry of a Cumulated Double-Bond Compound.

X. Reactions of Isocyanates and Carbodiimide with Acetylenic Compounds,” J. Org. Chem., 1970, 35, 2136-

2140.

205 Braunstein, P.; Nobel, D. “Transition-Metal-Mediated Reactions of Organic Isocyanates,” Chem. Rev.,

1989, 89, 1927-1945.

206 Ng, S.-S.; Ho, C.-Y.; Schleicher, K.D.; Jamison, T.F. “Nickel-catalyzed coupling reactions of alkenes,”

Pure Appl. Chem., 2008, 80, 929-939.

207 Hoberg, H.; Sümmermann, K. “Diazanickelacyclopentanone aus Nickel(0), Iminen und Isocyanaten,” J.

Organomet. Chem., 1983, 253, 383-389.

208 Hoberg, H.; Sümmermann, K.; Milchereit, A. “C-C Verknüpfung von Alkenen Mit Isocyanaten am

Nickel(0),” J. Organomet. Chem., 1984, 288, 237-248.

209 Hoberg, H.; Sümmermann, K.; Milchereit, A. “CC Bond Formation of Alkenes with Isocyanates on Ni0

Complexes- a New Synthesis of Acrylamides,” Angew. Chem. Int. Ed. Eng., 1985, 24, 325-326.

210 Hoberg, H.; Hernandez, E. “Nickel(0)-catalysed Synthesis of Unsaturated Carboxylic Acid Anilides from

Ethane and Phenyl Isocyanate,” J. Chem. Soc., Chem. Commun., 1986, 544-545.

211 Hoberg, H.; Sümmermann, K.; Hernandez, E.; Ruppin, C.; Guhl, D. “Nickel(0)-katalysierte Synthese von

Crotonsäureanilid aus Propen und Phenylisocyanat,” J. Organomet. Chem., 1988, 344, C35-C38.

212 Hoberg, H.; Guhl, D. “Nickel(0) induzierte und katalysierte CC-Verknüpfungen von Phenylisocyanat mit

funktionalisierten Alkenen,” J. Organomet. Chem., 1989, 375, 245-257.

213 Hoberg, H.; Guhl, D. “Ni0-induzierte CC-Verknüpfungen von fluorsubstituieren Alkenen mit

Phenylisocyanat,” J. Organomet. Chem., 1989, 378, 279-292.

214 Hoberg, H.; Guhl, D.; Betz, P. “Nickel(0) induzierte und katalysierte CC-Verknüpfungen zwischen

Phenylisocyanat und vinylsubstituierten Heteroaromaten,” J. Organomet. Chem., 1990, 387, 233-246.

215 Tillack, A.; Selke, R.; Fischer, C.; Bilda, D.; Kortus, K. “Eine Ni0-induzierte CC-Kupplungsreaktion von

Phenylisocyanat mit Cyclopenten in Gegenwart von chiralen Phosphor-Liganden,” J. Organomet. Chem.,

1996, 518, 79-84.

216 Hoberg, H.; Bärhausen, D.; Mynott, R.; Schroth, G. “Neurtige Ni0-katalysierte Reaktionen zwischen 1,3-

Butadien und Phenylisocyanat,” J. Organomet. Chem., 1991, 410, 117-126.

217 Hoberg, H.; Bärhausen, D. “Ni0-katalysierte Herstellung von δ-Lactamen aus Isopren und

Phenylisocyanat,” J. Organomet. Chem., 1990, 397, C20-C23.

218 Hoberg, H.; Hernandez, E. “Nickel(0)-Catalyzed Synthesis of Sorbanilide from 1,3-Pentadiene and Phenyl

Isocyanate,” Angew. Chem. Int. Ed. Eng., 1985, 24, 961-962.

Page 119: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

103

219 Hoberg, H.; Nohlen, M. “Ni0-induzierte CC-Verknüpfungen von Phenylisocyanat mit cyclischen,

fünfgliedrigan Alkenen, katalytische Herstellung von β,γ-ungesättigten Carbonsäureaniliden,” J. Organomet.

Chem., 1991, 412, 225-236.

220 Hoberg, H.; Hernandez, E. “Intermoleculare CC-Verknüpfungen von Azanickelacyclopentanone, α,ω-

Disäureamide aus Alkenen und Phenylisocyanat,” J. Organomet. Chem., 1990, 387, 233-246.

221 Duong, H.A.; Cross, M.J.; Louie, J. “Nickel-Catalyzed Cycloaddition of Alkynes and Isocyanates,” J. Am.

Chem. Soc., 2004, 126, 11438-11439.

222 Ozawa, T.; Horie, H.; Kurahashi, T.; Matsubara, S. “Nickel-catalyzed [2+2+1] cycloadditions of alkynes,

acrylates, and isocyanates,” Chem. Commun., 2010, 46, 8055-8057.

223 Morimoto, M.; Nishida, Y.; Miura, T.; Murakami, M. “Nickel-catalyzed [2+2+2] Cycloaddition Reaction

of Isocyanates with 1,3,-Dienes,” Chem. Lett., 2013, 42, 550-552.

224 Miura, T.; Morimoto, M.; Murakami, M. “Enantioselective [2+2+2] Cycloaddition Reaction of

Isocyanates and Allenes Catalyzed by Nickel,” J. Am. Chem. Soc., 2010, 132, 15836-15838.

225 Duong, H.A.; Louie, J. “Regioselectivity in nickel(0)/phosphine catalyzed cycloadditions of alkynes and

isocyanates,” J. Organomet. Chem., 2005, 690, 5098-5104.

226 D’Souza, B.R.; Louie, J. “Nickel-Catalyzed Cycloadditive Couplings of Enynes and Isocyanates,”

Organic Lett., 2009, 11, 4168-4171.

227 Zhang, K.; Chopade, P.R.; Louie, J. “Coupling of vinyl aziridines and phenyl isocyanate,” Tetrahedron

Lett., 2008, 49, 4306-4309.

228 Miura, T.; Mikano, Y.; Murakami, M. “Nickel-Catalyzed Synthesis of 1,3,5-Trisubstituted Hydantoins

from Acrylates and Isocyanates,” Organic Lett., 2011, 13, 3560-3563.

229 Duong, H.A.; Louie, J. “A nickel(0) catalyzed cycloaddition of alkynes and isocyanates that affords

pyrimidine-diones,” Tetrahedron, 2006, 62, 7552-7559.

230 Shi, J.; Guo, Z.; Wei, X.; Liu, D.; Lappert, M. “Lithium Dibenzylamide, a Simple, Selective, and Highly

Efficient Catalyst for Isocyanate Cyclotrimerization to Isocyanurate,” Synlett., 2011, 13, 1937-1939.

231 Duong, H.A.; Cross, M.J.; Louie, J. “N-Heterocyclic Carbenes as Highly Efficient Catalysts for the

Cyclotrimerization of Isocyanates,” Org. Lett., 2004, 6, 4679-4681.

232 Tang, J.; Mohan, T.; Verkade, J.G. “Selective and Efficient Syntheses of Perhydro-1,3,5-triazine-2,4,6-

triones and Carbodiimides from Isocyanates Using ZP(MeNCH2CH2)3N Catalysts,” J. Org. Chem., 1994, 59,

4931-4938.

233 Dekamin, M.G.; Varmira, K.; Farahmand, M.; Sagheb-Asl, S.; Karimi, Z. “Organocatalytic, rapid, and

facile cyclotrimerization of isocyanates using tetrabutylammonium phthalimide-N-oxyl and

tetraethylammonium-2-(carbamoyl) benzoate under solvent-free conditions,” Catalysis Communications,

2010, 12, 226-230.

234 Bloodworth, A.J.; Davies, A.G. “Organometallic Reactions. Part III. Trimerisation of Isocyanates by

Organotin Alkoxides and Oxides,” J. Chem. Soc., 1965, 6858-6863.

235 Peters, S.J.; Kassabaum, M.E.; Nocella, M.K.; McDonald, R. “Spectroscopic Characterization of 1-

Naphthyl Isocyanate Anion Radical and of Tris(1-naphthyl) Isocyanurate Atropisomers,” Eur. J. Org. Chem.,

2015, 6040-6046.

Page 120: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

104

236 Wang, W.-H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. “CO2 Hydrogenation to Formate

and Methanol as an Alternative to Photo- and Electrochemical CO2 Reduction,” Chem. Rev., 2015, 115,

12936-12973.

237 Ashley, A.E.; Thompson, A.L.; O’Hare, D. “Non-Metal-Mediated Homogeneous Hydrogenation of CO2

to CH3OH,” Angew. Chem. Int. Ed., 2009, 48, 9839-9843.

238 Goeppert, A.; Czaun, M.; Jones, J.-P.; Prakash, G.K.S.; Olah, G. “Recycling of carbon dioxide to methanol

and derived products- closing the loop,” Chem. Soc. Rev., 2014, 43, 7995-8048.

239 Sordakis, K.; Tang, C.; Vogt, L.K.; Junge, H.; Dyson, P.J.; Beller, M.; Laurenczy, G. “Homogeneous

Catalysis for Suitable Hydrogen Storage in Formic Acid and Alcohols,” Chem. Rev., 2018, 118, 372-433.

240 Li, Y.-N.; Ma, R; He, L.-N.; Diao, Z.-F. “Homogeneous hydrogenation of carbon dioxide to methanol,”

Catal. Sci. Technol., 2014, 4, 1498-1512.

241 http://news.ihsmarkit.com/press-release/country-industry-forecasting-media/global-methanol-demand-

growth-driven-methanol-olefi

242 Olah, G.; Goeppert, A.; Prakash, G.K.S. Beyond Oil and Gas: The Methanol Economy; VCH, Weinheim:

2006; Chapter 10-13, pp. 168-253.

243 IHS Markit Report, 2012, http://www.formacare.org/applications/

244 Le Berre, C.; Serp, P.; Kalck, P.; Torrence, G.P. Acetic Acid. In Ullmann’s Encyclopedia of Industrial

Chemistry, Wiley-VCH: Verlag, Germany, 2013; pp. 1-34.

245 Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies,

F.; Kurr, P.; Kniep, B.-L.; Tovar, M.; Fischer, R.S.; Nørskov, J.S.; Schlögl, R. “The Active Site of Methanol

Synthesis over Cu/ZnO/Al2O3 Industrial Catalysts,” Science, 2012, 336, 893-897.

246 Cañete, B.; Gigola, C.E.; Brignole, N.B. “Synthesis Gas Processes for Methanol Production via CH4

Reforming with CO2, H2O, and O2,”Ind. Eng. Chem. Res., 2014, 53, 7103-7112.

247 Ott, J.; Gronemann, V.; Pontzen, F.; Fiedler, E.; Grossmann, G.; Kersebohm, D.B.; Weiss, G. Witte, C.

Methanol. In Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH: Verlag, Germany, 2012; pp. 1-

27.

248 Xiu, S.; Shahbazi, A. “Bio-oil production and upgrading research: A review,” Renewable and Sustainable

Energy Reviews, 2012, 16, 4406-4414.

249 Palo, D.R.; Dagle, R.A.; Holladay, J.D. “Methanol Steam Reforming for Hydrogen Production,” Chem.

Rev., 2007, 107, 3992-4021.

250 Zhang, Y.; Xioa, J.; Shen, L. “Simulation of Methanol Production from Biomass Gasification in

Interconnected Fluidized Beds,” Ind. Eng. Chem. Res., 2009, 48, 5351-5359.

251 Feng, W.; Ji, P.; Chen, B.; Zheng, D. “Analysis of Methanol Production from Biomass Gasification,”

Chem. Eng. Technol., 2011, 34, 307-317.

252 Holmgren, K.M.; Berntsson, T.; Andersson, E.; Rydberg, T. “System aspects of biomass gasification with

methanol synthesis-Process concepts and energy analysis,” Energy, 2012, 45, 817-828.

253 http://carbonrecycling.is/george-olah/

Page 121: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

105

254 Courtemanche, M.-A.; Legare, M.-A.; Maron, L.; Fontaine, F.-G. “Reducing CO2 to Methanol Using

Frustrated Lewis Pairs: On the Mechanism of Phosphine-Borane-Mediated Hydroboration of CO2,” J. Am.

Chem. Soc., 2014, 136, 10708-10717.

255 Fontaine, F.-G.; Courtemanche, M.-A.; Legare, M.-A. “Transition-Metal-Free Catalytic Reduction of

Carbon Dioxide,” Chem. Eur. J., 2014, 20, 2990-2996.

256 Huang, F.; Lu, G.; Zhao, L.; Li, H.; Wang, Z.-X. “The Catalytic Role of N-Heterocyclic Carbene in a

Metal-Free Conversion of Carbon Dioxide into Methanol: A Computational Mechanism Study,” J. Am.

Chem. Soc., 2010, 132, 12388-12396.

257 Riduan, S.N.; Zhang, Y.; Wing, J.Y. “Conversion of Carbon Dioxide into Methanol with Silanes over N-

Heterocyclic Carbene Catalysts,” Angew. Chem. Int. Ed., 2009, 48, 3322-3325.

258 Chakraborty, S.; Zhang, J.; Krause, J.A.; Guan, H. “An Efficient Nickel Catalyst for the Reduction of

Carbon Dioxide with Borane,” J. Am. Chem. Soc., 2010, 132, 8872-8873.

259 Das Neves Gomez, C.; Blondiaux, E.; Thuery, P.; Cantat, T. “Metal-Free Reduction of CO2 with

Hydroboranes: Two Efficient Pathways at Play for the Reduction of CO2 to Methanol,” Chem. Eur. J., 2014,

20, 7098-7106.

260 Riduan, S.N.; Ying, J.Y.; Zhang, Y. “Mechanistic Insights into the Reduction of Carbon Dioxide with

Silanes over N-Heterocyclic Carbene Catalysts,” ChemCatChem, 2013, 5, 1490-1496.

261 Lim, C.-H.; Holder, A.M.; Hynes, J.T.; Musgrave, C.B. “Reduction of CO2 to Methanol Catalyzed by

Biomimetic Organo-Hydride Produced from Pyridine,” J. Am. Chem. Soc., 2014, 136, 16081-16095.

262 Huff, C.A.; Sanford, M.S. “Cascade Catalysis for the Homogeneous Hydrogenation of CO2 to Methanol,”

J. Am. Chem. Soc., 2011, 133, 18122-18125.

263 Rodríguez-Lugo, R.E.; Trincado, M.; Vogt, M.; Tewes, F.; Santiso-Quinones, G.; Grützmacher, H. “A

homogeneous transition metal complex for clean hydrogen production from methanol-water mixtures,” Nat.

Chem., 2013, 5, 342-347.

264 Tominaga, K.-I.; Sasaki, Y.; Watanabe, T.; Saito, M. “Ruthenium complex catalysed hydrogenation of

carbon dioxide to carbon monoxide, methanol and methane,” J. Chem. Soc., Chem. Commun., 1993, 629-

631.

265 Tominaga, K.-I.; Sasaki, Y.; Watanabe, T.; Saito, M. “Homogeneous Hydrogenation of Carbon Dioxide

to Methanol Catalyzed by Ruthenium Cluster Anions in the Presence of Halide Anions,” Bull. Chem. Soc.

Jpn., 1995, 68, 2837-2842.

266 Grasemann, M.; Laurenczy, G. “Formic acid as a hydrogen source- recent developments and future

trends,” Energy & Environmental Science, 2012, 5, 8171-8181.

267 Miller, A.J.M.; Heinekey, D.M.; Mayer, J.M.; Goldberg, K.I. “Catalytic Disproportionation of Formic

Acid to Generate Methanol,” Angew. Chem. Int. Ed., 2013, 52, 3981-3984.

268 Savourey, S.; Lefèvre, G.; Berthet, J.-T.; Thuéry, P.; Genre, C.; Cantat, T. “Efficient Disproportionation

of Formic Acid to Methanol Using Molecular Ruthenium Catalysts,” Angew. Chem. Int. Ed., 2014, 53,

10466-10470.

269 Sordakis, K.; Tsurusaki, A.; Iguchi, M.; Kawanami, H.; Himeda, Y.; Laurenczy, G. “Carbon Dioxide to

Methanol: The Aqueous Catalytic Way at Room Temperature,” Chem. Eur. J., 2016, 22, 15604-15608.

Page 122: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

106

270 Balaraman, E.; Gunanathan, C.; Zhang, J.; Shimon, L.J.W.; Milstein, D. “Efficient hydrogenation of

organic carbonates, carbamates, and formates indicates alternative routes to methanol based on CO2 and CO,”

Nat. Chem., 2011, 3, 609-614.

271 Balaraman, E.; Ben-David, Y.; Milstein, D. “Unprecedented Catalytic Hydrogenation of Urea Derivatives

to Amines and Methanol,” Angew. Chem. Int. Ed., 2011, 50, 11702-11705.

272 Han, Z.; Rong, L.; Wu, J.; Zhang, L.; Wang, Z.; Ding, K. “Catalytic Hydrogenation of Cyclic Carbonates:

A Practical Approach from CO2 and Epoxides to Methanol and Diols,” Angew. Chem. Int. Ed., 2012, 51,

13041-13045.

273 Wesselbaum, S.; vom Stein, T.; Klankermayer, J.; Leitner, W. “Hydrogenation of Carbon Dioxide to

Methanol by Using a Homogeneous Ruthenium-Phosphine Catalyst,” Angew. Chem. Int. Ed., 2012, 51, 7499-

7502.

274 Wesselbaum, S.; Moha, V.; Meuresch, M.; Brosinski, S.; Thenert, K.M.; Kothe, J.; vom Stein, T.; Englert,

U.; Hoelscher, M.; Klankermayer, J.; Leitner, W. “Hydrogenation of carbon dioxide to methanol using a

homogeneous ruthenium-Triphos catalyst: from mechanistic investigations to multiphase catalysis,” Chem.

Sci., 2015, 6, 693-704.

275 Rezayee, N.M.; Huff, C.A.; Sanford, M.S. “Tandem Amine and Ruthenium-Catalyzed Hydrogenation of

CO2 to Methanol,” J. Am. Chem. Soc., 2015, 137, 1028-1031.

276 Kothandaraman, J.; Goeppert, A.; Czaun, M.; Olah, G.A.; Prakash, G.K.S. “Conversion of CO2 from Air

into Methanol Using a Polyamine and a Homogeneous Ruthenium Catalyst,” J. Am. Chem. Soc., 2016, 138,

778-781.

277 Kar, S.; Sen, R.; Goeppert, A.; Prakash, G.K.S. “Integrative CO2 Capture and Hydrogenation to Methanol

with Reusable Catalyst and Amine: Toward a Carbon-Neutral Methanol Economy,” J. Am. Chem. Soc., 2018,

140, 1580-1583.

278 Khusnutdinova, J.R.; Garg, J.A.; Milstein, D. “Combining Low-Pressure CO2 Capture and Hydrogenation

To Form Methanol,” ACS Catal., 2015, 5, 2416-2422.

279 Kar, S.; Goeppert, A.; Kothandaraman, J.; Prakash, G.K.S. “Manganese-Catalyzed Sequential

Hydrogenation of CO2 to Methanol via Formamide,” ACS Catal., 2017, 7, 6347-6351.

280 Schneidewind, J.; Adam, R.; Baumann, W.; Jackstell, R.; Beller, M. “Low-Temperature Hydrogenation

of Carbon Dioxide to Methanol with a Homogeneous Cobalt Catalyst,” Angew. Chem. Int. Ed., 2017, 56,

1890-1893.

281 Tanaka, R.; Yamashita, M.; Nozaki, K. “Catalytic Hydrogenation of Carbon Dioxide Using Ir(III)-Pincer

Complexes,” J. Am. Chem. Soc., 2009, 131, 14168-14169.

282 Zhang, Y.; MacIntosh, A.D.; Wong, J.L.; Bielinski, E.A.; Williard, P.G.; Mercado, B.Q.; Hazari, N.;

Bernskoetter, W.H. “Iron catalyzed CO2 hydrogenation to formate enhanced by Lewis acid co-catalysts,”

Chem. Sci., 2015, 6, 4291-4299.

283 Bielinski, E.A.; Lagaditis, P.O.; Zhang, Y.; Mercado, B.Q.; Würtele, C.; Bernskoetter, W.H.; Hazari, N.;

Schneider, S. “Lewis Acid-Assisted Formic Acid Dehydrogenation Using a Pincer-Supported Iron Catalyst,”

J. Am. Chem. Soc., 2014, 136, 10234-10237.

284 Bielinski, E.A.; Förster, M.; Zhang, Y.; Bernskoetter, W.H.; Hazari, N.; Holthausen, M.C. “Base-Free

Methanol Dehydrogenation Using a Pincer-Supported Iron Compound and Lewis Acid Co-catalyst,” ACS

Catal., 2015, 2, 2404-2415.

Page 123: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

107

285 Jayarathne, U.; Hazari, N.; Bernskoetter, W.H. “Selective Iron-Catalyzed N-Formylation of Amines using

Dihydrogen and Carbon Dioxide,” ACS Catal., 2018, 8, 1338-1345.

286 Jayarathne, U.; Zhang, Y.; Hazari, N.; Bernskoetter, W.H. “Selective Iron-Catalyzed Deaminative

Hydrogenation of Amides,” Organometallics, 2017, 36, 409-416.

287 Unpublished work by Dr. Yuanyuan Zhang

Page 124: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

108

Chapter 2:

The Application of an Iron Catalyst in the Transformation of Methanol into Formamides

and Ureas via Acceportless Dehydrogenation Reactions

Reprinted (adapted) with permission from Lane, E.M.; Uttley, K.B.; Hazari, N.;

Bernskoetter, W.H. “Iron-Catalyzed Amide Formation from the Dehydrogenative

Coupling of Alcohols and Secondary Amines,” Organometallics, 2017, 36, 2020-2025.

Copyright 2017 American Chemical Society.

Reprinted (adapted) with permission from Lane, E.M.; Hazari, N.; Bernskoetter, W.H.

“Iron-Catalyzed Urea Synthesis: Dehydrogenative Coupling of Methanol and Amines,”

Chem. Sci., 2018, 9, 4003-4008. Published by the Royal Society of Chemistry.

Page 125: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

109

2.1: Introduction: Methanol as a Carbon Feedstock

As mentioned in Chapter 1, methanol is currently produced on an industrial scale

from a synthesis gas mixture (syngas; H2/CO2/CO) using alumina-supported mixed copper

and zinc oxide heterogeneous catalysts at elevated temperature and pressure (Scheme 2.1.1,

reaction (1)).1,2 While most syngas is derived from coal and natural gas,3 it can be

renewably sourced from biomass (via gasification),4-6 albeit at lower yields than its

nonrenewable counterparts.7,8 As of 2016, global methanol demand had reached 80 million

metric tons and since then has continued to rise.9,10 Although methanol is often used as a

solvent, >95% of methanol produced is subsequently converted into other valuable

chemicals,9 making it an abundant and potentially renewable carbon feedstock (Scheme

2.1.2).3 As raw material feedstocks typically constitute 60-70% of manufacturing

expenses,3 the low cost of methanol provides a significant economic advantage in this

respect. Formaldehyde accounts for the most substantial percentage of methanol demand

(27%),9 however, its value is largely based on its role as a highly active intermediate on

the pathway from methanol to more functionalized species. For example, further

transformation of formaldehyde yields high-demand products like resins for building

Scheme 2.1.1: (1) Industrial methanol production from syngas and (2) industrial

formaldehyde production from methanol.

Page 126: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

110

materials, industrial and automotive plastics, insulation, and pharmaceuticals.3,11 Industrial

production of formaldehyde is achieved through methanol oxidation/dehydrogenation

using either a silver or iron-molybdenum oxide heterogeneous catalyst in a fixed-bed

reactor at high temperatures and pressures (Scheme 2.1.1, reaction (2)).12-14 These

conditions are appropriate for formaldehyde, but they are incompatible with the additional

functionalization necessary to give formaldehyde-derived desirable products. This is a

common issue amongst many of the first generation methanol-sourced compounds in

Scheme 2.1.2, and as a result, commodity chemical production from methanol is relegated

to a multi-step process. An alternative method to achieve the goal of methanol

functionalization is to instead use metal-catalyzed dehydrogenative coupling under milder

conditions to convert methanol to the end product (via a formaldehyde intermediate) in one

pot, turning it into a more effective C1 carbon source.

Scheme 2.1.2: Some of the major chemical species derived from methanol and their

uses.

Page 127: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

111

Metal-catalyzed alcohol dehydrogenation effectively transforms an alcohol into a

more reactive aldehyde or ketone species without stoichiometric amounts of harsh, toxic,

or waste-producing oxidizing agents.15,16 The absence of such reagents opens the door to a

host of previously unavailable manipulations, as homogeneous organometallic catalysts

often suffer ligand degradation under oxidative conditions and therefore their applicability

to such processes was previously limited.17 The first known example of metal-promoted

alcohol dehydrogenation was introduced in the late 1800’s by Guerbet and was enacted

under harshly basic and high temperature conditions (130-180 °C) with heterogeneous

catalysts.18 Significant advancement in the field has occurred since then and is extensively

covered in several reviews.16-21 The concept of dehydrogenation as a form of alcohol

activation is a much more recent development, however, and while a host of examples now

exist, it was the early work by the groups of Fujita and Yamaguchi,22 Ramón and Yus,23

and Williams24,25 that generalized the method for a variety of transformations and formally

established the idea. By this process, a nucleophilic alcohol is transformed into an

electrophilic aldehyde or ketone, priming it for subsequent reaction with a different

nucleophile and further functionalization.18,23 In this manner a variety of bond-forming

reactions (such as C-C, C-N, and C-O)16,18,24,24,82 can be enacted to yield valuable

functionalized products from mild, cheap, and easy to handle starting materials. Alcohol

dehydrogenation followed by tandem coupling therefore accesses a diverse assortment of

useful products in a more environmental and economical manner.

In traditional alcohol dehydrogenation, the product of the reaction of the generated

aldehyde/ketone with other substrates also serves as an internal hydrogen acceptor to re-

activate the hydrogenated catalyst.23 However, this interferes with the generation of

Page 128: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

112

unsaturated products. One solution to this problem is to use a sacrificial hydrogen

acceptor26 like an alkene (e.g. tert-butyl ethylene)27 or a ketone (e.g. acetone),28 but these

additives result in additional reaction cost and waste generation.17 It is therefore

advantageous to incorporate a catalyst that can release the hydrogen it removes from the

alcohol as dihydrogen, referred to as acceptorless alcohol dehydrogenation. This is

typically accomplished using higher reaction temperatures, which enhance the kinetics and

thermodynamics of the reaction by speeding it up and by providing an entropic driving

force via the production of two species rather than one (respectively). In addition,

performing the dehydrogenation reaction at reflux temperatures effectively removes the

dihydrogen from solution, further favoring the process.17,29 However, even this byproduct

has value, as reversible dihydrogen generation from methanol has potential alternative

energy applications and there have been extensive investigations into the use of methanol

as a hydrogen storage material due to its high energy density (see chapter 1 for more

information).

Methanol dehydrogenation presents an additional challenge due to a greater

energetic demand for this process (ΔH = +84 kJ/mol) compared to that for higher alcohols

(e.g. ethanol ΔH = +68 kJ/mol)14,30,31 The formaldehyde intermediate initially produced is

also very unstable and will readily decompose to CO and H2 if it is not trapped effectively.65

As a result, the number of catalysts capable of methanol activation and functionalization is

limited, and those incorporating cheap and abundant base metals are even rarer.62,65,82 In

this chapter, a five-coordinate iron (II) PNP pincer complex for the catalytic acceptorless

dehydrogenative coupling of methanol with amines to give commercially valuable

formamides and ureas is introduced. Its impressive activity in this respect matches and

Page 129: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

113

in the case of amidation even exceeds the performance of precious-metal catalysts,

representing significant advancement in this approach to methanol transformation.

Page 130: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

114

2.2: Iron-Catalyzed Amide Formation from the Dehydrogenative Coupling of

Alcohols and Secondary Amines

2.2.1: Introduction

Amides are key functional groups in organic synthesis, biochemistry, and

pharmaceuticals.32 For example, it is estimated that approximately 25% of drug

molecules contain amides.33 Current methods for large-scale amide synthesis

typically involve the reaction of carboxylic acids33,34 or more commonly their

activated derivatives (such as acid chlorides, acid anhydrides, and acyl azides)35-37

with stoichiometric amounts of amines (Scheme 2.2.1), and additional inorganic or

organic promoters (e.g. bases, boronic acids, or borate esters) are often used. As a

result, these procedures are not atom efficient, they generate considerable chemical

waste, and they use relatively expensive starting materials.32 The dehydrogenative

coupling of alcohols and amines to form amides (with dihydrogen as the only

byproduct, Scheme 2.2.2) offers an atom-economical and more environmentally

Scheme 2.2.1: Metal-free methods for amide production.

Page 131: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

115

benign alternative to obtaining these materials. Although coinage metals such as

silver38 and gold39-42 have demonstrated some effectiveness as heterogeneous

catalysts for dehydrogenative amide synthesis, these systems normally require high

catalyst loadings, harsh reaction conditions, long reaction times, and have limited

substrate scopes. The sole heterogenous copper example exhibits a slightly wider

scope but otherwise suffers from similar issues, as well as the necessity of an

oxidatively harsh peroxide additive.43 Homogenous transition metal catalysts

incorporating rhenium,44 rhodium,45 and (most extensively) ruthenium have shown

greater activity.46-63 The most prominent example, a ruthenium-pincer catalyst

published in Science in 2007 by Milstein and co-workers, achieved TONs of nearly

1000 for the dehydrogenative coupling of primary amines with sterically accessible

alcohols (Figure 2.2.1).47 This spurred the development of a host of other

predominantly precious-metal catalysts for dehydrogenative amide synthesis, but

nearly all suffered from the limitations of requiring an added base, needing a

stoichiometric amount of a hydrogen acceptor, giving low TONs (<100), or a

combination of these factors. Notably, most existing catalysts perform best with

Scheme 2.2.2: Metal-promoted pathways for dehydrogenative amidation.

Page 132: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

116

primary amine substrates, and there are only a limited number of catalysts (all but

two are ruthenium-based) that are compatible with a range of secondary amine

substrates.51,52,54,56,58-60,62,64,65 To date, only two of these catalysts have achieved

TONs greater than 50 for the formation of tertiary amides under base-free and/or

hydrogen acceptor-free conditions (Figure 2.2.2).56,59 Furthermore, the development

of promotors incorporating cheaper, less toxic metals is lacking, as only three base-

metal catalysts (one copper and two manganese) have been described for

dehydrogenative amidation (Figure 2.2.1).64-66 While these recent studies

demonstrate some promise of base-metal-mediated coupling of amines and alcohols,

Figure 2.2.1: A top-performing ruthenium catalyst (a) and all previous base-metal

catalysts (b) for dehydrogenative amide formation from amines and alcohols.

Page 133: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

117

they exhibit limited tertiary amide production (TONs <50) and leave open the

possibility for substantial improvement.

As previously mentioned, smaller alcohols like methanol have to overcome a

larger activation energy for dehydrogenation, which makes them incompatible with

many catalysts.31 However, formamide production from the dehydrogenative

coupling of methanol and amines is a crucial target. Formamides not only appear in

agrochemicals, dyes, and fragrances,67,68 but they serve as reagents69 (e.g. in the

Vilsmeier-Haack reaction)70 and are key organic intermediates, such as in the

production of formamidines71 and isocyanides.72,73 They are also important amine

protecting groups in protein biosynthesis.74-76 Perhaps their most valuable trait,

however, is their high biological activity, which makes them integral to various

pharmaceuticals.62,77 Formamides are typically synthesized from amines and excess

formylating agents (chloral, phenyl formate, formic acid), which are

environmentally hazardous, generate copious waste, and display a small range of

utility.78,79 Most of the few existing methods for using methanol as a “green” C1

building block in place of these hazardous formylating reagents in formamide

syntheses require harsh conditions, long reaction times, and show limited scope.41-

43,62 Of the many previously reported homogeneous catalysts for amide formation

from alcohols and amines, only four (three ruthenium-based and one manganese-

based) were capable of formamide production (Scheme 2.2.2, R=H). All four

exhibited low activities, three58,62,65 with TONs ≤ 50 (Figure 2.2.2) and one

ruthenium catalyst80 with TONs < 100 but which required a base additive, displayed

Page 134: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

118

poor selectivity, and was only investigated with one amine substrate. Therefore, a

catalyst capable of the dehydrogenative coupling of methanol and amines in the

Figure 2.2.2: Efficient ruthenium, manganese, and iron catalysts for the dehydrogenative

coupling of alcohols and 2° amines.

Page 135: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

119

absence of other reagents is both desirable and rare, especially one incorporating a

base metal.

Several independent research endeavours (including within our laboratories) have

reported the successful application of iron-pincer catalysts of the type

(RPNP)Fe(H)(CO) (RPNP = N[CH2CH2(PR2)]2, R = iPr, Cy) in the dehydrogenation

of alcohols to esters,81 ketones,81 or carbon dioxide.82,83 Recently, we also described

the use of the five-coordinate iron (II) complex (iPrPNP)Fe(H)(CO) (1) in the

catalytic base-free hydrogenation of amides,84 which suggests it may have the

potential to catalyze the microscopic reverse reaction (the dehydrogenative

formation of amides). The following discussion summarizes the impressive ability

of 1 to catalyze the dehydrogenation of alcohols in the presence of secondary amines

to selectively produce amides in the absence of secondary hydrogen acceptors or

base promotors (Figure 2.2.2). Complex 1 is the most productive base-metal catalyst

yet developed for intermolecular dehydrogenative amide synthesis and gives TONs

that exceed those previously described for base or precious metals with respect to

tertiary amide formation.85

2.2.2: Results and Discussion

Catalytic Studies

Due to the success of 4-formylmorpholine as a substrate in prior amide

hydrogenation experiments,84 investigations began with attempts to couple

morpholine and methanol in the presence of 1. An initial experiment was performed

using a 4:1 amine:alcohol ratio and 0.1 mol% of 1 in 10 mL of dioxane at 80 °C for

8 hours. Despite the lack of optimization, this reaction yielded a promising TON of

Page 136: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

120

297 for the formation of 4-formylmorpholine. This TON is already several times

that exhibited by any other base-metal catalyst for this reaction. Using this reaction

as a benchmark, the solvent, reaction volume, time, temperature, and substrate ratio

were all varied to generate an optimized set of reaction conditions (Tables A2.1-

A2.5). Using these optimized conditions (illustrated in Table 2.2.1), a TON of 503

was achieved for the formation of 4-formylmorpholine (entry 3), which rivals the

highest TONs reported for ruthenium catalysts in dehydrogenative amidation using

secondary amines and far exceeds those of base-metal catalysts. The TON was

confirmed by NMR analysis of the 4-formylmorpholine product and by collection

of the dihydrogen produced (86% of expected H2 was collected). Encouraged by the

activity of 1 with traditionally more difficult secondary amines, the substrate scope

of dehydrogenative amidation was further explored. Overall, 1 is extremely effective

at amide production using cyclic secondary amines (entries 1-3), with a maximum

TON of 600 observed using 1,2,3,4-tetrahydroisoquinoline. However, the catalytic

activity appears strongly correlated to the size of the substituents on the amine. For

example, in comparing sterically restricted cyclic amines to substrates with even

small linear substituents, such as ethyl, there exists a significant drop in TON down

to approximately 200 (entry 4). This effect is further demonstrated as the amine

substituents increase in size (entries 5-7), with only trace conversion observed for

the phenyl and isopropyl derivatives. Attempts to generate diamides starting from

cyclic amides (entries 8 and 9) were also unsuccessful, although previous studies

from our laboratories suggest that this may be due to a competing 1,2-addition

reaction between secondary amides and 1.84

Page 137: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

121

Entry Amine TON

1

600b

2

564b

3

503b

4 (CH3CH2)2NH 213

5 (C6H5CH2)2NH 126

6 (C6H5)2NH <10

7 [(CH3)2CH]2NH <10

8

0b

9

0b,c

Table 2.2.1: Dehydrogenative Amidation of

Methanol Catalyzed by 1.a

a Reaction conditions: 3 µmol catalyst (0.1 mol%),

3 mmol alcohol, and 12 mmol amine in 5 mL THF

at 80 °C for 8 hrs. Each entry is an average of two

trials and the TON was determined by GC analysis

of amide production unless otherwise indicated. bDetermined by 1H NMR spectroscopy.

cOnly one

trial.

Page 138: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

122

Initial studies into extending the scope of amines to include primary amine

substrates have indicated a propensity for 1 to further substitute the initially formed

formamides to yield ureas, which is the subject of the next section in this chapter.

Despite the aforementioned steric limitations on the amine substrates, the ability of

1 to dehydrogenatively activate methanol and thereby catalyze the production of a

range of formamides with TONs of up to 600 in the absence of any additives is

unrivalled by base- and precious-metal catalysts alike. These results therefore

represent a significant step forward with respect to the use of methanol as a carbon

feedstock for more complex and commercially valuable organic species via

homogeneously catalyzed dehydrogenative coupling reactions.

The success of 1 in dehydrogenating a challenging substrate like methanol

motivated an investigation into the alcohol substrate scope for amidation (Table

2.2.2). Despite the fact that the primary focus of this discussion has been on the use

of methanol as a carbon source, access to a suite of alcohols capable of this

transformation would maintain the advantages of this synthetic approach while

further diversifying the product pool. Similar to the observations with amines, the

performance of 1 is highly sensitive to the steric profile of the alcohol (entries 1-3).

This effect on catalytic activity remains even when the larger substituent is slightly

removed from the dehydrogenation site (entries 4 and 5). The absence of catalysis

in the presence of 2,2,2-trifluoroethanol (entry 6) is likely due to its relative acidity

and its electron-withdrawing substituent. Interestingly, methanol far outperforms the

other alcohols explored. When this observation is combined with the results on the

effect of sterics on the amine substrate, it indicates that there is a strong steric effect

Page 139: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

123

at both sides of the product carbonyl moiety. To some extent these results are

unsurprising, as large groups would potentially interfere with metal-catalyzed

dehydrogenation steps in which the substrates or intermediates must pass the

isopropyl substituents of the pincer ligand. Although the overall substrate scope

exhibited by 1 remains quite sterically limited, the high productivity observed for

small substrates suggests that the inherent activity of the iron center is significant

and could perhaps be leveraged by further attenuation of the ancillary ligand

environment.

Mechanistic Investigations

Entry Alcohol TON

1 CH3OH 503

2 CH3CH2OH 50

3 CH3(CH2)5OH 13

4 C6H5(CH2)2OH 10

5 C6H5CH2OH 10

6 CF3CH2OH 0b

Table 2.2.2: Alcohol Dehydrogenation in the

Presence of Morpholine Catalyzed by 1.a

aReaction conditions: 3 µmol catalyst (0.1 mol%),

3 mmol alcohol, and 12 mmol amine in 5 mL THF

at 80 °C for 8 hrs. Each entry is an average of two

trials and TON was determined by GC analysis of

amide production unless otherwise indicated. bDetermined by 1H NMR spectroscopy.

Page 140: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

124

Recent investigations of transition-metal-catalyzed dehydrogenative amidation

have nearly all proposed the same general pathway for amide formation (Scheme

2.2.1).47,49,61 Each begins with the dehydrogenation of the alcohol to the

corresponding aldehyde followed by trapping either with amine or an additional

equivalent of alcohol to generate a hemiaminal or hemiacetal intermediate,

respectively. Dehydrogenation of an intermediate hemiaminal would directly afford

amide, while oxidation of the hemiacetal would yield an ester which could then

produce amide via a transamidation process. However, the exact mechanisms for

dehydrogenative amidation are likely catalyst and system dependant. For example,

many ruthenium catalysts employ base additives to achieve optimum performance,

which can potentially serve many different roles. These include enhancing

coordination of the alcohol to the metal,50 facilitating amine-to-aldehyde proton

transfer, obviating hemiaminal formation,61 or promoting ester transamidation.54

Prior studies within our laboratories demonstrate that 1 is quite capable of

promoting ester formation, achieving a TON of 107 over just 12 minutes for the

methanol to methylformate conversion under conditions similar to those used in this

work for amidation.82 Additional amidation experiments using catalyst 1 and

morpholine with methyl formate in place of alcohol yielded substantial quantities 4-

formylmorpholine (Scheme 2.2.3, Table A2.7).86 These observation suggest that the

lower pathway illustrated in Scheme 2.2.1 is at least viable for 1-catalyzed

amidation. However, further mechanistic experiments substituting benzaldehyde

for alcohol during the amidation process similarly afforded the corresponding amide

product (Scheme 2.2.3, Table A2.8), which also validates the hemiaminal-dependent

Page 141: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

125

pathway for amide production.86 Since these preliminary investigations indicate that

both of the Scheme 2.2.1 pathways are competent for 1, it is clear that a more

exacting and extensive mechanistic undertaking will be require to elucidate the

precise path of 1-catalyzed amidation.

During the course of the mechanistic and catalytic experiments described above,

it was noted that when the catalytic experiments were assembled, the solution of 1

and amine immediately changed color from deep red to yellow-orange upon addition

of the alcohol substrate at room temperature. When the solution was heated during

catalysis, the color would quickly revert back to a shade of deep red resembling its

original hue. In order to ascertain any mechanistic implications that this color change

might entail, in situ NMR spectroscopic studies using morpholine and methanol

were performed. Multiple experiments showed that the color change and

corresponding changes to the 31P and 1H NMR spectra were exclusively dependent

on the addition of alcohol (Figure A2.1), since treatment of 1 with only morpholine

produced no observable reaction. The reaction of 1 with methanol was examined

using variable-temperature NMR spectroscopy, which indicated an equilibrium

between 1 and the iron-methoxy species (iPrPN(H)P)Fe(H)(OCH3)(CO) (2) formed

Scheme 2.2.3: Iron-catalyzed coupling of benzaldehyde or methyl formate with

morpholine.

Page 142: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

126

by the addition of methanol across the Fe-N bond. At all observed temperatures

between 22 and 70 °C, only one 31P NMR resonance was seen, although its chemical

shift changes with temperature (Figure 2.2.3). This is indicative of a rapid

equilibration between 1 and 2 on the timescale of the NMR experiment, and the

position of the resonance represents an average value for the chemical shifts of the

two species present in solution. The change in the 31P NMR chemical shift suggests

that 1 is favored at high temperatures and 2 is preferred at lower temperatures, as

expected for a bimolecular to unimolecular reaction. The identity of 2 was confirmed

using 13CH3OH labeling and low-temperature NMR studies, which slowed the

Figure 2.2.3: Variable temperature 31P NMR

spectra indicate an equilibrium between 1 and 2

in the presence of 6 equiv. of methanol. Sample

dissolved in benzene-d6.

1

2 22 °C

50 °C

60 °C

70 °C

Page 143: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

127

exchange rate between free and iron-activated methanol. At -80 °C, two distinct

peaks for the bound and free 13CH3 fragment were observed in the 13C {1H} spectrum

and a sharp resonance for 2 was seen at 93.2 ppm in the 31P NMR spectrum (Figure

A2.2). A proton-coupled 13C spectrum showed that the labeled carbon is still

attached to three protons, supporting the presence of an iron-methoxy species

(Figure A2.2). The exact position of the 31P{1H} NMR peak for 2 depends both on

temperature and the number of equivalents of methanol due to the equilibrium

described above.

The formation 2 via deprotonation of methanol is likely an off-cycle reaction that

competes with catalytic amidation. As mentioned above and illustrated in Scheme

2.2.1, the dehydrogenation of an alcohol to an intermediate aldehyde has been

implicated in most related amidation processes.47,49,61 Our prior experimental and

computational studies of 1-catalyzed methanol dehydrogenation have found that this

reaction proceeds via a concerted transfer of the hydrogenation atoms associated

with the methanol O-H and C-H bonds. Dehydrogenation pathways involving

intermediates analogous to complex 2 were found to be significantly higher in

energy.82 Thus, while formation of 2 may have some kinetic relevance to catalytic

amidation, it is likely not directly on its pathway and the high-temperature catalytic

conditions presumably minimize the significance of 2 in amide formation.

2.2.3: Summary

The five-coordinate iron complex 1 exhibits the highest productivity yet

reported for base-metal-catalyzed intermolecular dehydrogenative coupling of

alcohols and amines to amides. This iron-mediated amidation displays an intriguing

Page 144: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

128

preference for secondary amines, serving as a complement to the leading precious-

metal catalysts, which work best with primary amines. In fact, the TONs of up to

600 for tertiary amide production far exceed the performance of any catalyst to date.

Perhaps most importantly, it displays notable success in the notoriously difficult

production of formamides from methanol and amines. In this manner, these results

represent a significant advancement toward a more widespread use of methanol as a

versatile and potentially renewable carbon building block in organic syntheses.

Page 145: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

129

2.3: Iron-Catalyzed Urea Synthesis: Dehydrogenative Coupling of Methanol and

Amines

2.3.1: Introduction

Ureas are frequent intermediates in organic synthesis and appear in resin

precursors, dyes, and agrochemicals.87,88 However, their most prominent role is in

pharmaceuticals, where they are key functional groups in a host of medicinal

compounds89,90 such as antibacterial,91 antiatherosclerotic,92 antidepressant,93 and

anticancer94,95 drugs. The most prevalent metal-free methods for synthesizing ureas use

highly toxic or dangerous starting materials, such as phosgene and its derivatives,

isocyanates, azides,96-98 or carbon monoxide (CO).97,99,100 Such CO reactions also

necessitate elevated temperatures and pressures or the concurrent use of strong

oxidants.97,101-106 Carbon dioxide (CO2) can be used as a non-toxic urea precursor, but these

reactions typically require high temperatures and pressures, expensive dehydrating agents,

Scheme 2.3.1: Metal-free pathways for urea formation.

Page 146: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

130

or strong bases.97,107-109 Additionally, all current metal-free methods produce

stoichiometric amounts of inorganic salts as waste (Scheme 2.3.1).

Transition metal catalysis is a potential strategy for improving the preparation of

ureas (Scheme 2.3.2). Of the most highly explored methods, the synthesis of symmetric

ureas via the metal-catalyzed oxidative carbonylation of amines requires high catalyst

loadings, generally gives low yields, and often forms significant amounts of side products

(such as oxamides, carbamate esters, or CO2). Furthermore, elevated pressures of CO and

harsh oxidizing conditions are needed.100,107,110-119 Urea generation through the metal-

catalyzed reaction of CO2 and substituted amines suffers from high temperatures or

pressures, base additives, poor reaction yields and scope, or some combination of these

factors.107,120-130 Metal-catalyzed dehydrogenative coupling approaches for urea synthesis

avoid pressurization concerns and have the advantage of producing dihydrogen (H2) as the

only byproduct. Unfortunately, the few examples of dehydrogenative coupling of

formamides and amines131-134 and amide cross-coupling135,136 require high catalyst loadings

Scheme 2.3.2: Metal-catalyzed urea synthesis from non-methanol sources.

Page 147: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

131

(2-5 mol%) and most also need a stoichiometric peroxide additive. In addition, the

preparation of the formamide starting materials is often tedious and typically employs

harsh formylating reagents.139 A desirable alternative is to instead dehydrogenatively

couple methanol with an amine to form a formamide intermediate. Subsequent

dehydrogenative coupling of the formamide with another equivalent of amine generates a

urea (and H2; Scheme 2.3.3). The paucity of catalysts suitable for even the first step of this

transformation was discussed in the previous section, and is demonstrated by the fact that

while numerous metal catalysts are capable of dehydrogenative amidation,137 only four

exhibit any activity with respect to coupling methanol and amines to formamides in the

absence of harsh oxidative conditions.138-141 Only the five-coordinate iron catalyst 1 does

so with turnover numbers (TONs) greater than 50.141 The incorporation of primary amines

into dehydrogenative amidation and ureation reactions faces the additional challenge of

potential imine formation via dehydration from a plausible hemiaminal intermediate.142

While imines can be valuable and dedicated dehydrogenative imination catalysts exist,143-

145 in the case of ureation they represent undesirable side products that detrimentally affect

selectivity. As primary amines are required for dehydrogenative ureations that proceed

through isocyanate intermediates, the minimization or elimination of imination further

restricts the suite of suitable catalysts.

Scheme 2.3.3: Metal-promoted pathway for dehydrogenative urea formation from

methanol and amines.

Page 148: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

132

Recently, Kim and Hong described the sole example of catalytic urea formation

directly from methanol and amines (Figure 2.3.1).146 Their precious metal ruthenium

catalyst exhibits high selectivities and excellent yields for the production of symmetric and

unsymmetric ureas, with TONs of up to 190 and 15, respectively, and its applicability

toward reversible147 hydrogen storage has been demonstrated.148 However, the synthesis of

unsymmetric ureas requires high catalyst loadings and a complicated two-step process. To

date, no other base-metal catalysts for the production of ureas from methanol and amines

have been described, with the most closely related examples instead relying on formamide

starting materials133 or isocyanate reagents.149 Base-metal catalysts for dehydrogenative

urea synthesis are particularly desirable for pharmaceutical applications where there are

stringent requirements regarding product separation from toxic metals.150 Previous work in

our laboratories as well as results introduced in the amidation section of this chapter have

verified the catalytic ability of iron-pincer complexes of the type (RPNP)Fe(H)(CO) (RPNP

= N[CH2CH2(PR2)]2, R = iPr, Cy) for both methanol dehydrogenation151 and the

dehydrogenative amidation of alcohols with secondary amines,141 suggesting the

applicability of this class of compounds toward ureation. In this section, the use of complex

1 ((iPrPNP)Fe(H)(CO), Figure 2.3.1) in the first base-metal-catalyzed dehydrogenative

Figure 2.3.1: Ruthenium and iron catalysts for symmetric and unsymmetric urea

formation from the dehydrogenative coupling of methanol and amines.

Page 149: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

133

coupling of methanol and primary amines to selectively form ureas will be discussed. This

system exhibits excellent activities in the absence of any additives and has been employed

to isolate ureas on a scale of several hundred milligrams, corresponding to TONs of up to

160 for symmetric ureas and 176 for unsymmetric ureas.

2.3.2: Results and Discussion

Catalytic Studies

Due to the relative insolubility of 1,3-dicyclohexylurea in most organic solvents

and the corresponding ease of post-reaction isolation, initial investigations into the ability

of 1 to catalyze the formation of symmetric ureas employed methanol and cyclohexylamine

as the substrates. A promising TON of 23 for the formation of 1,3-dicyclohexylurea was

observed using a one to four molar ratio of methanol to amine and 0.5 mol% of 1 in 5 mL

tetrahydrofuran (THF) at 80 °C for 8 hours. Empirical optimization of the reaction resulted

in minimal improvements (Tables A2.9-A2.13) except for increasing the reaction

temperature, which nearly tripled the TON for urea formation to 66 (Table 2.3.1, entry 12).

While the total yield of urea is modest (ca 33%), sterics may play a limiting kinetic role

given the need to attach two amine substrates to the methanol-derived carbonyl carbon of

the urea.

The encouraging preliminary results prompted examination of the substrate scope

for dehydrogenative ureation (Table 2.3.1). The most productive substrates were terminal

alkylamines, with pentylamine giving the highest TON of 160 (80% yield, entry 1). Small

steric changes such as using a branched alkylamine (isobutylamine, entry 2) or elongating

the alkyl chain (heptylamine, entry 3) did not significantly alter catalyst performance.

Likewise, capping the alkylamine chain with a methoxy group (2-methoxyethylamine,

Page 150: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

134

Entry Amine TONc Yield (%)

1 160 80%

2 150 75%

3 147 74%

4 144 72%

5 12d 6%

6b 156e 78%

7 140 70%

8 126 63%

9b 90 45%

10 123 62%

11 116 58%

12b 66 33%

13 22e,f 11%

14 0 --

15 0 --

aReaction conditions: 3 mmol methanol, 12 mmol amine, 0.5

mol% 1, 5 mL THF at 120 °C for 8 hours. Each entry is an

average of two trials unless otherwise indicated. bAverage of

three trials. cBased on yield of isolated urea of >99% purity

(as determined by 1H NMR spectroscopy) unless otherwise

indicated. d >98% purity. e >97% purity. fMixture of isomers,

~75:25 cis:trans.

Table 2.3.1: Dehydrogenative Symmetric Urea

Formation from Methanol and Amines Catalyzed by 1.a

Page 151: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

135

entry 4) afforded good yields. However, switching from a terminal amine (entry 3) to an

internal amine (2-aminoheptane, entry 5) significantly decreased the TON. While

amidation proceeded favorably (TON = 53), it is possible that the sterically hindered

internal amine significantly increased the barrier either for attack by a second equivalent

of amine or for the dehydrogenation of intermediates necessary to form the corresponding

symmetric urea.

Electronic influences were probed using a series of benzylamine derivatives

(entries 6-10) and it was found that the presence of electron-donating substituents such as

methoxy and methyl in the para position of the benzyl moiety enhanced the TON (entries

6 and 7, respectively) compared to unsubstituted benzylamine (entry 8). An electron-

withdrawing substituent such as trifluoromethyl slightly decreased the TON (entry 9).

These substituent changes affect the nucleophilicity of the amine substrate which likely

explains this trend in TON. Attempts to decrease steric hindrance by switching from

benzylamine to 2-phenethylamine had little effect (entries 8 versus 10, respectively),

although it is possible that the increased flexibility in the short carbon chain could

counteract some of the desired steric relief. Entries 11 and 12 demonstrate that while

smaller ring structures can still perform reasonably well, there is significant steric

hindrance that increases rapidly with size, as the change of a cyclopentyl group to a

cyclohexyl group dropped the TON by nearly half. The attempted synthesis of a cyclic urea

(entry 13) from a diamine was successful, although it gave a poor yield compared to its

monoamine counterpart (entry 12). This could be partially due to the cis/trans mixture of

the starting amine (~60:40), as formation of the strained trans product is less favorable

compared to its cis counterpart (though both cis and trans products were observed), thereby

Page 152: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

136

detracting from catalytic performance. However, analogous experiments by Kim and Hong

displayed a similar reduced activity toward diamines that was attributed to amine

coordination effects, which could also be involved here.146 Finally, aniline and its methyl

derivative (entries 14 and 15) reacted poorly under these conditions, giving very little of

the corresponding formamides and no conversion to the desired ureas. This is attributed to

a lack of nucleophilicity from the poor electron donation of the aryl substituent, which has

been a common obstacle for catalytic dehydrogenative coupling reactions using amines.152

Addition of a co-catalytic amount of exogenous base did little to overcome these limitations

(Table A2.14). Overall, 1 represents one of only two examples of metal complexes capable

of catalyzing urea synthesis from alcohols and primary amines. It has the advantage of

containing a cheaper, more abundant base metal while still displaying good yields in the

production of symmetric ureas from a range of substrates, without the formation of imines.

Furthermore, in several cases 1 affords isolated pure urea product on a synthetically useful

scale.

Encouraged by the performance of 1 in the synthesis of symmetric ureas, further

investigations addressed the preparation of unsymmetric ureas,153 which are most prevalent

as key functional groups in pharmaceuticals.93 While unsymmetric ureas can be acquired

via the transamidation of ureas, either by metal-free97 or metal-catalyzed153 means, these

methods have limited scope, can require base or reductant additives, and suffer from the

same toxicity issues in generating the urea starting materials. Kim and Hong accessed these

products from methanol and amines using a one-pot, two-step method that, while effective,

required significant and sequential catalyst loadings (6 mol% total, added in two portions)

and was restricted to benzylamine as a substrate. The direct reaction of methanol with two

Page 153: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

137

different primary amines has been shown to result in a distribution of symmetric and

unsymmetric ureas,146 so our approach focused on selectively forming unsymmetric ureas

from the reactions of formamides and amines (Table 2.3.2). As previously mentioned, the

few metal-catalyzed examples of this process exhibit very low TONs (<50).132-134 While

starting with formamides is not ideal due to cost and ease of synthesis, the potential for

generating these materials from the dehydrogenative coupling of methanol and amines was

Entry R R’ Yieldb (%) Conv.c (%) Sel.d (%)

1 85% 89% 96%

2 86% 93% 92%

3

49% 79% 68%

4 85% 86% 82%

5

11%

28%

78%

6

0% 16% --

Table 2.3.2: Dehydrogenative Unsymmetric Urea Formation from Formamides and

Amines Catalyzed by 1.a

aReaction conditions: 3 mmol formamide, 3 mmol amine, 0.5 mol% 1, 5 mL THF at 120 °C for 16 hours.

Each entry is an average of two trials. bIsolated yield. cBased on formamide consumption. dSelectivity:

percentage unsymmetric urea (compared to symmetric ureas) in final product.

Page 154: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

138

illustrated in the previous section of this chapter and may serve to alleviate some of the

reservations associated with this aspect of the method.

Re-optimization of the reaction conditions for producing unsymmetric rather than

symmetric ureas using 1 resulted in a change of the ratio of starting materials (to 1:1

formamide:amine) and in reaction time from 8 hours to 16 hours (Tables A2.15-A2.17).

Initial experiments involving benzylformamide and cyclohexylamine gave excellent yields

of the desired unsymmetric urea (Table 2.3.2, entry 1) with high selectivity. However, it

was found that some scrambling of the starting formamide or the unsymmetric product (or

both) had occurred to give trace amounts of both symmetric ureas (Scheme 2.3.4). Further

experiments under catalytic conditions showed that scrambling of the starting formamide

can occur in small amounts in the absence of catalyst, but suggested an iron-catalyzed

enhancement of this process (Table A2.18). NMR experiments also indicated that while

hydrogenation of the unsymmetric urea product to the opposing formamide and amine pair

could occur, it was a very minor process (ca 1.5% conversion) under the conditions studied

(10 mol% 1, THF-d8, 120 °C, 16 hours, 1 atm H2; Figure A2.3). Additional NMR

experiments revealed that the urea product can undergo further reactions with both amines

Scheme 2.3.4: Potential scrambling pathways in unsymmetric urea formation.

Page 155: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

139

and formamides in solution, providing yet another pathway to the observed scrambling

(Figures A2.4-A2.6). While insightful, these experiments did not distinguish whether

scrambling of the starting formamide or of the product urea was the predominant process

for producing the undesirable symmetric ureas (Figure A2.7). Catalytic trials involving the

opposing reactant pair of cyclohexylformamide and benzylamine (Table 2.3.2, entry 3)

revealed a significant decrease in selectivity, presumably due to a corresponding increase

in the rate of scrambling compared to unsymmetric ureation.

Efforts to enhance yield through minor substituent manipulation did not have a

significant effect. While 4-methoxybenzylamine and pentylamine were both high-

performing substrates for the formation of symmetric ureas, there was little advantage to

the N-(4-methoxybenzyl)formamide and pentylamine combination over benzylformamide

and cyclohexylamine in either TON or selectivity (Table 2.3.2, entry 2). This indicated that

starting from combinations of formamides and amines that minimized scrambling (for

example, benzylformamide or its derivatives rather than cyclohexylformamide) was far

more important than small electronic or steric changes in substituents. As a result, other

experiments with respect to unsymmetric urea production focused on species that were

reminiscent of medicinally-relevant ureas (entries 4-6).93

Although the initial reaction of isobutylformamide and ethanolamine did not yield

urea, changing the –OH group on the amine to a methoxy group (entry 4) gave excellent

turnover, albeit with poorer selectivity than for formamides containing benzyl derivatives.

The failure of ethanolamine was attributed to its preference for forming an iron-alkoxide

species with the catalyst.141 The production of unsymmetric ureas containing an aniline

moiety on one side is highly desirable in a variety of medical applications including for

Page 156: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

140

antitumor and anticonvulsant agents,91,94,95 however, formanilide was a poor substrate for

dehydrogenative coupling and scrambling to form the more active pentylformamide was

observed as the primary process (entry 5). A similar lack of nucleophilicity is reflected in

the amine of entry 6 as only formamide scrambling was observed with benzylformamide

(a known active substrate). Overall, while the possibility of obtaining unsymmetric ureas

in high yields with good selectivities was demonstrated, it was found that kinetic control

was far more important than qualitative changes in sterics or electronics. As a result, a full

substrate scope would not be as informative and was not performed. Regardless, 1 exhibits

TONs over three times greater than the existing catalysts for this method of obtaining

unsymmetric ureas132-134 and the one-pot two-step method starting from alcohol and

amine,146 and has the added benefits of being both base-metal-catalyzed and additive free.

The occurrence of reagent and product scrambling in the generation of unsymmetric ureas

is a drawback to this application of catalyst 1. However, this issue is not unique to iron and

further understanding of ways to control scrambling outside of reagent selection will

greatly expand the substrate scope and applicability of metal-mediated unsymmetric urea

formation.

Mechanistic Investigations

There are two main pathways proposed for urea formation, as shown in Scheme

2.3.5. Kim and Hong postulated that after initial dehydrogenative amidation (that can

proceed via either a hemiaminal or hemiacetal intermediate), their ruthenium-catalyzed

ureation followed path (a) to generate an aminal, followed by metal-catalyzed

dehydrogenation to yield the urea product.146 Alternatively, isocyanate intermediates

(usually metal-bound) have been implicated in other metal-mediated urea formation

Page 157: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

141

reactions starting from amides,132,136 as well as in the majority of metal-free ureation

cases.96,97 These examples would suggest path (b), whereby urea generation stems from

metal-catalyzed dehydrogenation of a formamide to an isocyanate, followed by

nucleophilic attack of an amine. Mechanistic investigations were therefore performed

using 1 to elucidate whether iron-catalyzed urea formation proceeds via an aminal or an

isocyanate intermediate. Our attempts to make a tetrasubstituted urea species

(R2NC(O)NR’2) from a tertiary amide and a secondary amine were unsuccessful (Figure

A2.8, Table A2.19). This could be due to the inability of the tertiary amide to undergo

dehydrogenation to the isocyanate, although steric hindrance in generating a putative

aminal intermediate cannot be ruled out. Therefore, a similar analysis was performed using

two synthetic routes for the preparation of 1,1,3-tripentylurea (Scheme 2.3.6). While the

NMR-scale reaction of pentylformamide and dipentylamine displayed approximately 80%

conversion to the desired urea,86 the alternative combination of dipentylformamide and

pentylamine did not produce any coupling product (Figures A2.9-A2.10). The main

Scheme 2.3.5: Proposed mechanistic pathways for metal-catalyzed urea formation

starting from methanol and primary amines.

Scheme 2.3.6: Formamide-dependent synthesis of 1,1,3-tripentylurea.

Page 158: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

142

difference between these two routes is the ability of the starting formamide to form an

isocyanate intermediate, as the sterics of the final product are identical. Further evidence

for an isocyanate intermediate was obtained through the reaction of cyclohexylformamide

with 1 in the absence of other reagents. In this reaction the iron-dihydride species

(presumably generated from formamide dehydrogenation) was observed by NMR

spectroscopy, along with cyclohexyl isocyanate (confirmed by GC analysis) as the sole

organic product (Figure A2.11). These results indicate that iron-catalyzed dehydrogenative

ureation most likely proceeds through an isocyanate intermediate as shown in Scheme

2.3.5b.

2.3.3: Summary

Complex 1 displays excellent catalytic ability for urea formation via the

dehydrogenative coupling of methanol and amines. While most precious-metal catalysts

for the coupling of alcohols and primary amines stop at the formamide, this base-metal

species has a rare propensity to undergo further coupling to the urea. The performance of

1 in the synthesis of symmetric ureas from methanol and primary amines rivals that of the

only other (ruthenium catalyzed) example, and its TONs in the synthesis of unsymmetric

ureas from formamides and primary amines are several times better than those of any other

existing catalysts for that synthetic route. In addition, 1 performs these dehydrogenative

reactions without the use of an exogenous base or a hydrogen acceptor, enhancing its

environmental and economic profile. In contrast to the proposed behavior of the ruthenium

catalyst, mechanistic studies suggest that in this case the final ureation step proceeds via

an isocyanate intermediate. Trapping the transient isocyanate species with other reagents

(such as alcohols to give urethanes)154 may provide alternative synthetic pathways to a suite

Page 159: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

143

of desirable compounds starting from dehydrogenative coupling. Despite limitations with

respect to scrambling and steric sensitivity, this works represents yet another example of

the superiority of 1 with respect to methanol activation and subsequent transformation into

high-value products.

Page 160: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

144

2.4: Chapter Summary

In this chapter, the superior ability of the iron pincer complex (iPrPNP)Fe(H)(CO)

(complex 1) in catalytic dehydrogenative coupling reactions for organic synthesis was

established. Complex 1 overcame the inherent energetic challenges to methanol activation

that thus far have limited its more widespread incorporation as a C1 carbon source in

dehydrogenative organic transformations. As a result, 1 emerges as the most active base-

metal catalyst for dehydrogenative amidation, with TONs that exceed its precious metal

competition with respect to secondary amines. In addition, 1 is the first base-metal catalyst

for dehydrogenative urea formation from methanol and primary amines and represents one

of only two catalysts for this method of urea synthesis, both exhibiting comparable TONs.

No base or hydrogen acceptor additives are needed in either amidation or ureation, making

this method of methanol utilization applicable to a wider range of functionalized species.

The significance of these results rests in the pivotal role of formamide and urea

functionalities in medicinal applications and the potential for sourcing such valuable

chemicals from a cheap and abundant carbon source in a more environmentally benign

manner using a non-precious-metal catalyst. That being said, understanding and solving

the severe steric limitations of catalyst 1 (possibly through modifications of the ancillary

ligand environment), improving its reactivity with less nucleophilic substrates, and

developing methods of increasing product selectivity in ureation are key issues that must

be addressed before its more widespread use in formamide and urea production is possible.

Additionally, while the performance of catalyst 1 compares with and often exceeds its

precious-metal competition for select substrates, the activities of base-metal catalysts in

general in this field of application are an order of magnitude worse than those of the best

Page 161: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

145

precious-metal catalysts. Ways of circumventing this problem (such as increased

recyclability of base-metal catalysts) will need to be established before implementation of

these methods on a large scale can be realized.

Page 162: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

146

2.5 References

1 Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies, F.;

Kurr, P.; Kniep, B.-L.; Tovar, M.; Fischer, R.S.; Nørskov, J.S.; Schlögl, R. “The Active Site of Methanol

Synthesis over Cu/ZnO/Al2O3 Industrial Catalysts,” Science, 2012, 336, 893-897.

2 Cañete, B.; Gigola, C.E.; Brignole, N.B. “Synthesis Gas Processes for Methanol Production via CH4

Reforming with CO2, H2O, and O2,”Ind. Eng. Chem. Res., 2014, 53, 7103-7112.

3 Olah, G.; Goeppert, A.; Prakash, G.K.S. Beyond Oil and Gas: The Methanol Economy; VCH, Weinheim:

2006; Chapter 10-13, pp. 168-253.

4 Zhang, Y.; Xioa, J.; Shen, L. “Simulation of Methanol Production from Biomass Gasification in

Interconnected Fluidized Beds,” Ind. Eng. Chem. Res., 2009, 48, 5351-5359.

5 Feng, W.; Ji, P.; Chen, B.; Zheng, D. “Analysis of Methanol Production from Biomass Gasification,” Chem.

Eng. Technol., 2011, 34, 307-317.

6 Holmgren, K.M.; Berntsson, T.; Andersson, E.; Rydberg, T. “System aspects of biomass gasification with

methanol synthesis-Process concepts and energy analysis,” Energy, 2012, 45, 817-828.

7 Xiu, S.; Shahbazi, A. “Bio-oil production and upgrading research: A review,” Renewable and Sustainable

Energy Reviews, 2012, 16, 4406-4414.

8 Palo, D.R.; Dagle, R.A.; Holladay, J.D. “Methanol Steam Reforming for Hydrogen Production,” Chem.

Rev., 2007, 107, 3992-4021.

9 IHS Markit Report, 2015, http://www.methanol.org/wp-content/uploads/2016/07/Marc-Alvarado-Global-

Methanol-February-2016-IMPCA-for-upload-to-website.pdf

10 IHS Chemical Bulletin, 2016, http://www.methanol.org/wp-content/uploads/2016/07/IHS-

ChemicalBulletin-Issue3-Alvarado-Jun16.pdf

11 IHS Markit Report, 2012, http://www.formacare.org/applications/

12 Franz, A.W.; Kronemayer, H.; Pfeiffer D.; Pilz, R.D.; Reuss, G.; Disteldorf, W.; Gamer, A.O.; Hilt, A.

Ullmann’s Encyclopedia of Industrial Chemistry, 6th edition; VCH, Weinheim: 2001; pp. 619-653.

13 https://patents.google.com/patent/US7381851

14 Qian, M.; Liauw, M.A.; Emig, G. “Formaldehyde synthesis from methanol over silver catalysts,” Applied

Catalysis A: General, 2003, 238, 211-222.

15 Ley, S.V.; Madin, A. Comprehensive Organic Syntheses; Pergamon: Oxford, 1991; Vol. 7, Chapter 2.7-

2.8, pp. 348-389.

16 Gunanathan, C.; Milstein, D. “Applications of Acceptorless Dehydrogenation and Related Transformations

in Chemical Synthesis,” Science, 2013, 341, 1229712 (1-11).

17 Crabtree, R.H. “Homogeneous Transition Metal Catalysis of Acceptorless Dehydrogenative Alcohol

Oxidation: Applications in Hydrogen Storage and to Heterocycle Synthesis,” Chem. Rev., 2017, 117, 9228-

9246.

Page 163: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

147

18 Dobereiner, G.E.; Crabtree, R.H. “Dehydrogenation as a Substrate –Activating Strategy in Homogeneous

Transition-Metal Catalysis,” Chem. Rev., 2010, 110, 681-703.

19 Nielsen, M. “Hydrogen Production by Homogenous Catalysis: Alcohol Acceptorless Dehydrogenation,”

Hydrogen Production and Remediation of Carbon and Pollutants; Lichtfouse, E.; Schwarzbauer, J.; Robert,

D., Eds.; Springer: Heidelberg, 2015.

20 Werkmeister, S.; Neumann, J.; Junge, K.; Beller, M. “Pincer-Type Complexes for Catalytic

(De)Hydrogenation and Transfer (De)-Hydrogenation Reactions: Recent Progress,” Chem. Eur. J., 2015, 21,

12226-12250.

21 Gunanathan, C.; Milstein, D. “Bond Activation and Catalysis by Ruthenium Pincer Complexes,” Chem.

Rev., 2014, 114, 12024-12087.

22 Fujita, K.; Yamaguchi, R. “Cp*Ir Complex-Catalyzed Hydrogen Transfer Reactions Directed toward

Environmentally Benign Organic Synthesis,” Synlett, 2005, 560-571.

23 Guillena, G.; Ramón, D.J.; Yus, M. “Alcohols as Electrophiles in C-C Bond-Forming Reactions: The

Hydrogen Autotransfer Process,” Angew. Chem. Int. Ed., 2007, 46, 2358-2364.

24 Hamid, M.H.S.A.; Slatford, P.A., Williams, J.M.J. “Borrowing Hydrogen in the Activation of Alcohols,”

Adv. Synth. Catal., 2007, 349, 1555-1575.

25 Nixon, T.D.; Whittlesey, M.K.; Williams, J.M.J. “Transition metal catalyzed reactions of alcohols using

borrowing hydrogen methodology,” Dalton Trans., 2009, 753-762.

26 Choi, J.; MacArthur, A.H.R.; Brookhart, M.; Goldman, A.S. “Dehydrogenation and Related Reactions

Catalyzed by Iridium Pincer Complexes,” Chem. Rev., 2011, 111, 1761-1779.

27 Morales-Morales, D.; Redón, R.; Wang, Z.; Lee, D.W.; Yung, C.; Magnuson, K.; Jensen, C.M. “Selective

dehydrogenation of alcohols and diols catalyzed by a dihydrido iridium PCP pincer complex,” Can. J. Chem.,

2001, 79, 823-829.

28 Levy, R.; Azerraf, C.; Gelman, D.; Rueck-Braun, K.; Kapoor, P.N. “Cyclometalated phosphine-based

Ir(III) pincer complex as a catalyst for Oppenauer –type oxidation of alcohols,” Catal. Commun., 2009, 11,

298-301.

29 Fujii, T.; Saito, Y. “Thermocatalytic Formation of Molecular Hydrogen and Cyclo-octene from Cyclo-

octane by Rhodium Complexes," J. Chem. Soc., Chem. Commun., 1990, 757-758.

30 Lin, W.-H.; Chang, H.-F. “A study of ethanol dehydration reaction in a palladium membrane reactor,”

Catal. Today, 2004, 97, 181-188.

31 Moran, J.; Preetz, A.; Mesch, R.A.; Krische, M.J. “Iridium-catalysed direct C-C coupling of methanol and

allenes,” Nat. Chem., 2011, 3, 287-290.

32 Pattabiraman, V.R.; Bode, J.W. “Rethinking amide bond synthesis,” Nature, 2011, 480, 471-479.

33 Lanigan, R.M.; Starkov, P.; Sheppard, T.D. “Direct Synthesis of Amides from Carboxylic Acids and

Amines Using B(OCH2CF3)3,” J. Org. Chem., 2013, 78, 4512-4523.

34 Lundberg, H.; Tinnis, F.; Selander, N.; Adolfsson, H. “Catalytic amide formation from non-activated

carboxylic acids and amines,” Chem. Soc. Rev., 2014, 43, 2714-2742.

35 Montalbetti, C.; Falque, V. “Amide bond formation and peptide coupling,” Tetrahedron, 2005, 61, 10827-

10852.

Page 164: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

148

36 Lang, S.; Murphy, J.A. “Azide rearrangements in electron-deficient systems,” Chem. Soc. Rev., 2006, 35,

146-156.

37 Kolakowski, R.V.; Shangguan; N.; Sauers, R.R.; Williams, L.J. “Mechanism of Thio Acid/Azide

Amidation,” J. Am. Chem. Soc., 2006, 128, 5695-5702.

38 Shimizu, K.; Ohshima; K.; Satsuma, A. “Direct Dehydrogenative Amide Synthesis from Alcohols and

Amines Catalyzed by γ-Alumina Supported Silver Cluster,” Chem. Eur. J., 2009, 15, 9977-9980.

39 Xu, B.; Madix, R.J.; Friend, C.M. “Dual-Function of Alcohols in Gold-Mediated Selective Coupling of

Amines and Alcohols,” Chem. Eur. J., 2012, 18, 2313-2318.

40 Tanaka, S.; Minato, T.; Ito, E.; Hara, M.; Kim, Y.; Yamamoto, Y.; Asao, N. “Selective Aerobic Oxidation

of Methanol in the Coexistence of Amines by Nanoporous Gold Catalysts: Highly Efficient Synthesis of

Formamides,” Chem. Eur. J., 2013, 19, 11832-11836.

41 Ishida, T.; Haruta, M. “N-Formylation of Amines via the Aerobic Oxidation of Methanol over Supported

Gold Nanoparticles,” ChemSusChem, 2009, 2, 538-541.

42 Preedasuriyachai, P.; Kitahara, H.; Chavasiri, W.; Sakurai, H. “N-Formylation of Amines Catalyzed by

Nanogold under Aerobic Oxidation Conditions with MeOH or Formalin,” Chem. Lett., 2010, 39, 1174-1176.

43 Tumma, H.; Nagaraju, N.; Reddy, K.V. “A facile method for the N-formylation of primary and secondary

amines by liquid phase oxidation of methanol in the presence of hydrogen peroxide over basic copper

hydroxyl salts,” J. Mol. Catal. A.: Chem., 2009, 310, 121-129.

44 Schleker, P.P.M.; Honeker, R.; Klankermayer, J.; W. Leitner, W. “Catalytic Dehydrogenative Amide and

Ester Formation with Rhenium-Triphos Complexes,” Chem. Cat. Chem., 2013, 5, 1762-1764.

45 Zweifel, T.; Naubron, J.-V.; Grützmacher, H. “Catalyzed Dehydrogenative Coupling of Primary Alcohols

with Water, Methanol, or Amines,” Angew. Chem. Int. Ed., 2009, 48, 559-563.

46 Naota, T.; Murahashi, S.-I. “Ruthenium-Catalyzed Transformations of Amino Alcohols to Lactams,”

Synlett., 1991, 693-694.

47 Gunanathan, C.; Ben-David, Y.; Milstein, D. “Direct Synthesis of Amides from Alcohols and Amines with

Liberation of H2,” Science, 2007, 317, 790-792.

48 Nordstrøm, L.U.; Vogt, H.; Madsen, R. “Amide Synthesis from Alcohols and Amines by the Extrusion of

Dihydrogen,” J. Am. Chem. Soc., 2008, 130, 17672-17673.

49 Watson, A.J.A.; Maxwell, A.C.; Williams, J.M.J. “Ruthenium-Catalyzed Oxidation of Alcohols into

Amides,” Org. Lett., 2009, 11, 2667-2670.

50 Dam, J.H.; Osztrovszky, G.; Nordstrøm, L.U.; Madsen, R. “Amide Synthesis from Alcohols and Amines

Catalyzed by Ruthenium N-Heterocyclic Carbene Complexes,” Chem.Eur. J., 2010, 16, 6820-6827.

51 Zhang, Y.; Chen, C.; Ghosh, S.C.; Li, Y.; Hong, S.H. “Well-Defined N-Heterocyclic Carbene Based

Ruthenium Catalysts for Direct Amide Synthesis from Alcohols and Amines,” Organometallics, 2010, 29,

1374-1378.

52 Schley, N.D.; Dobereiner; G.E.; Crabtree, R.H. “Oxidative Synthesis of Amides and Pyrroles via

Dehydrogenative Alcohol Oxidation by Ruthenium Diphosphine Diamine Complexes,” Organometallics,

2011, 30, 4174-4179.

Page 165: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

149

53 Prades, A.; Peris, E.; Albrecht, M. “Oxidations and Oxidative Couplings Catalyzed by Triazolylidene

Ruthenium Complexes,” Organometallics, 2011, 30, 1162-1167.

54 Chen, C.; Zhang, Y.; Hong, S.H. “N-Heterocyclic Carbene Based Ruthenium-Catalyzed Direct Amide

Synthesis from Alcohols and Secondary Amines: Involvement of Esters,” J. Org. Chem., 2011, 76, 10005-

10010.

55 Prechtl, M.H.G.; Wobser, K.; Theyssen, N.; Ben-David, Y.; Milstein, D.; Leitner, W. “Direct coupling of

alcohols to form esters and amides with evolution of H2 using in situ formed ruthenium catalysts,” Catal. Sci.

and Technol., 2012, 2, 2039-2042.

56 Srimani, D.; Balaraman, E.; Hu, P.; Ben-David, Y.; Milstein, D. “Formation of Tertiary Amides and

Dihydrogen by Dehydrogenative Coupling of Primary Alcohols with Secondary Amines Catalyzed by

Ruthenium Bipyridine-Based Pincer Complexes,” Adv. Synth. Catal., 2013, 355, 2525-2530.

57 Malineni, J.; Merkens, C.; Keul, H.; Möller, M. “An efficient N-heterocyclic carbene based ruthenium-

catalyst: Application towards the synthesis of esters and amides,” Catal. Commun., 2013, 40, 80-83.

58 Ortega, N.; Richter, C.; Glorius, F. “N-Formylation of Amines by Methanol Activation,” Org. Lett., 2013,

15, 1776-1779.

59 Oldenhuis, N.J.; Dong, V.M.; Guan, Z. “Catalytic acceptorless dehydrogenations: Ru-MACHO catalyzed

construction of amides and imines,” Tetrahedron, 2014, 70, 4213-4218.

60 Kim, K.; Kang, B.; Hong, S.H. “N-Heterocyclic carbene-based well-defined ruthenium hydride complexes

for direct amide synthesis from alcohols and amines under base-free conditions,” Tetrahedron, 2015, 71,

4565-4569.

61 Xie, X.; Huynh, H.V. “Tunable Dehydrogenative Amidation versus Amination Using a Single Ruthenium-

NHC Catalyst,” ACS Catal., 2015, 5, 4143-4151.

62 Kang, B.; Hong, S.H. “Hydrogen Acceptor- and Base-Free N-Formylation of Nitriles and Amines using

Methanol as C1 Source,” Adv. Synth. Catal., 2015, 357, 834-840.

63 Selvamurugan, S.; Ramachandran, R.; Prakash, G.; Viswanathamurthi, P.; Malecki, J.G.; Endo, A.

“Ruthenium(II) carbonyl complexes containing bidentate 2-oxo-1,2-dihydroquinoline-3-carbaldehyde

hydrazone ligands as efficient catalysts for catalytic amidation reaction,” J. Organomet. Chem., 2016, 803,

119-127.

64 Zultanski, S.L.; Zhao, J.; Stahl, S.S. “Practical Synthesis of Amides via Copper/ABNO-Catalyzed Aerobic

Oxidative Coupling of Alcohols and Amines,” J. Am. Chem. Soc., 2016, 138, 6416-6419.

65 Chakraborty, S.; Gellrich, U.; Diskin-Posner, Y.; Leitus, G.; Avram, L.; Milstein, D. “Manganese-

Catalyzed N-Formylation of Amines by Methanol Liberating H2: A Catalytic and Mechanistic Study,”

Angew. Chem. Int. Ed., 2017, 56, 4229-4233.

66 Kumar, A.; Espinosa-Jalapa, N.A.; Leitus, G.; Diskin-Posner, Y.; Avram, L.; Milstein, D. “Direct Synthesis

of Amides by Dehydrogenative Coupling of Amines with either Alcohols or Esters: Manganese Pincer

Complex as Catalyst,” Angew. Chem. Int. Ed., 2017, 56, 14992-14996.

67 Chen, B.-C.; Bednarz, M.S.; Zhao, R.; Sundeen, J.E.; Chen, P.; Shen, Z.; Skoumbourdis, A.P.; Barrish,

J.C. “A new facile method for the synthesis of 1-arylimidizole-5-carboxylates,” Tetrahedron Lett., 2000, 41,

5453-5456.

Page 166: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

150

68 Kobayashi, K.; Nagato, S.; Kwakita, M.; Morikawa, O.; Konishi, H. “Synthesis of 1-Formyl-1,2-

dihydroquinoline Derivatives by Lewis Acid-Catalyzed Cyclization of o-(1-Hydroxy-2-alkenyl)phenyl

Isocyanides,” Chem. Lett., 1995, 24, 575-576.

69 Tlili, A.; Blondiaux, E.; Frogneux, X.; Cantat, T. “Reductive functionalization of CO2 with amines: an

entry to formamide, formamidine, and methylamine derivatives,” Green Chem., 2015, 17, 157-168.

70 Downie, I.M.; Earle, M.J.; Heaney, H.; Shuhaibar, K.F. “Vilsmeier formylation and glyoxylation reactions

of nucleophilic aromatic compounds using pyrophosphoryl chloride,” Tetrahedron, 1993, 49, 4015-4034.

71 Han, Y.; Cai, L. “An Efficient and Convenient Synthesis of Formamidines,” Tetrahedron Lett., 1997, 38,

5423-5426.

72 Jones, S.; Warner, C.J.A. “Trichlorosilane mediated asymmetric reductions of the C=N bond,” Org.

Biomol. Chem., 2012, 10, 2189-2200.

73 Jagtap, S.B.; Tsogoeva, S.B. “First enantioselective organocatalytic allylation of simple aldimines with

allyltrichlorosilane,” Chem. Commun., 2006, 4747-4749.

74 Wuts, P.G.M., Greene’s Protective Groups in Organic Synthesis, 5th ed., Wiley, Hoboken, N.J., 2014, pp.

991.

75 Kozak, M. “Comparison of initiation of protein synthesis in procaryotes, eucaryotes, and organelles,”

Microbiol. Rev., 1983, 47, 1-45.

76 Wiśniewski, J.R.; Zougman, A.; Mann, M. “N-Formylation of lysine is a widespread post-translational

modification of nuclear proteins occurring at residues involved in regulation of chromatin function,” Nucleic

Acids Res., 2008, 36, 570-577.

77 Hett, R.; Fang, Q.K.; Gao, Y.; Wald, S.A.; Senanayake, C.H. “Large-Scale Synthesis of Enantio- and

Diastereomerically Pure (R,R)-Formoterol,” Org. Process Res. Dev., 1998, 2, 96-99.

78 Olah, G.A.; Ohannesian, L.; Arvanagh, M. “Formylating Agents,” Chem. Rev., 1987, 87, 671-686.

79 Gerack, C.J.; McElwee-White, L. “Formylation of Amines,” Molecules, 2014, 19, 7689-7713.

80 Kothandaraman, J.; Kar, S.; Sen, R.; Goeppert, A.; Olah, G.A.; Prakash, G.K.S. “Efficient Reversible

Hydrogen Carrier System Based on Amine Reforming of Methanol,” J. Am. Chem. Soc., 2017, 139, 2549-

2552.

81 Chakraborty, S.; Lagaditis, P.O.; Förster, M.; Bielinski, E.A.; Hazari, N.; Holthausen, MC.; Jones, W.D.;

Schneider, S. “Well-Defined Iron Catalysts for the Acceptorless Reversible Dehydrogenation-Hydrogenation

of Alcohols and Ketones,” ACS Catal., 2014, 4, 3994-4003.

82 Bielinski, E.A.; Förster, M.; Zhang, Y.; Bernskoetter, W.H.; Hazari, N.; Holthausen, M.C. “Base-Free

Methanol Dehydrogenation Using a Pincer-Supported Iron Compound and Lewis Acid Co-Catalyst,” ACS

Catal., 2015, 5, 2404-2415.

83 Alberico, E.; Sponholz, P.; Cordes, C.; Nielsen, M.; Drexler, H.-J.; Baumann, W.; Junge, H.; Beller, M.

“Selective Hydrogen Production from Methanol with a Defined Iron Pincer Catalyst under Mild Conditions,”

Angew. Chem. Int. Ed., 2013, 52, 14162-14166.

84 Jayarathne, U.; Zhang, Y.; Hazari, N.; Bernskoetter, W.H. “Selective Iron-Catalyzed Deaminative

Hydrogenation of Amides,” Organometallics, 2017, 36, 409-416.

Page 167: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

151

85 For a prior intramolecular example, see: Peña-López, M.; Neumann, H.; and Beller, M. “Iron(II) Pincer-

Catalyzed Synthesis of Lactones and Lactams through a Versatile Dehydrogenative Domino Sequence,”

Chem. Cat. Chem, 2015, 7, 865-871.

86 See Appendix 2 for experimental details.

87 Abad, A.; Loveras, J.; Michelena, A. “Nitrogen fertilizer and foliar urea effects on drum wheat yield and

quality and on residual soil nitrate in irrigated Mediterranean conditions,” Field Crops Research, 2004, 87,

257-269.

88 Kocyigit-Kaymakcioglu, B.; Celen, A.O.; Tabanca, N.; Ali, A.; Khan, S.I.; Khan, I.A.; Wedge, D.E.

“Synthesis and Biological Activity of Substituted Urea and Thiourea Derivatives Containing 1,2,4-Triazole

Moieties,” Molecules, 2013, 18, 3562-3576.

89 Lam, P.Y.S.; Jadhav, P.K.; Eyermann, C.J.; Hodge, C.N.; Ru, Y.; Bacheler, L.T.; Meek, J.L.; Otto, M.J.;

Rayner, M.M.; Wong, Y.N.; Chang, C.-W.; Weber, P.C.; Jackson, D.A.; Sharpe, T.R.; Erickson-Viitanen, S.

“Rational design of potent, bioavailable, nonpeptide cyclic ureas as HIV protease inhibitors,” Science, 1994,

263, 380-384.

90 Wan, J.-P.; Zhou, J.; Mao, H.; Pan, Y.-J.; Wu, A.-X. “A new diastereoselective multicomponent, one-pot

strategy for the synthesis of 3-substituted isoindolinones via efficient C-C bond formation,” Tetrahedron,

2008, 64, 11115-11123.

91 Sikka, P.; Sahu, J.K.; Mishra, A.K.; Hashim, S.R. “Role of Aryl Urea Containing Compounds in Medicinal

Chemistry,” Med. Chem., 2015, 5, 479-483.

92 Matsuda, K. “ACAT Inhibitors as Antiatherosclerotic Agents: Compounds and Mechanisms,” Medicinal

Research Reviews, 1994, 14, 271-305.

93 Gallou, I. “UNSYMMETRICAL UREAS. SYNTHETIC METHODOLOGIES AND APPLICATION IN

RATIONAL DRUG DESIGN,” Org. Prep. Proced. Int., 2007, 39, 355-383.

94 Bankston, D.; Dumas, J.; Natero, R.; Riedl, B.; Monahan, M.-K.; Sibley, R. “A Scaleable Synthesis of

BAY 43-9006: A Potent Raf Kinase Inhibitor for the Treatment of Cancer,” Org. Proc. Res. Dev., 2002, 6,

777-781.

95 Bergman, J.A.; Woan, K.; Perez-Villarroel, P.; Villagra, A.; Sotomayor, E.M.; Kozikowski, A.P.

“Selective Histone Deacetylase 6 Inhibitors Bearing Substituted Urea Linkers Inhibit Melanoma Cell

Growth,” J. Med. Chem., 2012, 55, 9891-9899.

96 Babad, H.; Zeiler, A.G. “The Chemistry of Phosgene,” Chem. Rev., 1973, 73, 75-91.

97 Bigi, F.; Maggi, R.; Sartori, G. “Selected syntheses of ureas through phosgene substitutes,” Green Chem.,

2000, 2, 140-148.

98 Kulkarni, A.R.; Garai, S.; Thakur, G.A. “Scalable, One-Pot, Microwave-Accelerated Tandem Synthesis of

Unsymmetrical Urea Derivatives,” J. Org. Chem., 2017, 82, 992-999.

99 Srivastava, S.C.; Shrimal, A.K.; Srivastava, A. “Reactions of (η-methylcyclopentadienyl)manganese

tricarbonyl with primary amines,” J. Organomet. Chem., 1991, 414, 65-69.

100 Shelton, P.A.; Zhang, Y.; Nguyen, T.H.H.; McElwee-White, L. “NaIO4-oxidized carbonylation of amines

to ureas,” Chem. Commun., 2009, 947-949.

Page 168: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

152

101 Nudelman, N.S.; Lewkowicz, E.S.; Pérez, D.G. “One-Pot Synthesis of N,N,N’,N’-Tetrasubstituted Ureas

and Oxomalonamides by Oxidative Carbonylation of Lithium Amides at Atmospheric Pressure,” Synthesis,

1990, 917-920.

102 Franz, R.A.; Applegath, F. “A New Urea Synthesis. I. The Reaction of Ammonia, Carbon Monoxide, and

Sulfur,” J. Org. Chem., 1962, 26, 3304-3305.

103 Mizuno, T.; Mihara, M.; Iwai, T.; Ito, T.; Ishino, Y. “Practical Synthesis of Urea Derivatives from Primary

Amines, Carbon Monoxide, Sulfur, and Oxygen under Mild Conditions,” Synthesis, 2006, 2825-2830.

104 Mizuno, T.; Mihara, M.; Nakai, T.; Iwai, T.; Ito, T. “Solvent-Free Synthesis of Urea Derivatives from

Primary Amines and Sulfur under Carbon Monoxide and Oxygen at Atmospheric Pressure,” Synthesis, 2007,

3135-3140.

105 Mei, J.; Yang, Y.; Xue, Y.; Lu, S. “Selective formation of unsymmetric ureas by selenium-catalyzed

oxidative-reductive carbonylation with CO,” J. Mol.Catal. A: Chem., 2003, 191, 135-139.

106 Sonoda, N. “Selenium assisted carbonylation with carbon monoxide,” Pure & Appl. Chem., 1993, 65,

699-706.

107 Shi, F.; Deng, Y.; SiMa, T.; Peng, J.; Gu, Y.; Qiao, B. “Alternatives to Phosgene and Carbon Monoxide:

Synthesis of Symmetric Urea Derivatives with Carbon Dioxide in Ionic Liquids,” Angew. Chem. Int. Ed.,

2003, 42, 3257-3260.

108 Ogura, H.; Takeda, K.; Tokue, R.; Kobayashi, T. “A Convenient Direct Synthesis of Ureas from Carbon

Dioxide and Amines,” Synthesis, 1978, 394-396.

109 Yamazaki, N.; Iguchi, T.; Higashi, F. “Studies on the reactions of the N-phosphonium salts of pyridines-

XV: Direct carbonylation of amines with carbon dioxide by a hydrolysis-dehydration reaction with

phosphorous compounds,” Tetrahedron, 1975, 31, 3031-3034.

110 Giannoccaro, P.; Nobile, C.F.; Mastrorilli, P.; Ravasio, N. “Oxidative carbonylation of aliphatic amines

catalysed by nickel complexes,” J. Organomet. Chem., 1991, 419, 251-258.

111 Alper, H.; Vasapollo, G.; Hartstock, V.; Miekuz, M. “Conversion of Primary Amines to Carbamate Esters

Using Palladium Chloride and Di-tert-butyl Peroxide. Double Carbonylation of Secondary Amines,”

Organometallics, 1987, 6, 2391-2393.

112 Giannoccaro, P. “Palladium-catalysed N,N’-disubstituted urea synthesis by oxidative carbonylation of

amines under CO and O2 at atmospheric pressure,” J. Organomet. Chem., 1987, 336, 271-278.

113 Smirnova, E.S.; Muñoz Molina, J.M.; Johnson, A.; Bandeira, N.A.G.; Bo, C.; Echavarren, A.M.

“Polynuclear Gold [AuI]4, [AuI]8, and Bimetallic [AuI4AgI] Complexes: C-H Functionalization of Carbonyl

Compounds and Homogeneous Carbonylation of Amines,” Angew. Chem. Int. Ed., 2016, 55, 7487-7491.

114 Qian, F.; McCusker, J.; Zhang, Y.; Main, A.D.; Chlebowski, M.; Kokka, M.; McElwee-White, L.

“Catalytic Oxidative Carbonylation of Primary and Secondary Diamines to Cyclic Ureas. Optimization and

Substituent Studies,” J. Org. Chem., 2002, 67, 4086-4092.

115 Guan, Z.-H.; Lei, H.; Chen, M.; Ren, Z.-H.; Bai, Y.; Wang, Y.-Y. “Palladium-Catalyzed Carbonylation

of Amines: Switchable Approaches to Carbamates and N,N’-Disubstituted Ureas,” Adv. Synth. Catal., 2012.

354, 489-496.

116 Park, J.H.; Yoon, J.C.; Chung, Y.K. “Cobalt/Rhodium Heterobimetallic Nanoparticle-Catalyzed

Oxidative Carbonylation of Amines in the Presence of Carbon Monoxide and Molecular Oxygen to Ureas,”

Adv. Synth. Catal., 2009, 351, 1233-1237.

Page 169: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

153

117 McCusker, J.; Main, A.D.; Johnson, K.S.; Grasso, C.A.; McElwee-White, L. “W(CO)6-Catalyzed

Oxidative Carbonylation of Primary Amines to N,N’-Disubstituted Ureas in Single or Biphasic Solvent

Systems. Optimization and Functional Group Compatibility Studies,” J. Org. Chem., 2006, 65, 5216-5222.

118 Orito, K.; Miyazawa, M.; Nakamura, T.; Horibata, A.; Ushito, H.; Nagasaki, H.; Yuguchi, M.; Yamashita,

S.; Yamazaki, T.; Tokuda, M. “Pd(OAc)2-Catalyzed Carbonylation of Amines,” J. Org. Chem., 2006, 71,

5951-5958.

119 Leung, T.W.; Dombek, B.D. “Oxidative Carbonylation of Amines Catalysed by Metallomacrocyclic

Compounds,” J. Chem. Soc., Chem. Commun., 1992, 205-206.

120 Zoeckler, M.T.; Laine, R.M. “Homogeneous catalytic formation of carbon-nitrogen bonds. 2. Catalytic

activation of the silicon-nitrogen bond,” J. Org. Chem., 1983, 48, 2539-2543.

121 Shi, F.; Zhang, Q.; Ma, Y.; He, Y.; Deng, Y. “From CO Oxidation to CO2 Activation: An Unexpected

Catalytic Activity of Polymer-Supported Nanogold,” J. Am. Chem. Soc., 2005, 127, 4182-4183.

122 Ion, A.; Parvulescu, V.; Jacobs, P.; Vos, D.D. “Synthesis of symmetrical or asymmetrical urea compounds

from CO2 via base catalysis,” Green Chem., 2007, 9, 158-161.

123 Tamura, M.; Ito, K.; Nakagawa, Y.; Tomishige, K. “CeO2-catalyzed direct synthesis of dialkylureas from

CO2 and amines,” J. Catal., 2016, 343, 75-85.

124 Xu, M.; Jupp, A.R.; Stephan, D.W. “Stoichiometric Reactions of CO2 and Indium-Silylamides and

Catalytic Synthesis of Ureas,” Angew. Chem. Int. Ed., 2017, 56, 14277-14281.

125 For a recent comprehensive review of metal-free and metal-catalyzed CO2 to urea transformations, see:

Wang, H.; Xin, Z.; Li, Y., “Synthesis of Ureas from CO2,” Top. Curr. Chem., 2017, 375, 49.

126 An, H.; Zhang, L.; Zhao, X., Wang, Y. “Effect of preparation conditions on the catalytic performance of

Cu-Fe/ZrO2 for the synthesis of DPU from aniline and CO2,” Chem. Eng. J., 2014, 255, 266-273.

127 Sun, D.-L.; Ye, J.-H.; Fang, Y.-X.; Chao, Z.-S. “Green Synthesis of N,N’-Dialkylureas from CO2 and

Amines using Metal Salts of Oxalates as Catalysts,” Ind. Eng. Chem. Res., 2016, 55, 64-70.

128 Kimura, T.; Kamata, K.; Mizuno, N. “A Bifunctional Tungstate Catalyst for Chemical Fixation of CO2 at

Atmospheric Pressure,” Angew. Chem. Int. Ed., 2012, 51, 6700-6703.

129 Fournier, J.; Bruneau, C.; Dixneuf, P.H.; Lécolier, S. “Ruthenium-Catalyzed Synthesis of Symmetrical

N,N’-Dialkylureas Directly from Carbon Dioxide and Amines,” J. Org. Chem., 1991, 56, 4456-4458.

130 For a review of heterogeneous catalysts for CO2 to cyclic urea transformations, see: Tamura, M.; Honda,

M.; Nakagawa, Y.; Tomishige, K. “Direct conversion of CO2 with diols, aminoalcohols and diamines to

cyclic carbonates, cyclic carbamates and cyclic ureas using heterogeneous catalysts,” J. Chem. Technol.

Biotechnol., 2014, 89, 19-33.

131 Kotachi, S.; Tsuji, Y.; Kondo, T.; Watanabe, Y. “Ruthenium catalysed N,N’-Diarylurea Synthesis from

N-Aryl substituted Formamides and Aminoarenes,” J. Chem. Soc., Chem. Commun., 1990, 549-550.

132 Kondo, T.; Kotachi, S.; Tsuji, Y.; Watanabe, Y.; Mitsudo, T. “Novel Ruthenium-Complex-Catalyzed

Synthesis of Ureas from Formamides and Amines,” Organometallics, 1997, 16, 2562-2570.

133 Kumar, G.S.; Kumar, R.A.; Kumar, P.S.; Reddy, N.V.; Kumar, K.V.; Kantam, M.L.; Prabhakar, S.;

Reddy, K.R. “Copper catalyzed oxidative coupling of amines with formamides: a new approach for the

synthesis of unsymmetrical urea derivatives,” Chem. Commun., 2013, 49, 6686-6688.

Page 170: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

154

134 Krishnakumar, V.; Chatterjee, B.; Gunanathan, C. “Ruthenium-Catalyzed Urea Synthesis by N-H

Activation of Amines,” Inorg. Chem., 2017, 56, 7278-7284.

135 Jiang, H.; Lin, A.; Zhu, C.; Cheng, Y. “Copper-catalyzed C-N bond formation through C-H/N-H

activation: a novel approach to the synthesis of multisubstituted ureas,” Chem. Commun., 2013, 49, 819-821.

136 Li, X.; Li, B.; You, J.; Lan, J. “Copper-catalysed oxidative C-H/N-H cross-coupling between formamides

and amides through chelation-assisted N-H activation,” Org. Biomol. Chem., 2013, 11, 1925-1928.

137 For a representative example, see: Gunanathan, C.; Ben-David, Y.; Milstein, D. “Direct Synthesis of

Amides from Alcohols and Amines with Liberation of H2,” Science, 2007, 317, 790-792.

138 Ortega, N.; Richter, C.; Glorius, F. “N-Formylation of Amines by Methanol Activation,” Org. Lett., 2013,

15, 1776-1779.

139 Kang, B.; Hong, S.H. “Hydrogen Acceptor- and Base-Free N-Formylation of Nitriles and Amines using

Methanol as a C1 source,” Adv. Synth. Catal., 2015, 357, 834-840.

140 Chakraborty, S.; Gellrich, U.; Diskin-Posner, Y.; Leitus, G.; Avram, L.; Milstein, D. “Manganese-

Catalyzed N-Formylation of Amines by Methanol Liberating H2: A Catalytic and Mechanistic Study,”

Angew. Chem. Int. Ed., 2017, 56, 4229-4233.

141 Lane, E.M.; Uttley, K.B.; Hazari, N.; Bernskoetter, W.H. “Iron-Catalyzed Amide Formation from the

Dehydrogenative Coupling of Alcohols and Secondary Amines,” Organometallics, 2017, 36, 2020-2025.

142 Kumar, A.; Espinosa-Jalapa, N.A.; Leitus, G.; Diskin-Posner, Y.; Avram, L.; Milstein, D. “Direct

Synthesis of Amides by Dehydrogenative Coupling of Amines with either Alcohols or Esters: Manganese

Pincer Complex as Catalyst,” Angew. Chem. Int. Ed., 2017, 56, 14992-14996.

143 Gnanaprakasam, B.; Zhang, J.; Milstein, D. “Direct Synthesis of Imines from Alcohols and Amines with

Liberation of H2,” Angew. Chem. Int. Ed., 2010, 49, 1468-1471.

144 Mukherjee, A.; Nerush, A.; Leitus, G.; Shimon, L.J.W.; Ben-David, Y.; Espinosa-Jalapa, N.A.; Milstein,

D. “Manganese-Catalyzed Environmentally Benign Dehydrogenative Coupling of Alcohols and Amines to

Form Aldimines and H2: A Catalytic and Mechanistic Study,” J. Am. Chem. Soc., 2016, 138, 4298-4301.

145 Saha, B.; Wahidur Rahaman, S.M.; Daw, P.; Sengupta, G.; Bera, J.K. “Metal-Ligand Cooperation on a

Diruthenium Platform: Selective Imine Formation through Acceptorless Dehydrogenative Coupling of

Alcohols with Amines,” Chem. Eur. J., 2014, 20, 6542-6551.

146 Kim, S.K.; Hong, S.H. “Ruthenium-Catalyzed Urea Synthesis Using Methanol as the C1 Source,” Org.

Lett., 2016, 18, 212-215.

147 Balaraman, E.; Ben-David, Y.; Milstein, D. “Unprecedented catalytic hydrogenation of urea derivatives

to amines and methanol,” Angew. Chem. Int. Ed., 2011, 50, 11702-11705.

148 Kothandaraman, J.; Kar, S.; Sen, R.; Goeppert, A.; Olah, G.A.; Prakash, G.A.S. “Efficient Reversible

Hydrogen Carrier System Based on Amine Reforming of Methanol,” J. Am. Chem. Soc., 2017, 139, 2549-

2552.

149 Huang, X.; Zhuang, T.; Kates, P.A.; Gao, H.; Chen, X.; Groves, J.T. “Alkyl Isocyanates via Manganese-

Catalyzed C-H Activation for the Preparation of Substituted Ureas,” J. Am. Chem. Soc., 2017, 139, 15407-

15413.

150 Thayer, A.M. “Trace Metals Debate,” Chem. Eng. News, 2013, 91, 10-13.

Page 171: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

155

151 Chakraborty, S.; Lagaditis, P.O.; Förster, M.; Bielinski, E.A.; Hazari, N.; Holthausen, M.C.; Jones, W.D.;

Schneider, S. “Well-Defined Iron Catalysts for the Acceptorless Reversible Dehydrogenation-Hydrogenation

of Alcohols and Ketones,” ACS Catal., 2014, 4, 3994-4003.

152 For a few examples of catalytic dehydrogenative coupling reactions using aniline as a substrate, see: [a]

Ref. 134. [b] Ref. 45. [c] Ref. 47. [d] Ref. 4848.

153 Zhang, M.; Imm, S.; Bähn, S.; Neubert, L.; Neumann, H.; Beller, M. “Efficient copper(II)-catalyzed

transamidation of non-activated primary carboxamides and ureas with amines,” Angew. Chem. Int. Ed., 2012,

51, 3905-3909.

154 Alper, H.; Hartstock, F.W. “An Exceptionally Mild, Catalytic Homogeneous Method for the Conversion

of Amines into Carbamate Esters,” J. Chem. Soc., Chem. Commun., 1985, 1141-1142.

Page 172: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

156

Appendix 1

1.2.A.i.: Nickel-Promoted CO2-Alkene Coupling for the Stereoselective Production of

Substituted Cyclic Anhydrides

General Considerations

All manipulations were carried out under a nitrogen or argon atmosphere using standard

Schlenk, vacuum, cannula, or glovebox techniques. Ligand 9 and NaBAr4F were prepared

according to previously published procedures.1,2 Bulk solvents were dried and

deoxygenated using literature procedures.3 All chemicals were purchased from VWR,

Aldrich, Fisher Scientific, Strem, or Cambridge Isotope Laboratories. NMR solvents were

dried over 3 Å molecular sieves and then used without further manipulation, or sodium and

then vacuum transferred prior to use. 1H, 13C, and 31P NMR spectra were recorded on

Bruker DRX 400 or Avance 600 MHz spectrometers at ambient temperature, unless

otherwise noted. Chemical shifts are reported in ppm; J values are given in Hz. In some

cases, the inequivalent 31P nuclei have coupling constants too small to be observed due to

the combination of 2J and 3J couplings and are therefore denoted as singlets. 1H and 13C

chemical shifts are referenced to residual solvent signals; 31P NMR chemical shifts are

referenced to an external standard of H3PO4. Probe temperatures were calibrated using

ethylene glycol and methanol as previously described.4 GCMS analyses were performed

using an Agilent gas chromatogram with mass spectrometer/Agilent CP-chiraSil-DEX

chiral gas chromatography column.

Page 173: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

157

Figure A1.1: Structure key for ligand numbering. iPr = isopropyl. tBu = tert-butyl.

1 2

3 4 5

6 7

8 9

Page 174: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

158

Ligand Number Name

1 (R)-2,2’-bis[bis(4-methoxy-3,5-di-tert-

butyl-phenyl)phosphino]-4,4’,6,6’-

tetramethoxybiphenyl =

(R)-DTBM-GARPHOS

2 1,2-bis((2S,5S)-2,5-

dimethylphospholano)ethane =

(S,S)-Me-BPE

3 1,2-bis((2S,5S)-2,5-

diisopropylphospholano)ethane =

(S,S)-iPr-BPE

4 (-)-1,2-bis((2S,5S)-2,5-

diisopropylphospholano)benzene =

(S,S)-iPr-DUPHOS

5 (1R,1’R,2S,2’S)-2,2’-di-tert-butyl-

2,3,2’,3’-tetrahydro-1H,1’H-

(1,1’)biisophosphindolyl =

(1R,1’R,2S,2’S)-DuanPhos

6 (S,S)-1,2-bis(tert-

butylmethylphosphino)ethane =

(S,S)-tBu-BISP*

7 (R)-1-[(Sp)-2-

(dicyclohexylphosphino)ferrocenyl]

ethyldi-tert-butylphosphine =

(R)-Fc(PCy2)(PtBu2)

8 (R)-1-[(Sp)-2-

(dicyclohexylphosphino)ferrocenyl]

ethyldicyclohexylphosphine =

(R)-Fc(PCy2)2

9 1,1’-bis((2S,5S)-2,5-

diisopropylphospholano)ferrocene =

(S,S)-Fc-iPr-BPE

Synthesis of Ligand-Ni(COD) Species and Select NMR Data

Scheme A1.1: Synthesis of Ligand-Ni(COD) species.

Table A1.1: Naming Key for Ligand Numbering.

Page 175: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

159

Synthesis of 1-Ni(COD):

A J. Young NMR tube was treated with 1 (0.010 g, 0.008 mmol), Ni(COD)2 (0.003 g, 0.011

mmol), and C6D6 (0.5 mL). The solution was heated overnight at 60 ̊ C and then filtered

through Celite. 1H NMR (400 MHz, C6D6): 1.44 (d, J = 14 Hz), 2.90 (s), 3.41 (t, J = 8),

6.71 (s), 7.56 (s), 7.73 (s). 31P{1H} NMR (400 MHz, C6D6): 36.0 (s).

Synthesis of 2-Ni(COD):

A 20 mL scintillation vial was charged with Ni(COD)2 (0.020 g, 0.073 mmol) and C6D6,

with stirring. The solution was transferred to a J. Young NMR tube and then 1 equivalent

of 2 (20 μL, 0.073 mmol) was injected using a microsyringe. 1H NMR (400 MHz, C6D6):

0.58 (m), 0.80 (m), 1.31 (m), 1.50 (m), 1.73 (m), 2.40 (m), 2.72 (m), 4.04 (m, 2-Ni(COD)),

5.06 (m, 2-Ni(COD)). 31P{1H} NMR (300 MHz, C6D6): 68.7 (s).

Synthesis of 3-Ni(COD):

A 20 mL scintillation vial was charged with Ni(COD)2 (0.051 g, 0.19 mmol) and THF (3

mL). A separate scintillation vial was charged with 3 (0.069 g, 0.19 mmol) and THF (3

mL). The ligand solution was added dropwise with stirring to the Ni(COD)2 solution, then

stirred overnight at room temperature. The THF was removed under vacuum and the

resulting oily yellow solid was dissolved in pentane and cooled to -35 ̊ C. The solution was

filtered through Celite and the pentane was removed under vacuum to yield a yellow solid.

1H NMR (400 MHz, C6D6): 0.74 (d, J = 6 Hz, CH(CH3)2), 0.84 (m), 0.96 (m), 1.02 (m),

1.26 (m), 2.39 (m), 2.73 (m), 4.13 (m, 3-Ni(COD)), 4.92 (m, 3-Ni(COD)). 31P{1H} NMR

(400 MHz, C6D6): 58.6 (s).

Page 176: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

160

Synthesis of 4-Ni(COD):

A 20 mL scintillation vial was charged with Ni(COD)2 (0.033 g, 0.12 mmol) and THF (3

mL). A separate scintillation vial was charged with 4 (0.050 g, 0.12 mmol) and THF (3

mL). The ligand solution was added dropwise, with stirring, to the Ni(COD)2 solution, then

transferred to a 50 mL round bottom flask fitted with a needle valve. The solution was

heated at 50 ̊ C overnight and then the solvent was removed under vacuum. The resulting

orange oil was dissolved in pentane, cooled to -35 ̊ C, and filtered through Celite. The

pentane was removed under vacuum to give an oily orange solid. 1H NMR (400 MHz,

C6D6): 0.52 (s), 0.65 (s), 0.83 (s), 1,10 (m), 1.22 (m), 1.36 (m), 1.50 (m), 1.62 (m), 2.43

(m), 2.74 (s), 5.06 (s), 7.56 (s). 31P{1H} NMR (400 MHz, C6D6): 56.3 (s).

Synthesis of 5-Ni(COD):

A 20 mL scintillation vial was charged with Ni(COD)2 (0.028 g, 0.10 mmol) and THF (3

mL). A separate scintillation vial was charged with 5 (0.039 g, 0.10 mmol) and THF (3

mL). The ligand solution was added dropwise with stirring to the Ni(COD)2 solution, then

stirred overnight at room temperature. The THF was removed under vacuum and the

resulting oily yellow solid was dissolved in pentane and cooled to -35 ̊ C. The solution was

filtered through Celite and the pentane was removed under vacuum to yield a yellow solid.

1H NMR (400 MHz, C6D6): 0.82 (d, J = 9 Hz, C(CH3)4), 2.27 (s), 2.80 (s, 5-Ni(COD)),

3.08 (s), 3.95 (m), 4.57 (s, 5-Ni(COD)), 7.05 (s, 4H, aryl), 7.23 (s, 4H, aryl). 31P{1H} NMR

(400 MHz, C6D6): 95.1 (s).

Page 177: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

161

Synthesis of 6-Ni(COD)

A 20 mL scintillation vial was charged with Ni(COD)2 (0.089 g, 0.32 mmol) and THF (3

mL). A separate scintillation vial was charged with 6 (0.076 g, 0.32 mmol) and THF (3

mL). The ligand solution was added dropwise with stirring to the Ni(COD)2 solution, then

stirred for four hours at room temperature. The THF was removed under vacuum and the

resulting solid was extracted with cold diethyl ether. The extractions were collected and

the ether was removed under vacuum to yield a brown oil. 1H NMR (400 MHz, C6D6): 0.94

(d, J = 5), 1.04 (d, J = 12), 1.20 (m), 1.39 (m), 2.17 (m), 2.42 (m), 2.68 (m), 4.07 (m), 4.20

(m), 4.48 (m), 4.55 (m). 31P{1H} NMR (400 MHz, C6D6): 51.9 (minor, s), 53.1 (major, s).

Synthesis of 7-Ni(COD)

Due to lack of reaction at room temperature and poor conversion followed by

decomposition at elevated temperatures (60-80 °C), this species was not isolated and was

instead formed in situ during coupling reactions. 31P{1H} NMR (400 MHz, THF/C6D6):

21.8 (d, J = 27 Hz, 1P), 80.2 (d, J = 27 Hz, 1P).

Synthesis of 8-Ni(COD)

A 20 mL scintillation vial was charged with Ni(COD)2 (0.011 g, 0.04 mmol) and THF (3

mL). A separate scintillation vial was charged with 8 (0.025 g, 0.04 mmol) and THF (3

mL). The ligand solution was added dropwise with stirring to the Ni(COD)2 solution, then

stirred overnight at room temperature. The THF was removed under vacuum and the

resulting oily solid was washed with pentane and re-dried to yield a red-orange powder. 1H

NMR (400 MHz, C6D6): 1.03 (m), 1.28 (m), 1.52 (m), 1.64 (m), 1.75 (m), 2.54 (m), 2.88

(m), 4.02 (m), 4.09 (s), 4.10 (d, J = 6), 4.43 (s), 4.62 (s), 5.08 (br s), 5.17 (br s). 31P{1H}

Page 178: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

162

NMR (400 MHz, C6D6): 25.1 (minor a, d, J = 28, 1P), 25.7 (minor b, d, J = 29, 1P), 26.1

(major, d, J = 24, 1P), 48.0 (minor b, d, J = 29, 1P), 49.0 (major, d, J = 24, 1P), 49.1 (minor

a, d, J = 28, 1P).

Synthesis of 9-Ni(COD)

A 20 mL scintillation vial was charged with Ni(COD)2 (0.012 g, 0.04 mmol) and THF (3

mL). A separate scintillation vial was charged with 9 (0.023 g, 0.04 mmol) and THF (3

mL). The ligand solution was added dropwise with stirring to the Ni(COD)2 solution, then

stirred overnight at room temperature. The THF was removed under vacuum yielding a

yellow powder. 1H NMR (400 MHz, C6D6): 0.75 (s), 0.99 (s), 1.05 (s), 1.30 (s), 1.35 (s),

1.41 (s), 1.57 (s), 1.68 (s), 1.79 (s), 2.01 (s), 2.28 (s), 2.40 (s), 2.51 (s), 3.07 (br s), 3.88 (s),

4.03 (s), 4.18 (s), 4.56 (s), 4.60 (s), 4.71 (s). 31P{1H} NMR (400 MHz, C6D6): 26.2 (major,

d, J = 32, 1P), 26.9 (minor, d, J = 34, 1P), 28.4 (s, overlapping with minor d), 29.2 (major,

d, J = 32, 1P).

Synthesis of Ligand-Ni-ethylene and Ligand-Ni-lactone Species

*Note: the table after the general methods sections gives the optimized reaction conditions

for each specific ligand

Scheme A1.2: Synthesis of Ligand-Ni-lactone species.

Page 179: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

163

General Method: Ligands 1-6 and 8-9

Dissolve 5-10 mg of Ligand-Ni(COD) and 1 equiv. NaBAr4F (if needed) in C6D6, acetone-

d6, or THF-d8 and transfer to a J. Young NMR tube (0-4 atm intended total pressure) or a

high-pressure NMR tube (5-13 atm intended total pressure). Sequentially add C2H4, then

CO2 or 13CO2. Heat up to 80 °C as necessary for 1-5 days.

Adaptation for ligand 7: Instead of the Ligand-Ni(COD) species, combine 5 mg Ni(COD)2

and 10 mg ligand with 1 equiv. NaBAr4F in the chosen NMR solvent.

Synthesis of Ligand-Ni-propylene and Ligand-Ni-lactone’ Species

General Method:

Dissolve 5-10 mg of Ligand-Ni(COD) and 1 equiv. NaBAr4F (if needed) in C6D6, acetone-

d6, or THF-d8 and transfer to a J. Young NMR tube (0-4 atm intended total pressure) or a

high-pressure NMR tube (5-13 atm intended total pressure). Sequentially add C3H6, then

CO2 or 13CO2. Heat up to 80 °C as necessary for 1-5 days.

Scheme A1.3: Synthesis of Ligand-Ni-lactone’ species.

Page 180: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

164

Ligand LNi(COD) NaBAr4F C2H4 CO2 Solvent Temp. Time

1 - - - - - - -

2 5 mg - 11 atm - C6D6 22 °C 2 hrs.

3 10 mg 18 mg 0.1 atm 0.1 atm Acetone-d6 22 °C 12 hrs.

4 10 mg 15 mg 2 atm 2 atm THF-d8 50 °C 12 hrs.

5 7 mg 11 mg 1.5 atm 1.5 atm C6D6 22 °C 12 hrs.

6 8 mg - 5 atm 5 atm C6D6 22 °C 48 hrs.

7 5 mg

Ni(COD)2

10 mg

ligand

16 mg 2 atm 2 atm THF-d8 22 °C 48 hrs.

8 10 mg 12 mg 5 atm 5 atm THF-d8 22 °C 48 hrs.

9 - - - - - - -

Ligand LNi(COD) NaBAr4F C3H6 CO2 Solvent Temp. Time

1 - - - - - - -

2 - - - - - - -

3 15 mg - 5 atm 5 atm C6D6 22 °C 48 hrs.

4 5 mg 8 mg 8 atm 2 atm THF-d8 22 °C 48 hrs.

5 10 mg 17 mg 5 atm 5 atm THF-d8 22 °C 48 hrs.

6 10 mg 22 mg 5 atm 5 atm THF-d8 22 °C 72 hrs.

7 - - - - - - -

8 - - - - - - -

9 12 mg 18 mg 5 atm 5 atm THF-d8 50 °C 12 hrs.

Ligand Ni-ethylene

(ppm)

Ni-lactone

(ppm)

Ni-propylene

(ppm)

Ni-lactone’

(ppm)

1 - - - -

2

(C6D6)

67.5

(s, 2P)

(121 MHz)

- - -

3

(acetone-d6)

55.8

(s, 2P)

(121 MHz)

58.1

(s, 1P)

64.8

(s, 1P)

(300 MHz)

53.7

(d, J = 69, 1P)

55.4

(d, J = 70, 1P)

major 49.1

56.5

(d, J = 13, 1P),

58.8

(d, J = 13, 1P)

60.1

Table A1.2: Optimized Experimental Conditions for Ni-lactone Syntheses

Table A1.3: Optimized Experimental Conditions for Ni-lactone’ Syntheses

Table A1.4: 31P{1H} NMR Shifts for Synthesized Ni Species. J Values are in Hz.

Page 181: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

165

(d, J = 66, 1P)

54.8

(d, J = 67, 1P)

minor (162 MHz)

(d, J = 12, 1P),

64.8

(d, J = 12, 1P)

60.9

(d, J = 3, 1P)

62.1

(d, J = 3, 1P)

(162 MHz)

4

(THF-d8)

60.2

(s, 2P)

(243 MHz)

59.0

(s, 1P)

63.4

(s, 1P)

(243 MHz)

54.8

(d, J = 69, 1P),

56.6

(d, J = 69, 1P)

(162 Mz)

-

5

(C6D6, C2H4)

(THF-d8, C3H6)

97.4

(s, 2P)

(162 MHz)

91.3 (d, J = 10,

1P)

105.7 (d, J = 9,

1P)

(162 MHz)

Possible: 94.7

(s, 2P)

(162 MHz)

-

6

(C6D6, C2H4)

(THF-d8, C3H6)

53.8

(s, 2P)

(162 MHz)

50.8 (s, 1P)

59.1 (s, 1P)

(162 MHz)

60.6 (s, 1P)

62.8 (s, 1P)

(162 MHz)

51.1 (d, J = 13,

1P))

64.3 (s, 1P)

(162 MHz)

7

(THF-d8)

26.6 (d, J = 24,

1P)

82.9 (d, J = 25,

1P)

(162 MHz)

- - -

8

(THF-d8)

28.0 (d, J = 21,

1P)

52.1 (d, J = 22,

1P)

(162 MHz)

29.8 (d, J = 24,

1P)

54.0 (d, J = 24,

1P)

(162 MHz)

- -

9

(THF-d8)

- - 31.7 (d, J = 29,

1P),

32.6 (d, J = 29,

1P)

major

27.6 (d, J = 29,

1P),

32.1 (d, J = 28,

1P)

minor

(162 MHz)

-

Page 182: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

166

Release of Cyclic Anhydride Product from Ni-lactone and Ni-lactone’ Species

General Method:

Take the Ligand-Ni-lactone and Ligand-Ni-lactone’ solutions and treat with 1 atm of CO

gas in a J. Young or high-pressure NMR tube. Allow to react for approximately 20 minutes

at room temperature. Analyze for SA or MSA using 1H and 13C NMR spectroscopy and

GCMS analysis.

Ligand Ligand-Ni(CO)2 (ppm) Solvent/Spectrometer

1 - -

2 - -

3 55.5 acetone-d6/400 MHz

4 58.9 acetone-d6/400 MHz

5 96.2 acetone-d6/400 MHz

6 53.2 acetone-d6/400 MHz

7 22.7

(d, J = 8)

88.3

(d, J = 8)

THF-d8/400 MHz

8 29.8

(d, J = 11)

52.7

(d, J = 11)

acetone-d6/400 MHz

9 28.6

41.4

THF-d8/600 MHz

Scheme A1.4: Release of cyclic anhydride product from lactone using CO gas.

Table A1.5: 31P{1H} NMR Shifts for Ligand-Ni(CO)2 Species. J Values are in Hz.

Page 183: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

167

Peaks/GC Information Used in the Identification of Succinic Anhydride and

Methylsuccinic Anhydride:

Succinic anhydride: 1H NMR (400 MHz, acetone-d6): 3.05 (s).

Methylsuccinic anhydride: 13C NMR (400 MHz, C6D6): 169.82 (C=O), 174.34 (C=O).

GCMS in acetone-d6: peaks at 16.577 and 16.866 minutes on chiral column (initial temp.:

40 ̊ C, ramp rate: 5 ̊ C/min, max. temp.: 150 ̊ C); m/z: 15, 28, 42, 55; (m+1)/z: 71.

Page 184: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

168

Select NMR Data: L-Ni(COD) Reactions with C2H4 or C3H6, CO2, and CO

*See Tables A1.2 and A1.3 for experimental conditions; see Schemes A1.2-A1.4 for

generic reaction depictions

Figure A1.2: 31P{1H} NMR spectrum in C6D6 showing limited conversion to 1-Ni(COD). 300 MHz

spectrometer.

Figure A1.3: 31P{1H} NMR spectrum in C6D6 showing limited conversion to 2-Ni-ethylene under 11

atm of C2H4. 300 MHz spectrometer.

2-Ni(COD)

2-Ni-2 2-Ni-ethylene

1-Ni(COD)

Ligand

Impurity

Free

Ligand

Page 185: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

169

Figure A1.4: 31P{1H} NMR spectra in acetone-d6 showing conversion from the 3-Ni(COD) species (A,

blue) to 3-Ni-ethylene (B, red), 3-Ni-lactone (C, green), and 3-Ni(CO)2 (D, purple). Blue = 600 MHz

spectrometer, others = 300 MHz.

3-Ni(COD) + NaBAr4F

+ C2H4

+ CO2

+ CO D

B

A

C C

Figure A1.5: 31P{1H} NMR spectra in acetone-d6 showing conversion from the 3-Ni(COD) species (A,

blue) to 3-Ni-propylene (B/B’, red), 3-Ni-lactone’ (C/C’/C”, green), and 3-Ni(CO)2 (D, purple). 400

MHz spectrometer.

A

D

B B’ B’

B

C” C” C’ C’ C C

3-Ni(COD) + NaBAr4F

+ C3H6

+ CO2

+ CO

Page 186: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

170

Figure A1.6: 31P{1H} NMR spectra in THF-d8 showing conversion from the 4-Ni(COD) species (A,

blue) to 4-Ni-ethylene (B, red), 4-Ni-lactone (C, green), and 4-Ni(CO)2 (D, purple). 600 MHz

spectrometer. * = unidentified side product.

A

B

C C

D

* *

4-Ni(COD) + NaBAr4F

+ C2H4

+ CO2

+ CO

Figure A1.7: 31P{1H} NMR spectra in THF-d8 showing conversion from the 4-Ni(COD) species (A,

blue) to 4-Ni-propylene (B, blue), and 4-Ni-CO2 adduct (C, red). 400 MHz spectrometer. * = unidentified

side product. Corresponding 4-Ni(CO)2 complex is the same as (D) in the purple spectrum in Figure

A1.6.

4-Ni(COD) + NaBAr4F + C3H6

*

A B

C

+ CO2

C

B

Page 187: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

171

Figure A1.8: 31P{1H} NMR spectra in C6D6 showing conversion from the 5-Ni(COD) species (A, blue)

to 5-Ni-ethylene (B, red), 5-Ni-lactone (C, green), and 5-Ni-(CO)2 (D, purple). 400 MHz spectrometer.

Purple spectrum in acetone-d6. NaBAr4F added after CO2 to enhance conversion.

5-Ni(COD)

+ C2H4

+ CO2 + NaBAr4F

+ CO

C C

B

A

5-Ni-5

Figure A1.9: 31P{1H} NMR spectra in THF-d8 showing conversion from the 5-Ni(COD) species (A,

blue) to potential 5-Ni-propylene (B, red), 5-Ni-CO2 adduct (C, green), and 5-Ni-(CO)2 (D, purple). 400

MHz spectrometer.

5-Ni(COD) + NaBAr4F

+ C3H6

+ 13CO2

+ CO

A

B

D

C C

D

Page 188: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

172

Figure A1.10: 31P{1H} NMR spectra in C6D6 showing conversion from the 6-Ni(COD) species (A, blue,

two diastereomers) to 6-Ni-ethylene (B, red), 6-Ni-lactone (C, green), and 6-Ni-(CO)2 (D, purple). 400

MHz spectrometer. Purple spectrum in acetone-d6. * = unidentified decomposition product.

6-Ni(COD)

+ C2H4

+ CO2

+ CO

A

A

B

C C

D

Figure A1.11: 31P{1H} NMR spectra in THF-d8 showing conversion from the 6-Ni(COD) species (A,

blue, two diastereomers) to 6-Ni-propylene (B, red), 6-Ni-lactone’ (C, red), and 6-Ni-(CO)2 (D, green).

400 MHz spectrometer. Green spectrum in acetone-d6.

6-Ni(COD) + NaBAr4F A

A

C

C

B B

D

+ C3H6 + 13CO2

+ CO

*

Page 189: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

173

Figure A1.12: 31P{1H} NMR spectra in THF-d8 showing conversion from the 7-Ni(COD) species (A,

blue) to 7-Ni-ethylene (B, red), no change upon CO2 addition (green), then to 7-Ni(CO)2 (D, purple). 400

MHz spectrometer. Blue spectrum in a mixture of THF/C6D6.

7-Ni(COD)

+ NaBAr4F + C2H4

+ CO2

+ CO

A A

B B

B

D D

B

Figure A1.13: 31P{1H} NMR spectra in THF-d8 showing conversion from the 8-Ni(COD) species (A,

blue) to 8-Ni-ethylene (B, red), 8-Ni-lactone (green), and 8-Ni(CO)2 (purple). 400 MHz spectrometer.

Purple spectrum in acetone-d6.

8-Ni(COD) + NaBAr4F

+ C2H4

+ CO2

+ CO

A A

B B

C C

D D

Page 190: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

174

Figure A1.14: 31P{1H} NMR spectra in THF-d8 showing conversion from the 9-Ni(COD) species (A,

blue) to 9-Ni-propylene (B/B’, red), and 9-Ni(CO)2 (green). Green = 600 MHz spectrometer, others =

400 MHz. * = unidentified impurity.

9-Ni(COD) + NaBAr4F

+ C3H6 + CO2

+ CO

D D

B’ B’

B B

A A

*

Page 191: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

175

Retrosynthesis: Ligand-Ni(COD) + Succinic Anhydride or Methylsuccinic Anhydride

Retrosyntheses of Ligand-Ni-lactone or Ligand-Ni–lactone’ species using SA or MSA

followed the same general procedure:

A J. Young NMR tube was charged with Ligand-Ni(COD) (0.010 g), succinic anhydride

or methylsuccinic anhydride (1.5 equiv.), and C6D6 or acetone-d6 (0.5 mL). The NMR tube

was heated at either 60 ̊ C or 70 ̊ C for 12 hours.

Retrosynthesis: Ligand-Ni(COD) + Acrylic Acid:

Retrosyntheses of Ligand-Nilactone species using acrylic acid followed the same general

procedure:

A J. Young NMR tube was charged with Ligand-Ni(COD) (0.010 g), acrylic acid (1 equiv.

or excess), and C6D6 or acetone-d6 (0.5 mL). The reaction solution was monitored

periodically at room temperature for up to 10 days.

Page 192: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

176

Select NMR Data: Reactions of 3-, 4-, and 5-Ni(COD) without Lewis Acid;

Retrosynthetic Reactions

Figure A1.15: 31P{1H} NMR spectra of the reaction of 3-Ni(COD) with C2H4 and CO2 in the absence

of Na salt. B = 3-Ni-ethylene, C = 3-Ni-lactone. C6D6, 300 MHz spectrometer.

3-Ni(COD) + C2H4

+ CO2

B

B

C C

Figure A1.16: 31P{1H} NMR spectrum of the reaction of 4-Ni(COD) with C2H4 and CO2 in the absence

of Na salt. A = 4-Ni(COD), B = 4-Ni-ethylene, C = 6-Ni-lactone. Acetone-d6, 300 MHz spectrometer.

B C C

A

4-Ni(COD) + C2H4 + CO2

Page 193: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

177

Figure A1.17: 31P{1H} NMR spectrum of the reaction of 5-Ni(COD) with C2H4 in the absence of Na

salt. A = 5-Ni(COD), B = 5-Ni-ethylene. C6D6, 400 MHz spectrometer.

A B

5-Ni-5

5-Ni(COD) + C2H4

Figure A1.18: 31P{1H} NMR spectra of the reaction of 3-Ni(COD) with C2H4 and CO2 (blue) or SA

(red) to generate a nickelalactone. C = 3-Ni-lactone (γ), C’ = suspected β-lactone, D = 3-Ni(CO)2.

Acetone-d6, 400 MHz spectrometer. Slight shift changes attributed to presence of Na salt.

3-Ni(COD) + NaBAr4F + C2H4 + CO2

3-Ni(COD) + SA

C C

C’ C’

C C

D

Page 194: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

178

Figure A1.19: 31P{1H} NMR spectra of the reaction of 4-Ni(COD) with C2H4 and CO2 (blue) or SA

(red) to generate a nickelalactone. A = 4-Ni(COD), B = 4-Ni-ethylene, C = 4-Ni-lactone, D = 4-Ni(CO)2.

Acetone-d6, 300 MHz spectrometer. * = unidentified impurity.

A

A

B

D

C C

C C *

4-Ni(COD) + C2H4 + CO2

4-Ni(COD) + SA

Figure A1.20: 31P{1H} NMR spectra of the reaction of 5-Ni(COD) with C2H4 and CO2 (blue) or SA

(red) to generate a nickelalactone. A = 5-Ni(COD), B = 5-Ni-ethylene, C = 5-Ni-lactone, D = 5-Ni(CO)2.

C6D6, 400 MHz spectrometer.

5-Ni(COD) + C2H4 + CO2

5-Ni(COD) + SA

C C

C C

D

A

B

Page 195: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

179

Figure A1.21: 31P{1H} NMR spectrum of the reaction of 3-Ni(COD) with acrylic acid to yield multiple

unidentified products inconsistent with lactone. Acetone-d6, 300 MHz spectrometer.

Figure A1.22: 31P{1H} NMR spectra of the reaction of 3-Ni(COD) with C3H6 and 13CO2 (blue) or MSA

(red, no Na salt; green with Na salt) to generate a nickelalactone. A = 3-Ni(COD), B = 3-Ni-propylene,

C/C’/C” = 3-Ni-lactone’, D = 3-Ni(CO)2. Coupling reaction in C6D6, MSA reactions in acetone-d6; 400

MHz spectrometer. Experiment not repeated with other Ligand-Ni(COD) species because of lack of

peak agreement between methods.

A

D

B B C

C’ C’ C” C”

3-Ni(COD) + NaBAr4F + C3H6 + CO2

3-Ni(COD) + NaBAr4F + MSA

3-Ni(COD) + MSA

C

A D

Page 196: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

180

1.2.A.ii.: The Use of CO Surrogates in Nickel-Promoted CO2-Alkene Coupling for the

Catalytic Production of Cyclic Anhydrides

General Considerations

All manipulations were carried out under a nitrogen or argon atmosphere using standard

Schlenk, vacuum, cannula, or glovebox techniques. The following species were prepared

by adaptations of previously reported procedures: dcpe-Ni(COD),5,6 dcpe-Ni-lactone’,5,6

and dcpe-Ni(CO2).7,8 Bulk solvents were dried and deoxygenated using literature

procedures.3 All chemicals were purchased from VWR, Aldrich, Fisher Scientific, Strem,

or Cambridge Isotope Laboratories. NMR solvents were dried over 3 Å molecular sieves

and then used without further manipulation, or sodium and then vacuum transferred prior

to use. 1H, 13C, and 31P NMR spectra were recorded on Bruker 300 MHz Avance II+, 300

MHz DRX, 500 MHz DRX or 600 MHz spectrometers at ambient temperature, unless

otherwise noted. Chemical shifts are reported in ppm; J values are given in Hz. In some

cases, the inequivalent 31P nuclei have coupling constants too small to be observed due to

the combination of 2J and 3J couplings and are therefore denoted as singlets. 1H and 13C

chemical shifts are referenced to residual solvent signals; 31P NMR chemical shifts are

referenced to an external standard of H3PO4. Probe temperatures were calibrated using

ethylene glycol and methanol as previously described.4 GCMS analyses were performed

using an Agilent gas chromatogram with mass spectrometer.

Synthesis of dcpe-Ni(COD):

In a 20-mL scintillation vial, dissolve 0.020 g Ni(COD)2 in ~2 mL THF. In a separate vial,

dissolve 1 equiv. dcpe ligand (0.030 g) in ~4 mL THF. Add the ligand solution to the nickel

solution dropwise, with stirring. Stir at room temperature for ~4 hours, then remove the

Page 197: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

181

solvent under vacuum. Recrystallize the resulting oily solid from pentane to yield a

powdery yellow solid. 1H NMR (C6D6, 400 MHz): 1.21-1.33 (m), 1.66 (s), 1.71 (s), 1.73

(s), 1.82-1.85 (m), 2.52 (m), 2.68 (m), 4.46 (s). 31P{1H} NMR (C6D6, 400 MHz): 59.0 (s,

2P).

Synthesis of dcpe-Ni-lactone’:

Dissolve 0.100 g dcpe-Ni(COD) in ~6 mL THF and transfer to a 25-mL heavy-walled glass

reaction vessel. Add 1.5 atm each propylene, then CO2. Stir at 50 °C overnight. De-gas the

solution and filter through Celite. Remove the solvent under vacuum to give an oily solid.

Wash the solid with cold toluene, then cold pentane (1-2 times) until the pentane wash is

colorless. Dry the resulting orange solid under vacuum. 1H NMR (C6D6, 500 MHz): 0.91-

1.01 (m), 1.10 (s), 1.16-1.40 (m), 1.55 (s), 1.65-1.69 (m), 1.82 (m), 2.02 (m), 2.12 (m), 2.34

(s), 2.52 (s), 2.68 (s), 2.82 (s). 31P{1H} NMR (C6D6, 400 MHz): 61.2 (s, 1P), 68.4 (s, 1P).

Page 198: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

182

Entry Ni C3H6, CO2 Temp./Time

1 7 mg 1.5 atm each 70 °C, 10 min.

2 10 mg 1.5 atm each 70 °C, 3 days

3 6 mg 1.5 atm each 50 °C, 1 hr.

60 °C, overnight

70 °C, overnight

4 7 mg 1.5 atm each 50 °C, overnight

60 °C, 2 days

70 °C, overnight

5 8 mg None 50 °C, overnight

70 °C, overnight

6 10 mg 0.5 atm each 22 °C, 3 hrs.

60 °C, 5 hrs.

70 °C, overnight

7 10 mg None 22 °C, 2-3 hrs.

8 8 mg 1 atm each 60 °C, 1 hr.

9 10 mg 1 atm each 22 °C, 1 hr.

10 10 mg 1 atm each 50 °C, 1 hr.

11 10 mg 1 atm each 50 °C, 5 hrs.

12 10 mg 0.5 atm each 60 °C, 2 hrs.

13 10 mg 0.5 atm each 60 °C, 2 hrs.

14 10 mg 0.5 atm each 22 °C, 2 hrs.

15 10 mg 1 atm each 22 °C, 6 hrs.

50 °C, 2 hrs.

General Procedure for CO Surrogate Reactions:

NMR-scale: Dissolve dcpe-Ni(COD) or dcpe-Ni-lactone’ in ~0.5 mL of the desired NMR

solvent with X equiv. [CO]. Transfer to a J. Young NMR tube (total gas pressure <4 atm)

Table A1.6: Experimental Conditions for Table 1.2.A.1 Entries

Scheme A1.5: Nickel-promoted production of cyclic anhydrides using CO surrogates

Page 199: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

183

or a high-pressure NMR tube (total gas pressure 4-13 atm). Add Y equiv. [B]. Quickly

freeze at -196 °C and add the desired pressure of propylene, then CO2 using a calibrated

gas bulb. Thaw and then heat as needed for the duration of the reaction.

Reaction-scale: Dissolve dcpe-Ni(COD) or dcpe-Ni-lactone’ in 2-6 mL’s C6H6 in a 20-

mL scintillation vial. Add X equiv. [CO], then Y equiv. [B]. If no gases pressure or heat is

needed, stir in the vial at room temperature for the duration of the reaction. Otherwise,

transfer the reactant mixture to a 25-mL heavy-walled glass reaction vessel, quickly freeze

at -196 °C and add the desired pressure of propylene, then CO2 using a calibrated gas bulb.

Thaw and then heat as needed for the duration of the reaction.

Optimization Tables for CO Surrogate Reactions:

Entry Ni Species [CO] [CO]

equiv.

[B] [B]

equiv.

Solvent MSA

(%)

1 A NFS 3 NEt3 4.5 THF-d8 0

2 A NFS 3 NEt3 4.5 C6D6 0

3b A NFS 3 - - C6D6 0

4c B NFS 3 NEt3 4 C6D6 6.30

Table A1.7: Determination of Starting Ni Speciesa

aReaction conditions: 8-17 mg Ni species, 3 equiv. NFS, and 0-4.5 equiv. NEt3 in the listed solvent with

1-1.5 atm each C3H6 and CO2 for 2 hrs.-3 days at room temperature. bNo gas added. c6 hrs. R.T., 1 hr. at

50 °C. MSA = methylsuccinic anhydride. Yield determined by GCMS using naphthalene as the standard.

Page 200: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

184

Entry Ni Species [CO] [CO]

equiv.

[B] [B]

equiv.

Solvent MSA

(%)

1 B PF 3 NEt3 4 30:70

C6D6:

THF-d8

0.01

2 B PF 3 NEt3 4 50:50

C6D6:

THF-d8

0.03

3 B PF 3 NEt3 4 C6D6 0.05

4 B PF 9 NEt3 12 50:50

C6D6:

THF-d8

0.68

5 B PF 9 NEt3 24 50:50

C6D6:

THF-d8

1.60

6 B PF 9 NEt3 24 C6H6 2.00

Entry Ni

Species

[CO] [CO]

equiv.

[B] [B]

equiv.

Solvent Gas

Pressure

MSA

(%)

1 B PF 3 NEt3 4 50:50

C6D6:

THF-d8

1.5 atm

each

0.03

2 B PF 3 NEt3 4 50:50

C6D6:

THF-d8

5 atm

each

0.02

3 B PF 9 NEt3 12 50:50

C6D6:

THF-d8

1.5 atm

each

0.68

4 B PF 9 NEt3 12 50:50

C6D6:

THF-d8

None 0.02

Table A1.9: Effect of Gas in Headspacea

aReaction conditions: 5-21 mg Ni species, X equiv. PF, and Y equiv. NEt3 in the listed solvent with C3H6

and CO2 at 50-70 °C for up to 2 days. MSA = methylsuccinic anhydride. Yield determined by GCMS

using naphthalene as the standard.

Table A1.8: Solvent Screena

aReaction conditions: 6-21 mg Ni species, X equiv. [CO], and Y equiv. [B] in the listed solvent with 0.5-

1.5 atm each C3H6 and CO2 at 50-70 °C for up to 3 days. MSA = methylsuccinic anhydride. Yield

determined by GCMS using naphthalene as the standard.

Table A1.9: Effect of Gas in Headspacea

Page 201: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

185

Entry Ni Species [CO] [CO]

equiv.

[B] [B]

equiv.

MSA

(%)

1 B TCPF 4 NEt3 4 4.80

2 B TCPF 4 NEt3 6 4.80

3 B TCPF 4 NEt3 8 5.20

4 B TCPF 4 NEt3 12 3.80

5 B TCPF 4 NEt3 16 7.30

6 B TCPF 4 NEt3 20 7.66

7 B TCPF 4 NEt3 40 7.47

8 B TCPF 4 NEt3 60 7.30

9 B TCPF 4 NEt3 100 4.60

10 B TCPF 4 NEt3 200 5.70

Entry Ni

Species

[CO] [CO]

equiv.

[B] [B]

equiv.

Gas Pressure MSA

(%)

1 B PF 3 NEt3 4 1.5 atm each 0.05

2 B PF 9 NEt3 12 None 0.09

3 B PF 9 NEt3 24 0.5 atm each 2.00

4 B TCPF 3 NEt3 4 1 atm each 5.60

5 B TCPF 4 NEt3 20 None 9.00

6 B TCPF 9 NEt3 24 0.5 atm each 5.40

7 B TCPF 9 NEt3 45 None 6.90

8 B TCPF 9 NEt3 100 0.5 atm each 4.00

9 B NFS 3 NEt3 4 0.5 atm each 6.30

10 B NFS 4 NEt3 10 1 atm each 9.00

11 B NFS 9 NEt3 24 0.5 atm each 7.40

Table A1.11: CO Surrogate Screena

aReaction conditions: 10 mg Ni species, X equiv. TCPF, and Y equiv. NEt3 in C6H6, no

gases, at room temperature for 2.5 hrs. MSA = methylsuccinic anhydride. Yield determined

by GCMS using naphthalene as the standard.

aReaction conditions: 8-10 mg Ni species, X equiv. [CO], and Y equiv. NEt3 in C6D6 or C6H6 with

indicated pressures of C3H6 and CO2 at 50-70 °C for up to 3 days. MSA = methylsuccinic anhydride.

Yield determined by GCMS using naphthalene as the standard.

Table A1.10: CO Surrogate:Base Ratio Screena

Page 202: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

186

Entry Ni

Species

[CO] [CO]

equiv.

[B] [B]

equiv.

Gas Pressure MSA

(%)

1 B PF 9 NEt3 24 0.5 atm each 2.00

2 B TCFP 4 NEt3 20 None 9.00

3 B NFS 4 NEt3 10 1 atm each 9.00

4b B PF 9 DBU 12 1 atm each -

5 B TCPF 3 DBU 4 1 atm each 0.57

6 B NFS 3 DBU 4 1 atm each 0.77

7 B PF 9 DMAP 12 0.5 atm each -

8 B TCPF 9 DMAP 12 0.5 atm each 4.00

9b B NFS 9 DMAP 12 0.5 atm each 6.30

10 B TCPF 4 - - 1 atm each 5.20

Table A1.12: Base Screena

aReaction conditions: 8-10 mg Ni species, X equiv. of [CO], and Y equiv. [B] in C6D6 or C6H6 with

indicated pressures of C3H6 and CO2 at 50-70 °C for up to 3 days. bRoom temperature for 1-2 hrs. MSA

= methylsuccinic anhydride. Yield determined by GCMS using naphthalene as the standard.

Page 203: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

187

Select NMR Spectra:

Figure A1.23: 31P{1H} NMR spectra showing attempts at reaction rate control using C6D6/THF-d8

solvent mixtures. 6-10 mg dcpe-Ni-lactone’, 3 equiv. PF, 4 equiv. NEt3. Blue = 12 hrs. at 70 °C, red = 6

hrs. at 50 °C, green = 24 hrs. at 50 °C, purple = 12 hrs. at 60 °C. B = dcpe-Ni-lactone’, * = dcpe-Ni(CO)2.

300 MHz spectrometer.

B B

*

*

*

*

B B

B

B B

B

C6D6

50:50 C6D6:THF-d8

50:50 C6D6:THF-d8

30:70 C6D6:THF-d8

Page 204: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

188

Figure A1.24: 1H NMR spectra in C6D6 showing lack of decomposition of MSA under CO surrogate

reaction conditions over time. 10 mg MSA, 12 mg TCPF, 20 µl NEt3, 1 atm each C3H6 and CO2. * =

MSA peaks. # = NEt3 peaks. R.T. = room temperature. 600 MHz spectrometer.

* * * * #

30 min., R.T.

1.5 hrs., R.T.

1 hr., 60 °C

2 hrs., 60 °C

12 hrs., 60 °C

Page 205: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

189

Figure A1.25: 1H NMR spectra in C6D6 showing lack of reaction of MSA with side products of CO

surrogate reactions. 10 mg MSA, 8 mg (1 equiv.) phenol, overnight at room temperature. * = MSA peaks.

# = phenol peaks. 500 MHz spectrometer. No evidence of reaction by GCMS.

*

* *

*

#

#

#

#

Page 206: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

190

Figure A1.26: 1H NMR spectrum in C6D6 showing lack of reaction of MSA with side products of CO

surrogate reactions. 10 mg MSA, 8 mg (1 equiv.) phenol, 24 µl (2 equiv.) NEt3, overnight at room

temperature. * = MSA peaks. # = phenol peaks. o = NEt3 peaks. 500 MHz spectrometer. No evidence of

reaction by GCMS.

# #

# * *

*

o o

Page 207: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

191

Preparation of dcpe-Ni(CO2):

1. [(Cy3P)2Ni]2(N2)

Make a suspension of 0.160 g Ni(acac)2 (acac = acetylacetonate) in ~5 mL toluene and

transfer to a 100-mL Schlenk flask. Dissolve 0.349 g PCy3 (2 equiv.) in ~5 mL toluene and

add to the Ni solution. Stir briefly, then cool to -35 °C for ~30 minutes. Inject 0.4 mL

Al(CH3)3 (1.2 equiv.) while cold and under a flow of N2, which will cause an immediate

color change to a deep reddish brown. While still under a positive N2 flow, allow the

reaction solution to warm slowly to room temperature with stirring. After ~2 hours, remove

the N2 flow and stir overnight. Remove the solvent under vacuum and recrystallize the

resulting reddish-brown oil from diethyl ether at -35 °C. Filter the resulting mixture and

wash the dark red crystalline product with cold diethyl ether. 31P{1H} NMR (C6D6, 300

MHz): 46.2 (s, 4P).

2. (Cy3P)2Ni(CO2)

Dissolve 0.095 g of the Ni-N2 dimer from (1) in ~2 mL toluene and transfer to a 25-mL

heavy-walled glass reaction vessel, giving a deep purple solution. Add ~1 atm CO2 at -196

°C using a calibrated gas bulb, resulting in an immediate solution color change to reddish

orange. Stir the mixture at room temperature for ~8 hours, which allows an orange

precipitate to accumulate. Filter the resulting mixture to yield an orange solid. 31P{1H}

NMR (C6D6, 300 MHz): 37.0 (s, 2P).

3. dcpe-Ni(CO2)

Page 208: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

192

Add 0.010 g (Cy3P)2Ni(CO2) and 0.06 g dcpe (dry) to a J. Young NMR tube, then vacuum

transfer in ~0.5 mL C6D6 at -196 °C. Add ~2 atm CO2 using a calibrated gas bulb while

the mixture is still frozen. Thaw and mix. *Note: the reaction is immediate and goes to

completion. The product is not stable in solution and over the course of ~12 hours it

partially reverts back to the Ni-N2 dimer and partially decomposes. 31P{1H} NMR (C6D6,

300 MHz): 43.9 (s, 2P).

Figure A1.27: 31P{1H} NMR spectra in C6D6 showing the products of the stepwise synthesis of dcpe-

Ni(CO2). 300 MHz spectrometer. * = free PCy3.

[(Cy3P)2Ni]2(N2)

(Cy3P)2Ni(CO2)

dcpe-Ni(CO2)

*

Page 209: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

193

1.2.B.: Nickel-Promoted Coupling of Isocyanates and Alkenes for the Production of

N-Substituted Acrylamides

General Considerations

All manipulations were carried out under a nitrogen or argon atmosphere using standard

Schlenk, vacuum, cannula, or glovebox techniques. The dcpe-Ni(COD) complex was

prepared according to previously published procedures.5,6 Bulk solvents were dried and

deoxygenated using literature procedures.3 All chemicals were purchased from VWR,

Aldrich, Fisher Scientific, Strem, or Cambridge Isotope Laboratories. NMR solvents were

dried over 3 Å molecular sieves and then used without further manipulation, or sodium and

then vacuum transferred prior to use. 1H, 13C, and 31P NMR spectra were recorded on

Bruker 300 MHz Avance II+, 300 MHz DRX, 500 MHz DRX or 600 MHz spectrometers

at ambient temperature, unless otherwise noted. Chemical shifts are reported in ppm; J

values are given in Hz. In some cases, the inequivalent 31P nuclei have coupling constants

too small to be observed due to the combination of 2J and 3J couplings and are therefore

denoted as singlets. 1H and 13C chemical shifts are referenced to residual solvent signals;

31P NMR chemical shifts are referenced to an external standard of H3PO4. Probe

temperatures were calibrated using ethylene glycol and methanol as previously described.4

GCMS analyses were performed using an Agilent gas chromatogram with mass

spectrometer. Product identities were confirmed by comparison to authentic samples or

previously reported spectra.9-14 All high-pressure coupling reactions were performed using

a Parr 5500 series compact reactor with glass insert.

*For procedure for dcpe-Ni(COD) preparation, see Appendix 1, section 1.2.A.ii.

Page 210: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

194

Nickel-Promoted Coupling Reactions between Ethylene and Isocyanates

General NMR Procedure:

Dissolve 10-15 mg dcpe-Ni(COD) and 0-2 equivalents Lewis acid or base additive in C6D6

or THF-d8 and transfer to a J. Young (total reaction pressure <4 atm) or high-pressure (total

reaction pressure <13 atm) NMR tube. Add 1-2 atm ethylene, then 1-10 equivalents

isocyanate at -196 °C via calibrated gas bulb. Allow the reaction to warm to room

temperature, then heat anywhere from 22-70 °C for up to 2 days. Add ~1 atm CO gas to

release any bound organic products and analyze using NMR spectroscopy and/or GCMS

using a naphthalene standard (add 10 µl of a 0.1M naphthalene solution in the reaction

solvent to the reaction mixture and dilute to a total volume of 0.5 mL).

Adjusted procedure for phenyl isocyanate:

Due to its boiling point, this substrate could not undergo vacuum transfer. Instead, after

preparing the dcpe-Ni(COD) NMR solution, add 1 atm ethylene at -196 °C via calibrated

gas bulb, warm to room temperature, and bring into the glovebox. Freeze the solution in

the coldwell, inject 1-10 equivalents phenyl isocyanate using a microsyringe, and quickly

remove the NMR tube from the glovebox while keeping the solution frozen. Re-add 1-2

Scheme A1.6: Nickel-promoted ethylene/isocyanate coupling pathway.

Page 211: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

195

atm ethylene, then warm the solution to room temperature. Proceed as indicated for other

isocyanates.

General Parr Procedure:

Dissolve 10-15 mg dcpe-Ni(COD) and 0-20 equivalents base additive in C6H6 or THF and

transfer to a Parr base. In the case of phenyl isocyanate, add 2-10 equivalents using a

microsyringe and seal the Parr reactor. For ethyl and tert-butyl isocyanate, vacuum transfer

2-10 equivalents into an ampule at -196 °C and then flame-seal the ampule. Carefully add

the sealed ampule to the dcpe-Ni(COD) solution in the Parr base in the glovebox, and then

seal the reactor. Outside the glovebox, pressurize the reactor to 300-1000 psi ethylene and

stir at 60 °C for 16 hours. Cool the reactor in an ice bath for ~45 minutes, then partially

vent the ethylene pressure to ~100 psi. Bring the Parr reactor back into the glovebox and

filter the product solution through Celite. Analyze by NMR spectroscopy and/or GCMS.

Page 212: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

196

Ni Species EtN=C=O tBuN=C=O PhN=C=O

A 60.0 ppm (s) 60.0 ppm (s) 60.0 ppm (s)

B 63.3 ppm (s) 63.3 ppm (s) 63.3 ppm (s)

C 53.1 ppm

(d, J = 10)

74.6 ppm

(d, J = 10)

- -

D 48.5 ppm

(d, J = 8)

71.6 ppm

(d, J = 8)

- -

E 60.5 ppm

(d, J = 48)

70.9 ppm

(d, J = 48)

59.9 ppm

(d, J = 55)

68.2 ppm

(d, J = 55)

51.9 ppm

(d, J = 48)

73.8 ppm

(d, J = 49)

F 58.4 ppm (s)

65.8 ppm (s)

62.4 ppm (s)

68.6 ppm (s)

57.2 ppm (s)

66.5 ppm (s)

*C6D6, 300 or 300+ MHz spectrometers; see labeling guide A-F on the next page

Table A1.13: 31P{1H} NMR Data for Nickel-Bound Species

Page 213: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

197

Select NMR Spectra:

NMR Labeling Guide-

A = dcpe-Ni(COD), B = dcpe-Ni-ethylene, C/D = Ni complexes with only isocyanate(s),

1. R = Et

dcpe-Ni(COD) + C2H4, 90 min., R.T.

+ Et-N=C=O

1 hr., R.T.

12 hrs., R.T.

3 hrs., 50 °C

12 hrs., 60 °C

12 hrs., 70 °C

A B

C C

E E

F F

Figure A1.28: 31P{1H} NMR spectra in C6D6 for the stepwise addition of 2 atm ethylene followed by 1

equiv. ethyl isocyanate. 300 MHz spectrometer.

E = dcpe-Ni-η2-acrylamide F = dcpe-Ni-metallocycle

Page 214: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

198

Figure A1.30: 31P{1H} NMR spectra in C6D6 for the stepwise addition of 2 atm ethylene followed by 10

equiv. ethyl isocyanate, at room temperature for 2 days. 300 MHz spectrometer.

dcpe-Ni(COD) + C2H4, 30 min.

+ EtN=C=O

+ 2 hrs.

+ 12 hrs.

+ 48 hrs.

A B

C C

E E

D D

Figure A1.29: 1H NMR spectrum in C6D6 after CO gas addition to the ethylene/ethyl isocyanate coupling

reaction from Figure A1.28. Red box indicates potential acrylamide product, green indicates potential

succinimide. 500 MHz spectrometer.

Page 215: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

199

Figure A1.31: 1H NMR spectrum in C6D6 for the confirmed N-ethyl acrylamide produced from nickel-

promoted ethylene/ethyl isocyanate coupling. (A) is the full spectrum, (B) is the expanded olefinic region.

Peak for –NH is likely too broad to observe. 500 MHz spectrometer.

A.

B.

-CH=CH2

-CH2

-CH3

Page 216: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

200

A.

-CH2CH3

Figure A1.32: 1H NMR spectrum in C6D6 for the confirmed isolated N-ethyl succinimide produced from

nickel-promoted ethylene/ethyl isocyanate/CO coupling. (A) is the full spectrum, (B) is the expanded

characteristic ethyl –CH2 peak. Peaks for ethyl –CH3 and ring –CH2 groups are in the overlapping 1-2.1

ppm region. 500 MHz spectrometer.

B.

Page 217: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

201

Figure A1.33: 31P{1H} NMR spectra in C6D6 for the coupling reaction of ethylene and ethyl isocyanate

(blue) versus the reaction of dcpe-Ni(COD) with N-ethyl acrylamide (red). Slightly shifted due to

reaction mixture, but matches the nickel η2-acrylamide complex. 300 MHz spectrometer.

E E

E

E

Figure A1.34: 31P{1H} NMR spectra in C6D6 for the coupling reaction of ethylene and ethyl isocyanate

using C2H4 (blue) versus 13C2H4 (red). Only the boxed peaks showed splitting changes due to

incorporation of 13C2H4. 300 MHz spectrometer.

dcpe-Ni(COD) + C2H4 + EtN=C=O

dcpe-Ni(COD) + H2C=CHC(O)NHEt

dcpe-Ni(COD) + C2H4 + EtN=C=O

dcpe-Ni(COD) + 13C2H4 + EtN=C=O

B

A

Page 218: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

202

A.

B.

J = 12 Hz J = 30 Hz J = 57 Hz

J = 21 Hz

J = 57 Hz

J = 2 Hz

Page 219: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

203

Figure A1.35: A. 31P{1H} and B. 13C{1H} NMR spectra (expanded sections, 1-3) in C6D6 for the

coupling reaction of ethylene and ethyl isocyanate using 13C2H4. Matching coupling constant denote

corresponding species. 31P: 300 MHz spectrometer, 13C: 500 MHz spectrometer.

J = 22 Hz

J = 12 Hz

J = 31 Hz

J = 32 Hz

Page 220: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

204

Figure A1.36: 1HNMR spectrum in C6D6 indicating the ethyl isocyanate cyclotrimerization product. 500

MHz spectrometer.

Figure A1.37: 31P{1H} NMR spectrum in C6D6 for the coupling reaction between 8 atm ethylene and 3

equiv. Et-N=C=O. 300 MHz spectrometer.

B

C C

D D

Page 221: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

205

Entry Et-N=C=O

(Equiv.)

C2H4 (psi) Products

1 5 300 η2-acrylamide: yes

Isocyanurate: yes

2 2 500 η2-acrylamide: no

Isocyanurate: no

3 10 500 η2-acrylamide: yes

Isocyanurate: yes

4b 10 500 η2-acrylamide: yes

Isocyanurate: yes

5 5 600 η2-acrylamide: yes

Isocyanurate: yes

6 10 750 η2-acrylamide: no

Isocyanurate: no

7 50 750 η2-acrylamide:

trace

Isocyanurate: yes

8c 50 750 η2-acrylamide:

trace

Isocyanurate: yes

9 2 1000 η2-acrylamide: no

Isocyanurate: no

10 5 1000 η2-acrylamide: no

Isocyanurate: no

11 10 1000 η2-acrylamide: no

Isocyanurate: no

Table A1.14: Nickel-promoted reaction of ethylene and ethyl isocyanate under high

pressure conditions.a

aReaction conditions: 10 mg dcpe-Ni(COD) in ~5 mL C6H6 with X equiv. EtN=C=O and Y psi of C2H4

at 60 °C for 16 hrs. bIn THF solvent. cWith 1 equiv. sodium 3-fluorophenoxide base.

Page 222: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

206

E E

A

B

*

Figure A1.38: 31P{1H} NMR spectra in C6D6 for the Parr coupling reaction between 300 psi (blue) or

100 psi (red) ethylene and 5 or 2 equiv. Et-N=C=O, respectively. * = unidentified side product. 300 MHz

spectrometer.

Figure A1.39: 1H NMR spectra in C6D6 for the Parr coupling reaction between 300 psi (blue) or 100 psi

(red) ethylene and 5 or 2 equiv. Et-N=C=O, respectively. Boxes indicate peaks for the isocyanurate. 500

MHz spectrometer.

Page 223: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

207

2. R = tBu

Free

COD

Free

COD

Free tBuNCO

-CH=CH2

Figure A1.40: 1H NMR spectrum in C6D6 for the coupling reaction between 1 atm ethylene and 10 equiv. tBu-N=C=O, after CO addition. 500 MHz spectrometer.

E E F

C

B

dcpe-Ni(COD) + C2H4 + tBuN=C=O

dcpe-Ni(COD) + tBuN=C=O

Figure A1.41: 31P{1H} NMR spectra in C6D6 for the coupling reaction between 1 atm ethylene and 10

equiv. tBu-N=C=O with simultaneous (blue) or stepwise (red and green) reagent additions. 300 MHz

spectrometer.

Page 224: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

208

3. R = Ph

* * *

Free

COD Free

COD

Free

PhNCO

Figure A1.42: 1H NMR spectrum in C6D6 for the coupling reaction between 1.5 atm ethylene and 2

equiv. Ph-N=C=O, after CO addition. Green box indicates potential succinimide, the red box shows lack

of vinylic peaks. 600 MHz spectrometer.

F F

dcpe-Ni(COD) + C2H4 + PhN=C=O

dcpe-Ni(COD) + C2H4 + PhN=C=O + NaBAr4F, immediate

dcpe-Ni(COD) + C2H4 + PhN=C=O + NaBAr4F, 1 hr. at 60 °C

Figure A1.43: 31P{1H} NMR spectra in C6D6 for the coupling reaction between 1.5 atm ethylene and 2

equiv. Ph-N=C=O, without (blue) and with (red and green) 1 equiv. Lewis acid. New species in red and

green spectra did not yield acrylamide. Suspected C/D-type complexes. 300 MHz spectrometer.

Page 225: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

209

Figure A1.44: 1H NMR spectrum in C6D6 for the coupling reaction between 1 atm ethylene and 2 equiv.

Ph-N=C=O in the presence of 1 equiv. Lewis acid, after CO addition. No acrylamide generated. 500 MHz

spectrometer.

F F

dcpe-Ni(COD) + C2H4 + PhN=C=O

dcpe-Ni(COD) + C2H4 + PhN=C=O + NEt3

12 hrs., 60 °C

Figure A1.45: 31P{1H} NMR spectra in C6D6 for the coupling reaction between 1 atm ethylene and 2

equiv. Ph-N=C=O, without (blue) and with (red) 2 equiv. NEt3. 300 MHz spectrometer.

Page 226: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

210

dcpe-Ni(COD) + C2H4 + PhN=C=O

+ DBU, 20 hrs., 60 °C

F F

B

Figure A1.46: 31P{1H} NMR spectra in THF-d8 for the coupling reaction between 1 atm ethylene and 2

equiv. Ph-N=C=O, before (blue) and after (red) addition of 2 equiv. DBU. New species in red did not

yield acrylamide. 300 MHz spectrometer.

Figure A1.47: 1H NMR spectrum in THF-d8 for the coupling reaction between 1 atm ethylene and 2

equiv. Ph-N=C=O in the presence of 2 equiv. DBU, after CO addition. Olefinic region. 2D NMR

indicates the boxed peaks are not acrylamide. 600 MHz spectrometer.

Page 227: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

211

B

A

F F

dcpe-Ni(COD) + C2H4 + PhN=C=O

+ NaOFPh, 12 hrs., 60 °C

Figure A1.48: 31P{1H} NMR spectra in THF-d8 for the coupling reaction between 1 atm ethylene and 4

equiv. Ph-N=C=O, before (blue) and after (red) addition of 2 equiv. NaOFPh (sodium 3-

fluorophenoxide). New species in red did not yield acrylamide. 300 MHz spectrometer.

Figure A1.49: 1H NMR spectrum in THF-d8 for the coupling reaction between 1 atm ethylene and 4

equiv. Ph-N=C=O in the presence of 2 equiv. sodium 3-fluorophenoxide, after CO addition. 2D NMR

indicates the boxed peaks are not acrylamide. 600 MHz spectrometer.

Page 228: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

212

F F

B

dcpe-Ni(COD) + C2H4 + PhN=C=O

+ NaOtBu, 12 hrs., 60 °C

Figure A1.50: 31P{1H} NMR spectra for the coupling reaction between 1 atm ethylene and 2 equiv. Ph-

N=C=O, before (blue) and after (red) addition of 2 equiv. NaOtBu. Blue = C6D6, red = THF-d8. 300 MHz

spectrometer.

A.

Page 229: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

213

C.

B.

Figure A1.51: 1H NMR (A), HSQC (B), and COSY (C) spectra in THF-d8 for the coupling reaction

between 1 atm ethylene and 2 equiv. Ph-N=C=O in the presence of 2 equiv. NaOtBu. Red boxes indicated

acrylamide vinylic peaks. 500 MHz spectrometer.

-CH=CH2 -CH=CH2

Page 230: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

214

1.3: Transition-Metal-Catalyzed Hydrogenation of CO2 to Methanol

General Considerations

All manipulations were carried out under a nitrogen or argon atmosphere using standard

Schlenk, vacuum, cannula, or glovebox techniques. Catalysts 1, 2, and 1-formate were

prepared as previously described.15,16 All other chemicals were purchased from Aldrich,

Fisher, Strem, Oakwood Chemicals, VWR, or Cambridge Isotope Laboratories. Liquid

amine substrates and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) were dried over calcium

hydride or sodium hydride, purified by vacuum transfer or distillation, and stored over 3 Å

molecular sieves. Lithium triflate (LiOTf) was gently heated overnight under vacuum to

remove trace water. Molecular sieves (3 Å) for catalytic reactions were rigorously dried

under vacuum for 48-72 hours. Bulk solvents were dried and deoxygenated using literature

procedures.3 NMR solvents were dried over 3 Å molecular sieves and then used without

further manipulation, or sodium and then vacuum transferred prior to use. Hydrogen and

carbon dioxide were purchased from Airgas and were used as received. 1H, 13C and 31P

NMR spectra were recorded on Bruker 300 MHz Avance II+, 300 MHz DRX, 500 MHz

DRX or 600 MHz spectrometers at ambient temperature, unless otherwise noted. Chemical

shifts are reported in ppm; J values are given in Hz. 1H and 13C chemical shifts are

referenced to residual solvent signals; 31P chemical shifts are referenced to an external

standard of H3PO4. Probe temperatures were calibrated using ethylene glycol and methanol

as previously described.4 Gas chromatography was performed on a Thermofisher Scientific

Trace 1300 Series gas chromatograph with a FID or on an Agilent gas chromatogram with

mass spectrometer, both using helium as the carrier gas. All high-pressure CO2

hydrogenation reactions were performed using a Parr 5500 series compact reactor with

Page 231: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

215

glass insert. X-ray crystallographic data was collected on a Bruker SMART CCD or a

Bruker Prospector diffractometer with Apex II detectors. Samples were collected in inert

oil and quickly transferred to a cold gas stream. The structures were solved from direct

methods and Fourier syntheses, then refined by full-matrix least-squares procedures with

anisotropic thermal parameters for all non-hydrogen atoms. Crystallographic calculations

were carried out using SHELXTL.

Page 232: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

216

General Procedure for Catalytic Reactions:

Step A: Inside of a glovebox, weigh out 2.02 g of 3Å molecular sieves and 21 mmol

morpholine into a glass Parr insert. Add 8 mL solvent and 15 µmol catalyst (300 µl of a

0.5M solution), seal the reactor, and remove from the glovebox. Pressurize with the desired

amount of CO2, then H2. Heat at the optimized temperature with stirring for 16 hours. Cool

the reactor in an ice bath for 45 minutes, then partially vent to 100 psi. Bring the reactor

back inside of the glovebox, fully vent it, and open it. Filter the reaction solution, and then

analyze an aliquot by GC-FID using mesitylene as a standard (100 µl of reaction solution

with 0.024M or 0.0024M mesitylene standard after final dilution to 1 mL total volume).

Keep the remaining step A solution in the glovebox in preparation for step B.

Step B: Inside of a glovebox, weigh out 1 mmol LiOTf and 2.5 mmol DBU into a glass

Parr insert. Add the volume of reaction solution from step A corresponding to the desired

amount of 4-formylmorpholine for step B and dilute with solvent to a final volume of 5

mL. Inject 10 µmol catalyst (100 µl of a 0.5M solution) and seal the reactor. Outside of the

glovebox, pressurize the reactor with the same amount of H2 as in step A and heat to the

same temperature (with stirring) for 16 hours. Cool the reactor in an ice bath for 45 minutes,

then vent and open. Filter the reaction solution and then analyze by GC-FID using a

mesitylene standard (100 µl of reaction solution with 0.024M or 0.0024M mesitylene

standard after final dilution to 1 mL total volume).

Page 233: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

217

Synthesis Procedure for Generating Morphylammonium Morphylcarbamate:

Transfer morpholine and tetrahydrofuran (THF) into a glass Parr insert and seal the reactor.

Pressurize with 1000 psi CO2 and stir for 16 hours at 120 °C. Cool the reactor in an ice

bath for 45 minutes, then open and quickly filter out the resulting white solid as the product.

X-ray quality crystals were obtained from a highly concentrated solution in 1,4-dioxane.

Figure A1.52: Molecular structure of morphylammonium morphylcarbamate with ellipsoids at 50%

probability. Select bond distances, Å: (C7)-(O1) = 1.27, (C7)-(O2) = 1.26, (C7)-(N1) = 1.39.

Page 234: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

218

Reaction Optimization Tables For:

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONd Additive

1Ac Morph.

(21)

100/900 15 4.87 - 325 Mol.

sieves

1Bc FM

(1.25)

0/900 10 0.44 0.81 81 -

2A Morph.

(21)

100/900 15 12.57 - 838 -

2B FM

(1.25)

0/900 10 1.20 0.14 14 -

3A Morph.

(21)

100/900 15 12.49 - 832 -

3B FM

(1.25)

0/900 10 1.48 - - Mol.

sieves

4A Morph.

(21)

100/900 15 4.84 - 322 Mol.

sieves

4B FM

(1.25)

0/900 10 1.00 - - Mol.

sieves

Table A1.15: Effect of Molecular Sievesa

aReaction conditions: Step A: 21 mmol morpholine with or without 2.02 g molecular sieves, 15 µmol

catalyst 1 (0.07 mol%) in 8 mL dioxane, 100 psi CO2, 900 psi H2, 120 °C, 16 hrs. Step B: 1.25 mmol

FM from step A solution, 10 µmol catalyst 1 (0.8 mol%), 1 mmol LiOTf, 2.5 mmol DBU, with or

without 2.02 g molecular sieves in 5 mL total volume of solution, 100 psi H2, 120 °C, 16 hrs. bEach

entry represents only one trial unless otherwise indicated. cAverage of two trials. dTurnover number

determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 235: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

219

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONd LiOTf

(mmol)

1Ac Morph.

(21)

100/900 15 4.87 - 325 -

1Bc FM

(1.25)

0/900 10 0.44 0.81 81 1

5A Morph.

(21)

100/900 15 5.23 - 348 -

5B FM

(1.25)

0/900 10 1.17 0.08 8 -

6A Morph.

(21)

100/900 15 1.86 - 124 1

6B FM

(1.25)

0/900 10 0.28 1.07 107 1

Table A1.16: Effect of LiOTfa

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL dioxane, with or without 1 mmol LiOTf, 100 psi CO2, 900 psi H2, 120 °C, 16 hrs. Step

B: 1.25 mmol FM from step A solution, 10 µmol catalyst 1 (0.8 mol%), with or without 1 mmol LiOTf,

2.5 mmol DBU in 5 mL total volume of solution, 100 psi H2, 120 °C, 16 hrs. bEach entry represents

only one trial unless otherwise indicated. cAverage of two trials. dTurnover number determined by GC-

FID analysis of formamide in step A and methanol in step B.

Page 236: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

220

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONd DBU

(mmol)

1Ac Morph.

(21)

100/900 15 4.87 - 325 -

1Bc FM

(1.25)

0/900 10 0.44 0.81 81 2.5

7A Morph.

(21)

100/900 15 5.02 - 334 -

7B FM

(1.25)

0/900 10 1.38 - - -

8Ac Morph.

(21)

100/900 15 11.90 - 793 2.5

8Bc FM

(1.25)

0/900 10 0.32 0.55 55 2.5

9Ac Morph.

(21)

100/900 15 4.80 - 320 -

9Bc FM

(1.25)

0/900 10 0.01 0.84 84 5.0

10Ac Morph.

(21)

100/900 15 9.16 - 610 5.0

10Bc FM

(1.25)

0/900 10 1.77 - - -

Table A1.17: Effect of DBUa

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL dioxane, with or without 2.5-5 mmol DBU, 100 psi CO2, 900 psi H2, 120 °C, 16 hrs.

Step B: 1.25 mmol FM from step A solution, 10 µmol catalyst 1 (0.8 mol%), 1 mmol LiOTf, with or

without 2.5-5 mmol DBU in 5 mL total volume of solution, 100 psi H2, 120 °C, 16 hrs. bEach entry

represents only one trial unless otherwise indicated. cAverage of two trials. dTurnover number

determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 237: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

221

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONd Carbamate

(mmol)

1Ac Morph.

(21)

100/900 15 4.87 - 325 -

1Bc FM

(1.25)

0/900 10 0.44 0.81 81 2.5

11A Morph.

(21)

100/900 15 5.94 - 396 -

11B FM

(1.25)

0/900 10 1.21 0.04 4 1.25

12Ae Carbamate

(3)

100/900 15 0.64 - 42

12Be Carbamate

(1.25)

0/900 10 0.11 - -

Table A1.18: Effect of Morphylammonium Morphylcarbamatea

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL dioxane, 100 psi CO2, 900 psi H2, 120 °C, 16 hrs. Step B: 1.25 mmol FM from step A

solution, 10 µmol catalyst 1 (0.8 mol%), 1 mmol LiOTf, 2.5 mmol DBU, with or without 1.25 mmol

morphyl carbamate in 5 mL total volume of solution, 100 psi H2, 120 °C, 16 hrs. bEach entry represents

only one trial unless otherwise indicated. cAverage of two trials. dTurnover number determined by GC-

FID analysis of formamide in step A and methanol in step B. eCarbamate substrate in both steps.

Page 238: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

222

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONc Solvent

1A Morph.

(21)

100/900 15 4.42 - 294 THF

1B FM

(1.25)

0/900 10 0.01 1.13 113

2A Morph.

(21)

100/900 15 4.30 - 286 Dioxane

2B FM

(1.25)

0/900 10 0.03 1.14 113

3A Morph.

(21)

100/900 15 3.30 - 220 Toluene

3B FM

(1.25)

0/900 10 0.01 1.29 129

4A Morph.

(21)

100/900 15 3.37 - 224 Ethyl

Acetate

4B FM

(1.25)

0/900 10 0.09 0.37 37

Table A1.19: Solvent Screena

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL solvent, 100 psi CO2, 900 psi H2, 120 °C, 16 hrs. Step B: 1.25 mmol FM from step A

solution, 10 µmol catalyst 1 (0.8 mol%), 1 mmol LiOTf, 2.5 mmol DBU in 5 mL total volume of

solution, 100 psi H2, 120 °C, 16 hrs. bAverage of two trials unless otherwise indicated. cTurnover

number determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 239: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

223

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONd Max.

TON

1A Morph.

(21)

100/900 15 4.42 - 294 1400

1B FM

(1.25)

0/900 10 0.01 1.13 113 125

5A Morph.

(21)

200/800 15 6.91 - 460 1400

5B FM

(1.25)

0/800 10 0.01 1.08 108 125

6Ac Morph.

(21)

200/800 15 6.86 - 457 1400

6Bc FM

(3.05)

0/800 10 0.25 2.34 234 305

7Ac Morph.

(21)

200/1200 15 9.44 - 629 1400

7Bc FM

(4.20)

0/1200 10 0.10 3.52 352 420

8A Morph.

(21)

200/1200 15 9.24 - 616 1400

8B FM

(4.00)

0/1200 10 0.05 3.40 340 400

Table A1.20: Pressure Screena

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL THF, X psi CO2, Y psi H2, 120 °C, 16 hrs. Step B: 1.25-4 mmol FM from step A

solution, 10 µmol catalyst 1 (0.25-0.8 mol%), 1 mmol LiOTf, 2.5 mmol DBU in 5 mL total volume of

solution, Y psi H2, 120 °C, 16 hrs. bAverage of two trials unless otherwise indicated. cOnly one trial. dTurnover number determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 240: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

224

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONc LiOTf,DBU

(mmol)

8A Morph.

(21)

200/1200 15 9.24 - 616 -

8B FM

(4.00)

0/1200 10 0.05 3.40 340 1, 2.5

9A Morph.

(21)

200/1200 15 9.19 - 612 -

9B FM

(4.00)

0/1200 10 0.20 3.27 327 1, 4.0

10A Morph.

(21)

200/1200 15 9.29 - 619 -

10B FM

(4.00)

0/1200 10 0.09 3.41 341 1.5, 2.5

11A Morph.

(21)

200/1200 15 9.28 - 618 -

11B FM

(4.00)

0/1200 10 0.12 3.20 320 1.5, 4.0

Table A1.21: LiOTf/DBU Screena

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL THF, 200 psi CO2, 1200 psi H2, 120 °C, 16 hrs. Step B: 4 mmol FM from step A

solution, 10 µmol catalyst 1 (0.25 mol%), 1-1.5 mmol LiOTf, 2.5-4 mmol DBU in 5 mL total volume

of solution, 1200 psi H2, 120 °C, 16 hrs. bAverage of two trials unless otherwise indicated. cTurnover

number determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 241: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

225

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONd Catalyst

8A Morph.

(21)

200/1200 15 9.24 - 616 1

8B FM

(4.00)

0/1200 10 0.05 3.40 340 1

12A Morph.

(21)

200/1200 15 3.20 - 213 2

12B FM

(4.00)

0/1200 10 - - - NA

13Ac Morph.

(21)

500/500 15 2.94 - 196 2

13Bc FM

(4.00)

0/500 10 0.01 1.29 129 NA

14Ac Morph.

(21)

500/500 5 1.67 - 334 2

14Bc FM

(4.00)

0/500 10 - - - NA

Table A1.22: Catalyst Screena

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 5-15 µmol catalyst 1 or 2

in 8 mL THF, 200 psi CO2, 1200 psi H2 (or 500:500), 120 °C, 16 hrs. Step B: 4 mmol FM from step A

solution, 10 µmol catalyst 1 (0.25 mol%), 1 mmol LiOTf, 2.5 mmol DBU in 5 mL total volume of

solution, 1200 psi H2, 120 °C, 16 hrs. bAverage of two trials unless otherwise indicated. cOnly one trial. dTurnover number determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 242: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

226

Entryb Substrate

(mmol)

CO2/H2

(psi)

Catalyst

(µmol)

FM

(mmol)

CH3OH

(mmol)

TONc Temp.

(°C)

8A Morph.

(21)

200/1200 15 9.24 - 616 120

8B FM

(4.00)

0/1200 10 0.05 3.40 340 120

15A Morph.

(21)

200/1200 15 16.76 - 1117 100

15B FM

(4.00)

0/1200 10 0.01 3.60 360 100

16A Morph.

(21)

200/1200 15 15.24 - 1016 80

16B FM

(4.00)

0/1200 10 0.44 3.22 322 80

Table A1.23: Temperature Screena

aReaction conditions: Step A: 21 mmol morpholine, 2.02 g molecular sieves, 15 µmol catalyst 1 (0.07

mol%) in 8 mL THF, 200 psi CO2, 1200 psi H2, Z °C, 16 hrs. Step B: 4 mmol FM from step A solution,

10 µmol catalyst 1 (0.25 mol%), 1 mmol LiOTf, 2.5 mmol DBU in 5 mL total volume of solution,

1200 psi H2, Z °C, 16 hrs. bAverage of two trials unless otherwise indicated. cTurnover number

determined by GC-FID analysis of formamide in step A and methanol in step B.

Page 243: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

227

Appendix 1 References

1 Crépy, K.V.L.; Imamoto, T. “PREPARATION OF (S,S)-1,2-BIS(tert-BUTYLMETHYLPHOSPHINO)

ETHANE ((S,S)-t-Bu-BISP*) AS A RHODIUM COMPLEX,” Organic Syntheses, 2005, 82, 22-29.

2 Yakelis, N.A.; Bergman, R.G. “Safe Preparation and Purification of Sodium Tetrakis[(3,5-

trifluoromethyl)phenyl]borate (NaBArF24): Reliable and Sensitive Analysis of Water in Solutions of

Fluorinated Tetraarylborates,” Organometallics, 2005, 24, 3579-3581.

3 Pangborn, A.B.; Giardello, M.A.; Grubbs, R.H.; Rosen, R.K.; Timmers, F.J. “Safe and Convenient

Procedure for Solvent Purification,” Organometallics, 1996, 15, 1518-1520.

4 Sandström, J. Dynamic NMR Spectroscopy; Academic Press: New York, 1982.

5 Hoberg, H.; Schaefer, D. “NICKEL(0)-INDUZIERTE C-C-VERKNÜPFUNG ZWISCHEN

KOHLENDIOXID UND ETHYLEN SOWIE MONO-ODER DI-SUBSTITUIERTEN ALKENEN,” J.

Organomet. Chem., 1983, 251, C51-C53.

6 Jin, D.; Williard, P.G.; Hazari, N.; Bernskoetter, W.H. “Effect of Sodium Cation on Metallocycle β-Hydride

Elimination in CO2-Ethylene Coupling to Acrylates,” Chem. Eur. J., 2014, 20, 3205-3211.

7 Jolly, P.W.; Jonas, K.; Krüger, C.; Tsay, Y.-H. “THE PREPARATION, REACTIONS, AND STRUCTURE

OF BIS[BIS-(TRICYCLOHEXYLPHOSPHINE)NICKEL] DINITROGEN, {[(C6H11)3P]2Ni}2N2,” J.

Organomet. Chem., 1971, 33, 109-122.

8 Aresta, M.; Nobile, C.F.; Albano, V.G.; Forni, E.; Manassero, M. “New Nickel-Carbon Dioxide Complex:

Synthesis, Properties, and Crystallographic Characterization of (Carbon dioxide)-

bis(tricyclohexylphosphine)nickel,” J.C.S. Chem. Comm., 1975, 636-637.

9 Maier, L. “The Direct Synthesis of Tris(N-Substituted Carbamoylethyl) Phosphine Oxides,” Helvetica

Chimica Acta, 1973, 56, 1252-1257.

10 Krivec, M.; Gazvoda, M.; Kranjc, K.; Polanc, S.; Kočevar, M. “A Way to Avoid Using Precious Metals:

The Application of High-Surface Activated Carbon for the Synthesis of Isoindoles via the Diels-Alder

Reaction of 2H-Pyran-2-ones,” J. Organic Chem., 2012, 77, 2857-2864.

11 Dokli, I.; Gredičak, M.; “Mechanochemical Ritter Reaction: A Rapid Approach to Functionalized Amides

at Room Temperature,” Chem. Eur. J., 2015, 2727-2732.

12 Garad, D.N.; Tanpure, S.D.; Mhaske, S.B. “Radical-mediated dehydrative preparation of cyclic imides

using (NH4)2S2O8-DMSO: application to the synthesis of vernakalant,” Beilstein J. Org. Chem., 2015, 11,

1008-1016.

13 Tang, X.-J.; Dolbier, W.R. “Efficient Cu-Catalyzed Atom Transfer Radical Addition Reactions of

Fluoroalkylsulfonyl Chlorides with Electron-deficient Alkenes Induced by Visible Light,” Angew. Chem. Int.

Ed., 2015, 54, 4246-4249.

14 Mo, F.; Yan, J.M.; Qiu, D.; Li, F.; Zhang, Y.; Wang, J. “Gold-Catalyzed Halogenation of Aromatics by N-

Halosuccinimides,” Angew. Chem. Int. Ed., 2010, 49, 2028-2032.

Page 244: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

228

15 Bielinski, E.A.; Lagaditis, P.O.; Zhang, Y.; Mercado, B.Q.; Würtele, C.; Bernskoetter, W.H.; Hazari, N.;

Schneider, S. “Lewis Acid-Assisted Formic Acid Dehydrogenation Using a Pincer-Supported Iron Catalyst,”

J. Am. Chem. Soc., 2014, 136, 10234-10237.

16 Werkmeister, S.; Junge, K.; Wendt, B.; Alberico, E.; Jiao, H.; Baumann, W.; Junge, H.; Gallou,F.; Beller,

M. “Hydrogenation of Esters to Alcohols with a Well-Defined Iron Complex,” Angew. Chem. Int. Ed., 2014,

53, 8722-8726.

Page 245: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

229

Appendix 2

2.2: Iron-Catalyzed Amide Formation from the Dehydrogenative Coupling of

Alcohols and Secondary Amines

General Considerations

All manipulations were carried out under a nitrogen or argon atmosphere using standard

Schlenk, vacuum, cannula, or glovebox techniques. Catalysts 1 and

(iPrPNHP)Fe(H)(HCO2)(CO) (1-formate) were prepared as previously described.1 Amide

products (for gas chromatography response factors) that were not commercially available

were prepared using previously reported procedures.2,3 All other chemicals were purchased

from Aldrich, Fisher, Strem, Synthonix, Oakwood Chemicals, VWR, or Cambridge

Isotope Laboratories. Liquid amine and alcohol substrates were dried over calcium hydride

or sodium hydride, purified by vacuum transfer or distillation, and stored over 3 Å

molecular sieves. Solid substrates were purified by sublimation, followed by

recrystallization (if necessary). Bulk solvents were dried and deoxygenated using literature

procedures.4 NMR solvents were dried over 3 Å molecular sieves and then used without

further manipulation, or sodium and then vacuum transferred prior to use. Hydrogen and

carbon dioxide were purchased from Airgas and were used as received. 1H, 13C and 31P

NMR spectra were recorded on Bruker 300 MHz Avance II+, 300 MHz DRX, 500 MHz

DRX or 600 MHz spectrometers at ambient temperature, unless otherwise noted. Chemical

shifts are reported in ppm; J values are given in Hz. 1H and 13C chemical shifts are

referenced to residual solvent signals; 31P chemical shifts are referenced to an external

standard of H3PO4. Probe temperatures were calibrated using ethylene glycol and methanol

Page 246: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

230

as previously described.5 Gas chromatography was performed on a Thermofisher Scientific

Trace 1300 Series gas chromatograph with a FID using helium as the carrier gas.

General Procedure for the Formation of Complex 2

In a dry box, 5 mg of complex 1 was dissolved in either C6D6 or THF-d8 in a 20-mL

scintillation vial and then transferred to a J-Young tube. The tube was then sealed, removed

from the glovebox, degassed, and 6 equivalents of methanol or 13C-labeled methanol were

added via calibrated gas bulb.

Select NMR data for 2 (at -80 °C) 1H NMR (THF-d8, 300 MHz): -24.3 (t, 1H, J = 52, Fe-

H); 13C{1H} NMR (THF-d8, 300 MHz): 59.58 (s, Fe-O-CH3); 13C NMR (THF-d8, 300

MHz): 59.57 (q, J = 132, Fe-O-CH3);31P{1H} NMR (THF-d8, 300 MHz): 93.21 (d, J = 32).

General Methods for Catalytic Alcohol Dehydrogenation in the Presence of Amines

In a dry box, a 100-mL Schlenk tube was loaded with 5 mL of tetrahydrofuran (THF), 12

mmol of amine, 3 µmol of catalyst, and 3 mmol of alcohol, then sealed. It was immediately

placed in an oil bath preheated to 80 °C and stirred for 8 hours. It was then cooled in an ice

bath for 30 minutes prior to analysis. If analyzed using NMR spectroscopy, 100 µl of

reaction solution were added to an NMR tube with 395 µl CDCl3 and 5 µl of mesitylene

standard and an NMR delay time of 60 seconds was used. If analyzed by GC, 100 µl of

reaction solution was diluted to 1 mL with THF and a mesitylene standard (0.024 M or

0.0024M after final dilution) was added.

Page 247: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

231

General Procedure for H2 Collection Studies

In a dry box, a 100-mL Schlenk flask was loaded with 2.5 mL THF, 6 mmol of morpholine,

1.5 µmol of catalyst 1, and 1.5 mmol methanol. The flask was fitted with a reflux condenser

and an adaptor with a stopcock, removed from the glovebox, and connected to a gas burette

setup that had been pre-sparged with N2. The connecting hoses/trap were subjected to two

evacuation/refill cycles, the trap was cooled in a dry ice/acetone bath, and a small amount

of vacuum was used to reset the starting water volume. The Schlenk flask was lowered into

an oil bath (preheated to 80 °C) and the system was allowed to equilibrate for 3.5 minutes,

after which a small amount of vacuum was used to restore the water volume in the buret

(with the connection to the reflux setup closed). The stopcock was then opened, the reaction

was allowed to proceed for 8 hours, and the change in the water level in the gas burette

was used to determine the TON. The procedure for determining TON has been previously

reported.1

Page 248: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

232

Reaction Optimization Tables for:

Entry Solvent TON

1 Acetonitrile 30

2 Ethyl Acetate 49b

3 Toluene 297b

4 Dioxane 344b

5 Tetrahydrofuran, 10 mL 392

6 Tetrahydrofuran, 5 mL 503

Table A2.1: Reaction Solventa

aReaction conditions: 3 µmol catalyst 1 (0.1 mol%), 3 mmol

alcohol, and 12 mmol amine in 10 mL solvent at 80 °C for 8

hrs. Each entry is an average of two trials unless otherwise

indicated. TON determined by NMR analysis of the production

of amide. bOnly one trial.

Page 249: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

233

Entry Time (hrs) TON

1 1 176b

2 4 383

3 8 503

4 16 553

Entry Temp. (°C) TON

1 R.T. 10

2 60 172

3 80 503

4 100 328

5 120 341

aReaction conditions: 3 µmol catalyst 1 (0.1 mol%), 3

mmol alcohol, and 12 mmol amine in 5 mL THF at Y °C

for 8 hrs. Each entry is an average of two trials unless

otherwise indicated. TON determined by NMR analysis

of the production of amide.

Table A2.3: Reaction Temperaturea

Table A2.2: Reaction Timea

aReaction conditions: 3 µmol catalyst 1 (0.1

mol%), 3 mmol alcohol, and 12 mmol amine in

5 mL THF at 80 °C for Z hrs. Each entry is an

average of two trials unless otherwise indicated.

TON determined by NMR analysis of the

production of amide. bAverage of three trials.

Page 250: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

234

Entry CH3OH

(mmol)

Morpholine

(mmol)

CH3OH:

Morpholine

TON (FM)

1 3 3 1:1 261

2 3 12 1:4 503

3 3 30 1:10 425

4 12 3 4:1 163

Entry Catalyst TON

1 None 0b

2 FeCl2 0b

3 1-formate 0c

4 1 503

Table A2.5: Catalyst Control Experimentsa

aReaction conditions: 3 µmol catalyst 1 (0.1 mol%), A mmol alcohol, and B mmol amine in 5 mL

THF at 80 °C for 8 hrs. Each entry is an average of two trials unless otherwise indicated. TON

determined by NMR analysis of the production of amide.

aReaction conditions: 3 µmol catalyst X (0.1 mol%),

3 mmol alcohol, and 12 mmol amine in 5 mL THF at

80 °C for 8 hrs. Each entry is an average of two trials

unless otherwise indicated. TON determined by NMR

analysis of the production of amide. bOnly one trial.

cWith 10 mol% LiBF4 or LiOTf.

Table A2.4: Alcohol/Amine Ratioa

Page 251: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

235

Entry Amine Additive (mol%) TON

1 Morpholine LiOTf, 10 370b

2 Morpholine None 503

3 Morpholine KOtBu, 1 216b

4 Morpholine KOtBu, 15 86b

5 Diethylamine None 176

6 Diethylamine KOtBu, 1 11

Table A2.6: Effect of Base or Lewis Acid Additivea

aReaction conditions: 3 µmol catalyst 1 (0.1 mol%), 3 mmol alcohol, 12 mmol amine, and X

mol% base (potassium tert-butoxide) or Lewis acid (lithium trifluoromethanesulfonate) in 5

mL THF at 80 °C for 8 hrs. Each entry is an average of two trials unless otherwise indicated.

TON determined by NMR analysis of the production of amide. bOnly one trial.

Page 252: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

236

Entry

KOtBu (mol%) TON

1

0 790

2

1 953

3 No catalyst 0 236

4 No catalyst 1 876

Entry

KOtBu (mol%) TON

1

0 463

2

1 860

3 No catalyst 0 563

Table A2.7: Ester Substratea

Table A2.8: Aldehyde Substratea

aReaction conditions: 3 µmol catalyst 1 (0.1 mol%), 3 mmol ester, 12

mmol amine, and X mol% base (potassium tert-butoxide) in 5 mL

THF at 80 °C for 8 hrs. Each entry is an average of two trials unless

otherwise indicated. TON determined by GC analysis of the

production of amide.

aReaction conditions: 3 µmol catalyst 1 (0.1 mol%), 3 mmol aldehyde, 12

mmol amine, and X mol% base (potassium tert-butoxide) in 5 mL THF

at 80 °C for 8 hrs. Each entry is an average of two trials unless otherwise

indicated. TON determined by GC analysis of the consumption of

aldehyde.

Page 253: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

237

NMR Studies of the Formation of 2

1

1 + 2 5

B

No reaction

Figure A2.1: Sequential addition of reagents A. Room temperature 1H NMR spectra, hydride region; blue spectrum is the starting

material, red spectrum is after addition of morpholine (no change); green spectrum is

after addition of methanol. B. Room temperature 31P{1H} NMR spectra, same

sample/addition sequence. A small amount of 5 is detected due to catalyst

decomposition. Sample dissolved in benzene-d6.

A

1

1 + 2

No reaction

Page 254: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

238

-80 °C Free

13CH

3OH 2 THF THF

-80 °C

22 °C

1+2

A

-80 °C

22 °C

-10 °C

-20 °C

2

3 4 5 Free

Ligand

1+2

B

3 2

1+2

C

-80 °C

22 °C

-10 °C

-20 °C

Page 255: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

239

Figure A2.2: Low temperature NMR experiments

Low-temperature NMR experiments with 2, including A. 13

C{1H} NMR

(red and blue) and 13C NMR (green) spectra, B. 31P{1H} NMR spectra, C. 1H NMR spectra, hydride region. The Fe-H peaks for 4 are not observed

due to its low concentration. Sample dissolved in THF-d8.

Page 256: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

240

2.3: Iron-Catalyzed Urea Synthesis: Dehydrogenative Coupling of Methanol and

Amines

General Considerations:

All manipulations were carried out under a nitrogen or argon atmosphere using standard

Schlenk, vacuum, cannula, or glovebox techniques. Catalyst 1 was prepared as previously

described.1 Formamides that were not commercially available were prepared using

previously reported procedures.6 All other chemicals were purchased from Aldrich, Fisher,

Strem, Oakwood Chemicals, VWR, or Cambridge Isotope Laboratories. Liquid amine and

alcohol substrates were dried over calcium hydride or sodium hydride, purified by vacuum

transfer or distillation, and stored over 3 Å molecular sieves. Solid substrates were purified

by sublimation, followed by recrystallization (if necessary). Bulk solvents were dried and

deoxygenated using literature procedures.4 NMR solvents were dried over 3 Å molecular

sieves and then used without further manipulation, or sodium and then vacuum transferred

prior to use. Hydrogen was purchased from Airgas and was used as received. 1H, 13C and

31P NMR spectra were recorded on Bruker 300 MHz Avance II+, 300 MHz DRX, 500

MHz DRX or 600 MHz spectrometers at ambient temperature, unless otherwise noted.

Chemical shifts are reported in ppm; J values are given in Hz. 1H and 13C chemical shifts

are referenced to residual solvent signals; 31P chemical shifts are referenced to an external

standard of H3PO4. Probe temperatures were calibrated using ethylene glycol and methanol

as previously described.5 Gas chromatography was performed on a Thermofisher Scientific

Trace 1300 Series gas chromatograph with FID using helium as a carrier gas.

Page 257: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

241

General Methods for Symmetric Urea Formation from Catalytic Methanol

Dehydrogenation in the Presence of Amines

In a glovebox, a 100-mL Schlenk tube was loaded with 5 mL of tetrahydrofuran (THF), 12

mmol of amine, 15 µmol of catalyst, and 3 mmol of alcohol, then sealed. It was

immediately placed in an oil bath preheated to 120 °C and stirred for 8 hours. It was then

cooled in an ice bath for 30 minutes. If the starting amine was benzylamine or its

derivatives, all of the following sample preparations and reaction workups were performed

in a glovebox due to the air sensitivity of the remaining starting material.

Analysis of formamide yield:

If analyzed by NMR, 100 µl of reaction solution were added to an NMR tube with 395 µl

CDCl3 and 5 µl of mesitylene standard and an NMR delay time of 60 seconds was used. If

analyzed by GC, 100 µl of reaction solution were diluted to 1 mL with THF and mesitylene

standard (0.024 M or 0.0024M after final dilution) was added.

Urea isolation:

The solvent was then removed from the reaction mixture using glovebox vacuum or rotary

evaporation. The resulting oily solid was transferred to a filtration frit and washed with

room temperature pentane in a glovebox for benzylamine-type starting materials and cold

pentane in air for alkylamine starting materials. The filtrate for alkylamine starting

materials was collected, dried, and re-washed until no more solid was recovered. In the

case of diaminocyclohexane, the product was an oil that was purified with silica gel column

chromatography using 20:1 CH2Cl2:methanol as the eluent according to previously

Page 258: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

242

established procedures.7 Ureas were identified by comparison to previously reported

literature data.8-14

General Methods for Unsymmetric Urea Formation from Catalytic Dehydrogenative

Coupling of Formamides and Amines

In a glovebox, a 100-mL Schlenk tube was loaded with 5 mL of tetrahydrofuran (THF), 3

mmol of amine, 15 µmol of catalyst, and 3 mmol of formamide, then sealed. It was

immediately placed in an oil bath preheated to 120 °C and stirred for 16 hours. It was then

cooled in an ice bath for 30 minutes. If the starting amine was benzylamine or its

derivatives, all of the following sample preparations and reaction workups were performed

in a glovebox due to the air sensitivity of the remaining starting material.

Analysis of formamide yield:

If analyzed by NMR, 100 µl of reaction solution were added to an NMR tube with 395 µl

CDCl3 and 5 µl of mesitylene standard and an NMR delay time of 60 seconds was used. If

analyzed by GC, 100 µl of reaction solution were diluted to 1 mL with THF and mesitylene

standard (0.024 M or 0.0024M after final dilution) was added.

Urea isolation:

The solvent was then removed from the reaction mixture using glovebox vacuum or rotary

evaporation. The resulting oily solid was transferred to a filtration frit and washed with

room temperature pentane in a glovebox for benzylamine-type starting materials and cold

pentane in air for alkylamine starting materials. The filtrate for alkylamine starting

materials was collected, dried, and re-washed until no more solid was recovered. In the

Page 259: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

243

case of isobutylformamide as the substrate, the product was an oil that was dried under

vacuum and then analyzed by NMR spectroscopy.

Page 260: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

244

Optimization Tables for Symmetric Urea Synthesis

Table A2.9: Catalyst Loadinga

Entry Amine Catalyst

Loading

(mol%)

TON (Urea) Yield (%)

1 Cyclohexylamine 0.1 33 3%

2 Cyclohexylamine 0.25 34 9%

3 Cyclohexylamine 0.5 23 12%

Table A2.10: Alcohol:Amine Ratioa

Entry Amine Alcohol:Amine TON (Urea)

1 Cyclohexylamine 1:1 7

2 Cyclohexylamine 1:4 23

3 Cyclohexylamine 1:6 29

4 Cyclohexylamine 1:10 34

Table A2.11: Solventa

Entry Amine Solvent TON (Urea)

1 Cyclohexylamine Ethyl acetate 12

2 Cyclohexylamine 1,4-Dioxane 18

3 Cyclohexylamine Tetrahydrofuran 23

4 Cyclohexylamine Toluene 28

aReaction conditions: X mol% catalyst 1, 3 mmol alcohol, and 12 mmol amine in 5 mL THF at 80°C for

8 hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from isolated

yield of urea.

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol alcohol, and X mmol amine in 5 mL THF at 80°C for

8 hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from isolated

yield of urea. TON error is ±13.

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol alcohol, and 12 mmol amine in 5 mL solvent at 80°C

for 8 hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from isolated

yield of urea.

Page 261: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

245

Table A2.12: Temperaturea

Entry Amine Temperature (°C) TON (Urea)

1 Cyclohexylamine 60b 5

2 Cyclohexylamine 80 23

3 Cyclohexylamine 80c 28

4 Cyclohexylamine 100 55

5d Cyclohexylamine 120 66

6 Cyclohexylamine 120c 66

Table A2.13: Timea

Entry Amine Time (hrs) TON (Urea)

1 Cyclohexylamine 1 11

2b Cyclohexylamine 4 44

3b Cyclohexylamine 8 66

4b Cyclohexylamine 16 56

Table A2.14: Addition of Basea

Entry Amine DBU Additive

(mol%)

TON (Urea)

1 4-(trifluoromethyl)

benzylamine

0 90

2b 4-(trifluoromethyl)

benzylamine

5 108

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol alcohol, and 12 mmol amine in 5 mL THF at X°C for

8 hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from isolated

yield of urea. b16 hrs. cToluene solvent. dAverage of three trials.

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol alcohol, and 12 mmol amine in 5 mL THF at 120°C

for X hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from isolated

yield of urea. bAverage of three trials.

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol alcohol, and 12 mmol amine in 5 mL THF at 120°C

for 8 hrs, with X mol% 1,8-diazabicyclo(5.4.0)undec-7-ene (DBU). Each entry is an average of three

trials unless otherwise indicated. TON determined from isolated yield of urea. bOnly one trial.

Page 262: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

246

Optimization Tables for Unsymmetric Urea Synthesis

Table A2.15: Timea

Entry R R’ Time (hrs) TON (Urea)

1 Benzyl Cyclohexyl 4b 0

2 Benzyl Cyclohexyl 4 68

3 Benzyl Cyclohexyl 8 156

4 Benzyl Cyclohexyl 16 170

Table A2.16: Temperaturea

Entry R R’ Temperature

(°C)

TON (Urea)

1 Benzyl Cyclohexyl 100 162

2 Benzyl Cyclohexyl 120 170

3 Benzyl Cyclohexyl 140 138

Table A2.17: Catalyst Loadinga

Entry R R’ Catalyst

Loading

(mol%)

TON

(Urea)

Yield (%)

1 Benzyl Cyclohexyl 0.5 170 85%

2 Benzyl Cyclohexyl 1 83 83%

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol formamide, and 3 mmol amine in 5 mL THF at 120°C

for X hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from isolated

yield of unsymmetric urea. bNo catalyst.

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol formamide, and 3 mmol amine in 5 mL THF at X°C

for 16 hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from

isolated yield of unsymmetric urea.

aReaction conditions: X mol% catalyst 1, 3 mmol formamide, and 3 mmol amine in 5 mL THF at 120°C

for 16 hrs. Each entry is an average of two trials unless otherwise indicated. TON determined from

isolated yield of unsymmetric urea.

Page 263: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

247

Table A2.18: Formamide Scramblinga

Entry R R’ Starting

Formamide

Left

(mmol)

Scrambled

Formamide

(mmol)

Catalyst Yield,

Unsymm.

Urea (%)

Sel.b

(%)

1 Cyclo-

hexyl

Benzyl 1.53 0.02 1, 0.5

mol%

34% 72%

2 Cyclo-

hexyl

Benzyl 2.70 0.04 None 0 NA

aReaction conditions: 0.5 mol% catalyst 1 or no catalyst, 3 mmol formamide, and 3 mmol amine in 5 mL

THF at 120°C for 4 hrs. Each entry represents only one trial. NA = not applicable. bSelectivity determined

by amount of unsymmetric urea as compared to the amounts of the two symmetric ureas.

Page 264: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

248

Mechanistic NMR Experiments

Figure A2.3: Hydrogenation of ureas

Room temperature 13C{1H} NMR spectra in THF-d8, 600 MHz, carbonyl region; blue spectrum is straight

after mixing at room temperature, red spectrum is after 16 hrs. at 120 °C. GC-FID showed only ~1.5%

conversion to formamide.

Immediate

16 hrs. @ 120 °C

1,3-dibenzylurea

Page 265: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

249

Figure A2.4: Reaction of urea with amine under H2

Room temperature 13C{1H} NMR spectrum in THF-d8, 600 MHz, carbonyl region; after 16 hrs. at 120°C.

The unsymmetric urea product is marked by a box.

1,3-dibenzylurea

Unsymmetric Urea

Page 266: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

250

Figure A2.5: Reaction of urea with formamide under H2

Room temperature 13C{1H} NMR spectrum in THF-d8, 600 MHz, carbonyl region; after 16 hrs. at 120

°C. The unsymmetric urea product is marked by a box. The peak at 160.6 ppm is a second rotamer of the

cyclohexylformamide reactant.

Cyclohexylformamide

1,3-dibenzylurea

Unsymmetric Urea

Page 267: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

251

Figure A2.6: Reaction of urea with formamide, no H2 Room temperature 13C{1H} NMR spectrum in THF-d8, 600 MHz, carbonyl region; after 16 hrs. at 120

°C. The unsymmetric urea product is marked by a box. The peak at 160.6 ppm is a second rotamer of the

cyclohexylformamide reactant.

1,3-dibenzylurea

Cyclohexylformamide

Unsymmetric Urea

Page 268: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

252

Figure A2.7: Reaction of formamides and amines Room temperature 13C{1H} NMR spectra in THF-d8, 600 MHz, carbonyl region; blue spectrum is straight

after mixing at room temperature, red spectrum is after 2 hrs. at 120 °C, green spectrum is after 5 hrs.

Immediate

2 hrs. @ 120 °C

5 hrs. @ 120 °C

Unsymmetric Urea

Benzylformamide

Cyclohexyl

formamide

Page 269: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

253

Table A2.19: Attempt to Make a Tetrasubstituted Urea: Standard Catalytic

Conditionsa

Entry Formamide Left Urea Produced

1 3 mmol 0

Immediate

16 hrs. @ 120 °C

Figure A2.8: Attempt to make a tetrasubstituted urea: NMR-scale

Room temperature 13C{1H} NMR spectra in THF-d8, 600 MHz carbonyl region; blue spectrum is straight

after mixing at room temperature, red spectrum is after 16 hrs. at 120 °C.

No reaction

aReaction conditions: 0.5 mol% catalyst 1, 3 mmol formamide, and 12 mmol amine in 5 mL THF at 80°C

for 8 hrs. Each entry represents only one trial.

No reaction

Page 270: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

254

Figure A2.9: Reaction of pentylformamide with dipentylamine

Room temperature 13C{1H} NMR spectra in THF-d8, 600 MHz, carbonyl region; blue spectrum is straight

after mixing at room temperature, red spectrum is after 1 hr. at 120 °C, green spectrum is after 2 hrs., and

purple spectrum is after 24 hrs. Carbonyl carbon for the trisubstiuted urea is shown by the blue box. The

other carbonyl carbon is from the starting formamide. GC-FID showed presence of the trisubstituted urea

and no dipentylformamide or pentylamine from scrambling.

Immediate

24 hrs. @ 120 °C

2 hrs. @ 120 °C

1 hr. @ 120 °C

Unsymmetric Urea

Page 271: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

255

No reaction

Immediate

1 hr. @ 120 °C

2 hrs. @ 120 °C

24 hrs. @ 120 °C

Figure A2.10: Reaction of dipentylformamide with pentylamine

Room temperature 13C{1H} NMR spectra in THF-d8, 600 MHz, carbonyl region; blue spectrum is straight

after mixing at room temperature, red spectrum is after 1 hr. at 120 °C, green spectrum is after 2 hrs., and

purple spectrum is after 24 hrs. The carbonyl carbon shown is from the starting formamide. GC-FID

showed no trisubstituted urea and no pentylformamide or dipentylamine from scrambling.

Page 272: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

256

A.

2 1

3

Page 273: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

257

B.

Figure A2.11: Dehydrogenation of a formamide to an isocyanate

A. Room temperature 1H NMR spectrum in THF-d8, hydride region, after 1 hr. at 120 °C. Presence of the

iron-dihydride (trans isomer) is marked by a box. B. 31P{1H} NMR spectrum in THF-d8 after 1 hr. at 120

°C. Presence of 5 and some free ligand due to catalyst decomposition. Arrows indicate phosphorous

resonances for both the cis (red) and trans (green) dihydride isomers. 300+ MHz. GC-FID showed

presence of cyclohexyl isocyanate.

2

1

Free

Ligand

5

4 3

Page 274: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

258

NMR Spectra for Synthesized Symmetric Ureas

All 1H NMRs are taken on a 300+ MHz spectrometer and all 13C{1H} NMRs are taken on

a 600 MHz spectrometer unless otherwise indicated. * = small amount of formamide

impurity

Page 275: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

259

1,3-di-n-pentylurea (Table 2.3.1, Entry 1)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

*

*

Page 276: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

260

1,3-bis(2-methylpropyl)urea (Table 2.3.1, Entry 2)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

*

*

Page 277: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

261

1,3-di-n-heptylurea (Table 2.3.1, Entry 3)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 278: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

262

1,3-bis(2-methoxyethyl)urea (Table 2.3.1, Entry 4)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

*

*

Page 279: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

263

1,3-bis(1-methylhexyl)urea (Table 2.3.1, Entry 5)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 280: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

264

1,3-bis(4-methoxybenzyl)urea (Table 2.3.1, Entry 6)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

*

*

Page 281: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

265

1,3-bis(4-methylbenzyl)urea (Table 2.3.1, Entry 7)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 282: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

266

1,3-dibenzylurea (Table 2.3.1, Entry 8)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

*

*

Page 283: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

267

1,3-bis(4-trifluoromethylbenzyl)urea (Table 2.3.1, Entry 9)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

*

Page 284: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

268

1,3-bis(2-phenylethyl)urea (Table 2.3.1, Entry 10)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 285: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

269

1,3-dicyclopentylurea (Table 2.3.1, Entry 11)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 286: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

270

1,3-dicyclohexylurea (Table 2.3.1, Entry 12)

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 287: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

271

Octahydro-benzoimidazol-2-one (Table 2.3.1, Entry 13)

Note: observe both cis and trans isomers of the product urea

1H NMR, CDCl3

13C{1H} NMR, CDCl3

Page 288: Homogeneous Nickel and Iron Complexes for Carbon Dioxide

272

Appendix 2 References

1 Bielinski, E.A.; Lagaditis, P.O.; Zhang, Y.; Mercado, B.Q.; Würtele, C.; Bernskoetter, W.H.; Hazari, N.;

Schneider, S. “Lewis Acid-Assisted Formic Acid Dehydrogenation Using a Pincer-Supported Iron Catalyst,”

J. Am. Chem. Soc., 2014, 136, 10234-10237.

2 Badioli, M.; Ballini, R.; Bartolacci, M.; Bosica, G.; Torregiani, E.; Marcantoni, E. “Addition of

Organocerium Reagents to Morpholine Amides: Synthesis of Important Pheromone Components of Achaea

janata,” J. Org. Chem., 2002, 67, 8938-8942.

3 Ramesh, P.; Fadnavis, N.W. “Ammonium Nitrate: A Biodegradable and Efficient Catalyst for the Direct

Amidation of Esters under Solvent-free Conditions,” Chem. Lett., 2015, 44, 138-140.

4 Pangborn, A.B.; Giardello, M.A.; Grubbs, R.H.; Rosen, R.K.; Timmers, F.J. “Safe and Convenient

Procedure for Solvent Purification,” Organometallics, 1996, 15, 1518-1520.

5 Sandström, J. Dynamic NMR Spectroscopy; Academic Press: New York, 1982.

6 Dhake, K.P.; Tambade, P.J.; Singhal, R.S.; Bhanage, B.M. “An efficient catalyst- and solvent-free N-

formylation ofaromatic and aliphatic amines,” Green Chemistry Letters and Reviews, 2011, 4, 151-157.

7 Kim, S.H.; Hong, S.H. “Ruthenium-Catalyzed Urea Synthesis Using Methanol as the C1 Source,” Org.

Lett., 2016, 18, 212-215.

8 Mizuno, T.; Mihara, M.; Nakai, T.; Iwai, T.; Ito, T. “Solvent-Free Synthesis of Urea Derivatives from

Primary Amines and Sulfur under Carbon Monoxide and Oxygen at Atmospheric Pressure,” Synthesis, 2007,

20, 3135-3140.

9 Artuso, E.; Degani, I.; Fochi, R.; Magistris, C. “Preparation of Mono, Di-, and Trisubstituted Ureas by

Carbonylation of Aliphatic Amines with S,S,-Dimethyl Dithiocarbonate,” Synthesis, 2007, 22, 3497-3506.

10 Park, J.H.; Yoon, J.C.; Chung, Y.K. “Cobalt/Rhodium Heterobimetallic Nanoparticle-Catalyzed Oxidative

Carbonylation of Amines in the Presence of Carbon Monoxide and Molecular Oxygen to Ureas,” Adv. Synth.

Catal., 2009, 351, 1233-1237.

11 Zhang, M.; Imm, S.; Bähn, S.; Neubert, L.; Neumann, H.; Beller, M. “Efficient Copper(III)-Catalyzed

Transamidation of Non-Activated Primary Carboxamides and Ureas with Amines,” Angew. Chem. Int. Ed.,

2012, 51, 3905-3909.

12 Balaraman, E.; Ben-David, Y.; Milstein, D. “Unprecedented Catalytic Hydrogenation of Urea Derivatives

to Amines and Methanol,” Angew. Chem. Int. Ed., 2011, 50, 11702-11705.

13 Smirnova, E.S.; Muñoz Molina, J.M.; Johnson, A.; Bandeira, N.A.G.; Bo, C.; Echavarren, A.M.

“Polynuclear Gold [AuI]4, [AuI]8, and Bimetallic [AuI4AgI] Complexes: C-H Functionalization of Carbonyl

Compounds and Homogeneous Carbonylation of Amines,” Angew. Chem. Int. Ed., 2016, 55, 7487-7491.

14 Chiarotto, I.; Feroci, M. “Selective and Environmentally Friendly Methodologies Based on the Use of

Electrochemistry for Fine Chemical Preparation: An Efficient Synthesis of N,N’-Disubstituted Ureas,” J.

Org. Chem., 2003, 68, 7137-7139.