189
Purdue University Purdue e-Pubs Open Access Dissertations eses and Dissertations 8-2016 Mass spectrometric characterization of remotely charged amino acids and peptides Damodar Koirala Purdue University Follow this and additional works at: hps://docs.lib.purdue.edu/open_access_dissertations Part of the Chemistry Commons is document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] for additional information. Recommended Citation Koirala, Damodar, "Mass spectrometric characterization of remotely charged amino acids and peptides" (2016). Open Access Dissertations. 788. hps://docs.lib.purdue.edu/open_access_dissertations/788

Mass spectrometric characterization of remotely charged

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Mass spectrometric characterization of remotely charged

Purdue UniversityPurdue e-Pubs

Open Access Dissertations Theses and Dissertations

8-2016

Mass spectrometric characterization of remotelycharged amino acids and peptidesDamodar KoiralaPurdue University

Follow this and additional works at: https://docs.lib.purdue.edu/open_access_dissertations

Part of the Chemistry Commons

This document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] foradditional information.

Recommended CitationKoirala, Damodar, "Mass spectrometric characterization of remotely charged amino acids and peptides" (2016). Open AccessDissertations. 788.https://docs.lib.purdue.edu/open_access_dissertations/788

Page 2: Mass spectrometric characterization of remotely charged

PURDUE UNIVERSITY GRADUATE SCHOOL

Thesis/Dissertation Acceptance

To the best of my knowledge and as understood by the student in the Thesis/Dissertation Agreement, Publication Delay, and Certification/Disclaimer (Graduate School Form 32), this thesis/dissertation adheres to the provisions of Purdue University’s “Policy on Integrity in Research” and the use of copyrighted material.

Damodar Koirala

MASS SPECTROMETRIC CHARACTERIZATION OF REMOTELY CHARGED AMINOACIDS AND PEPTIDES

Doctor of Philosophy

Paul Wenthold

Hilkka I. Kenttämaa

Yu Xia

Paul Wenthold

Timothy Zwier

Timothy Zwier 04/15/2016

Page 3: Mass spectrometric characterization of remotely charged
Page 4: Mass spectrometric characterization of remotely charged

MASS SPECTROMETRIC CHARACTERIZATION OF REMOTELY CHARGED

AMINO ACIDS AND PEPTIDES

A Dissertation

Submitted to the Faculty

of

Purdue University

by

Damodar Koirala

In Partial Fulfillment of the

Requirements of the Degree

of

Doctor of Philosophy

August 2016

Purdue University

West Lafayette, Indiana

Page 5: Mass spectrometric characterization of remotely charged

ii

To Koirala Family

Page 6: Mass spectrometric characterization of remotely charged

iii

ACKNOWLEDGMENTS

It certainly feels great to have the opportunity to thank individual or group that

inspired, helped, supported and cared me to be where I am today. I am thankful to God

for sending all these amazing people to be part of my life.

I want to begin thanking my parents for the quality education and their supported.

Being the youngest in the family, my brother and sister always made sure I am not losing

track on my passion. I am also thankful to my relatives who have helped our parents

when I am away from them.

I would like to thank my friends and advisors at University of Wisconsin-

Superior. The parties and walk to the bars with friends are always hinging on my mind.

The food and gathering at Tom Markee’s house was always great. I am very thankful for

Tom for being my mentor, you are a great teacher. I still can’t believe your wife Sue is

not with us anymore, she has been truly missed. I would also like to thank Profs Waxman

and Lane for their inspiration on learning chemistry in class and in laboratory.

I feel blessed to have a wife who is with me on every twist and turns. She has

helped me throughout my Ph.D. career starting from OP. We are blessed to have a

wonderful daughter who is always there to refresh us and make us feel like we are in

different world.

Page 7: Mass spectrometric characterization of remotely charged

iv

The Nepali community at Purdue is a very helpful community. It consists of very

helpful and educated people. They make me feel like I have never been away from my

family. They are always there during the time of difficulties. We had lots of fun together,

lots of parties, sports, travelling, casinos and celebrations. I had pleasure serving as a

treasurer (2014-15) and a president (2015-16) of Nepali Society at Purdue.

I am thankful for the colleagues at Department Chemistry at Purdue. Class of

2011 was great, we might have more parties and fun than any other group we had our

own club ‘Club 1106’. Bharat and Manish might be ones who were closer to me during

my stay at Purdue. Bharat and I shared lots of moment together from good to bad. We

prepared together for cumes, analytical seminar and OP. Manish and I also shared lots of

moment together. I spend most of first year at the apartment that he and Janak shared.

And we all moved to next apartment and lived together during our second year at Purdue.

Those moments were best, good food, FIFA games and NBA double header on TNT. I

also want to thank Manish and Neha for helping me on dipeptide synthesis. In addition, I

want to thank Hari Khatri dai for his support and help in several chemical synthesis and

NMR analysis. Hari dai devoted his valuable time on analyzing my sample and also

setting up some synthesis in this lab hood. Thank you all.

I am grateful for the support I received from my major advisor, Prof Paul G

Wenthold, during my stay at Purdue University. I must say he is a good man with big

brain and bigger heart. He has great passion in Chemistry and science in general. I can

never forget the ideas, motivations and enthusiasm that Paul brought to one on one or

group meetings. I always had positive productivity after meeting with him. Thank you.

Page 8: Mass spectrometric characterization of remotely charged

v

I would like to thank the former Wenthold Lab members that I worked with. Dr.

Nathan J. Rau was my mentor who taught me how to use flowing afterglow. I was very

inspired by your knowledge on Chemistry and your dedication to work perfectly.

Besides, the time we spend together at ASMS conference in Vancouver. Do you

remember uphill from conference hall to the hotel, it was very steep, right? I did not

worked with Dr. Ekram Hossain but we spend lots of time together in lab and outside.

Ekram is a very intellectual person with lots of courage to try new stuff. I am sure you

will be big time professor one day. One moment I recall being with you is when Bharat

dropped you to the wrong airport in Indianapolis. Beside these two, I enjoyed working

with undergraduate Deon Turner. Actually, he synthesized the amino acid studied in

chapter 6 of this thesis, and did some mass spectrometry studies on that compound. I

wish him good luck for bright future.

Now, I would like to remember current Wenthold Lab members. I was very happy

when Chris Haskin joined our lab because I was by myself for long period of time. Chris

changed the structure of lab by cleaning and organizing old stuff. He is junior to me but I

am surprised how smart he is, even Paul would ask him on organic reactions, Wow. Babu

joined our lab within a month after Chris joined, so then we became triplet. My friends

teased us using MS terminologies “single-quadrupole changed to triple-quadrupole”.

Babu has very enthusiastic personality with diverse knowledge on the field of Chemistry.

He looks like a youngster with lots of energy but she always tells me that he is older than

me, which I have hard time believing. If I was not graduating soon then I am sure we

would have collaborated and have at least a co-authored peer reviewed article. I also

enjoyed being around and working together with undergraduate students; James

Page 9: Mass spectrometric characterization of remotely charged

vi

Langford, Bryan Smith and Chris Brown. They are all smart guys with lots and lots of

talent on them. They all know world history almost in detail. James is very good at

communicating and telling stories about his research. He is a person who I witnessed

struggled a lot at the beginning of his research career and improved day by day to be a

systematic researcher. I worked with Bryan only for a semester when we struggle a lot on

synthesis. When Chris Haskin joined the group, Bryan is having lots of success and has

become independent researcher as well. The meeting at the far was awesome. I helped

Chris Brown on several incidences but never worked together in a project. In my

understanding, Chris Brown had good knowledge on how researches are performed even

before he joined or started using our lab. So when free, we spend most of our time on

discussion and future goal. Overall, I enjoyed the company of current lab member.

Thank you all for your continuous support and believing on me. I will do my best

to keep all the promises.

Page 10: Mass spectrometric characterization of remotely charged

vii

TABLE OF CONTENTS

Page

LIST OF TABLES ...............................................................................................................x

LIST OF FIGURES ........................................................................................................... xi

ABSTRACT ..................................................................................................................... xiv

CHAPTER ONE: GUIDE TO DISSERTATION ................................................................1

1.1 Guide to Dissertation ...............................................................................................1 1.2 References ................................................................................................................3

CHAPTER TWO: REVIEW ON GAS-PHASE STRUCTURES OF AMINO ACIDS .....4

2.1 Introduction ..............................................................................................................4 2.2 Experimental Techniques.........................................................................................8 2.3 Computational Studies ...........................................................................................10 2.4 Conformations of AAs in Gas Phase .....................................................................11

2.4.1 Group 1: AA with R = H i.e. (Gly) ............................................................. 12 2.4.2 Group 2: AAs with Aliphatic Side Chains ....................................................16 2.4.3 Group 3: AAs with Oxygen or Sulfur on Side Chain ...................................21 2.4.4 Group 4: AAs with Amine Substituted Side Chain ..................................... 24 2.4.5 Group 5: AAs with Carbonyl Substituted Side Chain ................................. 25 2.4.6 Group 6: AAs with Aromatic Side Chain .................................................... 26

2.5 Stabilization of Zwitterionic AAs in Gas Phase ....................................................28 2.6 Conclusion .............................................................................................................31 2.7 References ..............................................................................................................32

CHAPTER THREE: EXPERIMENTAL APPROACHES ................................................37

3.1 Introduction ............................................................................................................37 3.2 The Flowing-Afterglow Triple-Quadrupole Mass Spectrometer ..........................38

3.2.1 The Ion Source ..............................................................................................38 3.2.2 The Mass Analyzer .......................................................................................40 3.2.3 Ion Detection Region ....................................................................................41

Page 11: Mass spectrometric characterization of remotely charged

viii

Page 3.3 The LCQ-Deca Mass Spectrometer .......................................................................42

3.3.1 The Ion Source ..............................................................................................42 3.3.2 The Mass Analyzer .......................................................................................43 3.3.3 Ion Detection Region ....................................................................................46

3.4 References ..............................................................................................................47

CHAPTER FOUR: THE FLOWING AFTERGLOW STUDY OF PYRIDINE N-OXIDE NITRENE RADICAL ANIONS ......................................................................48

4.1 Introduction ............................................................................................................48 4.2 Computational Detail .............................................................................................53 4.3 Results ....................................................................................................................53

4.3.1 Reactivity Studies .........................................................................................54 4.3.1.1 Reaction with NO .............................................................................56 4.3.1.2 Reaction with CS2 .............................................................................58 4.3.1.3 Comparison of Isomers: Oxygen Nucleophilicity ............................64

4.4 Conclusion .............................................................................................................67 4.5 References ..............................................................................................................68

CHAPTER FIVE: MASS SPECTROMETRIC STUDY OF THE DECOMPOSITION PATHWAYS OF CANONICAL AMINO ACIDS AND Α-LACTONES IN THE GAS PHASE ...............................................................................................................................70

5.1 Introduction ............................................................................................................70 5.2 Experimental ..........................................................................................................74 5.3 Synthesis ................................................................................................................75 5.3.1 (Z)-4-(3-ethoxy-3-oxoprop-1-en-1-yl)benzenesulfonate(deprotonated

ethyl 4-sulfo-cis-cinnamate) .................................................................................75 5.3.2 4-(acryloyloxy)benzenesulfonate (PASO3

) .................................................76 5.4 Computational Methods .........................................................................................76 5.5 Results ....................................................................................................................77 5.5.1 Dissociation of PheSO3

...............................................................................79 5.5.2 Dissociation of the α –Lactone, αLSO3

......................................................86 5.6 Discussion ..............................................................................................................88 5.6.1 Amino Acid Dissociation ..............................................................................90 5.6.2 Dissociation of the α –Lactone .....................................................................94 5.7 Conclusions ............................................................................................................99 5.8 References ............................................................................................................101

Page 12: Mass spectrometric characterization of remotely charged

ix

Page

CHAPTER SIX: EFFECT OF CHARGE POLARITY ON AMINO ACID DISSOCIATION..............................................................................................................107

6.1 Introduction ..........................................................................................................107 6.2 Synthesis ..............................................................................................................108 6.3 Results and Discussion ........................................................................................108

6.3.1 Dissociation of PheNMe3+ .........................................................................109

6.3.1.1 Neutral Loss of Mass 17 Da and Formation of m/z 206 ion ...........111 6.3.1.2 Neutral Loss of Mass 44 Da and Formation of m/z 179 ion ...........113 6.3.1.3 Neutral Loss of Mass 45 Da and Formation of m/z 178 ion ...........118 6.3.1.4 Neutral Loss of Mass 74 Da and Formation of m/z 149 ion ...........118 6.3.1.5 Neutral Loss of Mass 87 Da and Formation of m/z 136 ion ...........118

6.4 Conclusion ...........................................................................................................119 6.5 References ............................................................................................................124

CHAPTER SEVEN: MOBILE C-H PROTONS IN A PROTON DEFICIENT PEPTIDE .......................................................................................................................126

7.1 Introduction ..........................................................................................................126 7.2 Experimental ........................................................................................................130 7.3 Synthesis ..............................................................................................................131 7.3.1 Synthesis of Amino Acid Esters .................................................................131 7.3.2 Synthesis of Gly-Phe*OH and Gly-Phe*OMe Dipeptides .......................131 7.3.3 Synthesis of Phe*-GlyOH and Phe*-GlyOMe Dipeptides .......................132 7.3.4 Synthesis of Diketopiperazine ....................................................................132 7.3.5 Synthesis of Gly-Phe* Oxazolone ..............................................................133 7.4 Results and Discussion ........................................................................................134

7.4.1 Analysis of Gly-Phe*OH and Phe*-GlyOH .............................................134 7.4.2 Analysis of Gly-Phe*OMe and Phe*-GlyOMe ........................................136 7.4.3 Deuterium Labelled Phe*-GlyOMe ...........................................................142 7.4.4 Computational Results ................................................................................146

7.5 Conclusion ...........................................................................................................148 7.6 References ............................................................................................................149

VITA ................................................................................................................................152

PUBLICATIONS .............................................................................................................153

Page 13: Mass spectrometric characterization of remotely charged

x

LIST OF TABLES

Table Page

2.1Significance of 20 α-amino acids .............................................................................. 6

4.1Reaction Efficiencies and Product Branching Ratios for Reactions of 3PNO and 4PNO with NO and CS2 .................................................................... 56

5.1Computed Reaction Energies for Dissociation Pathways of PheSO3 and

Phe .......................................................................................................................... 89

7.1Calculated Relative Energies of b2 Ions ................................................................ 147

Page 14: Mass spectrometric characterization of remotely charged

xi

LIST OF FIGURES

Figure Page

2.1 Structure of naturally occurring 20 α-amino acids ................................................... 5

2.2 The 4 types of possible intramolecular hydrogen bond in AAs ............................. 11

3.1 The flowing-afterglow triple-quadrupole mass spectrometer ................................ 38

4.1 Resonance effect due to N-Oxide motif (Left) and the resonance structures of the π2α and α2π states of 4PNO (Right) ............................................................ 52

4.2 The m/z 92 region of the mass spectra of 4PNO before (dashed) and after (solid) addition of CS2 ............................................................................................ 60

4.3 The m/z 92 region of the mass spectra of 4PNO reaction with CS2 (dashed) and after addition of NO (solid) ............................................................... 61

4.4 Calculated Mullikin charge distributions in 3PNO , 4PNO and pyridine-n-oxide. Charges on hydrogen have been summed into the heavy atoms. ............ 66

5.1 a) ESI and b) CID mass spectra of PheSO3. The CID for spectrum was carried out with a normalized collision energy of 25%. ......................................... 78

5.2 Low-energy conformations of the amino acid moiety of PheSO3 and their relative energies, in kcal/mol, computed at the B3LYP/6-31+G* and MP2/6-311+G** (in parentheses) level of theory .................................................. 79

5.3 CID spectra of m/z 227 ions obtained a) by CID of PheSO3 , and from sulfonated b) cis and c) trans cinnamic acid ........................................................... 83

5.4 Calculated energetics for the dissociation of PheSO3 (X = SO3) and Phe

(X = H). Values correspond to electronic energies and geometries calculated at the MP2/6-311+G**//B3LYP/6-31+G* level of theory a) not a stable geometry; corresponds to an amino acid geometry the same as that for zwitterionic PheSO3; b) computed utilizing a 3 kcal/mol barrier for the reverse reaction, obtained from a relaxed surface scan; see Supporting Information. ............................................................................................................ 91

Page 15: Mass spectrometric characterization of remotely charged

xii

Figure Page

5.5 Calculated energetics for the dissociation of αLSO3 (X = SO3

) and the non-sulfonated derivative (X = H). Values correspond to electronic energies and geometries calculated at the MP2/6-311+G**//B3LYP/6-31+G* level of theory, unless indicated a) not a stable geometry; corresponds to the geometry of the lactone with the CO2 removed ....................... 96

6.1 The 7 low energy conformations of PheNMe3+. Relative energies are

shown in kcal/mol ................................................................................................ 109

6.2 CID spectra of a) PheNMe3+ with m/z 223, b) d3-PheNMe3

+ with m/z 226 ....... 110

6.3 CID spectra of m/z 206 obtained from a) CID of PheNMe3+, b) authentic

cinnamic acid (1) .................................................................................................. 112

6.4 CID spectra of a) m/z 179 obtained from CID of PheNMe3+, b) 2 with m/z

179, c) d2-2 with m/z 181 ..................................................................................... 114

6.5 CID spectra of m/z 164 obtained from a) 2, b) PheNMe3+ .................................. 115

6.6 PES for decarboxylation of PheNMe3+, PheSO3

and Phe. The barriers and the molecular energy were were calculated at the MP2/6-311+G**//B3LYP/6-31+G* level of theory. ........................................................ 117

6.7 Proposed Fragmentation Pathways of PheNMe3+................................................ 120

6.8 Possible rearrangement products of PheNMe3+ and their relative energies

in kcal/mol calculated at the B3LYP/6-31+G* level of theory .......................... 121

6.9 CID spectrum of Ia with m/z 223 ........................................................................ 123

7.1 Formation of diketopiperazine from dipeptide with cis amide bond (top) and formation of oxazolone from dipeptide with trans amide bond (bottom) ..... 128

7.2 Mass Spectra: a) CID of Gly-Phe*OH and b) CID of Phe-Gly*OH ................. 135

7.3Proposed mechanism for formation of common structure from Phe*-GlyOH and Gly-Phe*OH .......................................................................... 136

7.4 Mass Spectra: a) CID of Gly-Phe*OMe and b) CID of Phe*-GlyOMe ............. 138

7.5 Mass Spectra: CID of b2 ions a) MS3 spectra of Gly-Phe*OMe and b) MS3 spectra of Phe*-GlyOMe c) CID of 1 d) CID of 2 ................................. 140

7.6 Mass Spectrum: CID of Phe*-Gly (1-13C) OMe .................................................. 142

Page 16: Mass spectrometric characterization of remotely charged

xiii

Figure Page

7.7 Mass Spectra: CID of deuterated sample a) d3-Phe*-GlyOMe b) Phe*-Gly (2, 2-d2) OMe c) d5- Phe*-Gly (2, 2-d2) OMe ............................ 145

7.8 Proposed fragmentation pathway of Phe*-GlyOR’ ............................................ 146

7.9 Lowest energy structures of 1, 3, 4, 2, and 5 respectively ................................... 148

Page 17: Mass spectrometric characterization of remotely charged

xiv

ABSTRACT

Koirala, Damodar. Ph. D., Purdue University, August 2016. Mass Spectrometric Characterization of Remotely Charged Amino Acids and Peptides. Major Professor: Paul G. Wenthold. Ion–molecule reactions in a flowing afterglow are used to examine the

electronic structure of 3- and 4-pyridinylnitrene-n-oxide radical anions. Reactions with

nitric oxide are generally similar to those reported previously for other aromatic

nitrene radical anions. In particular, phenoxide formation by nitrogen–oxygen

exchange is observed with both isomers. Oxygen atom abstraction by NO is also

observed with both isomers. Very significant differences in the reactivity are observed

in the reactions of the two isomers with carbon disulfide. The reactivity of the 3-n-

oxide isomer with CS2 is similar to that observed previously for nitrene radical anions,

and reactions of the n-oxide moiety are not observed, similar to what is expected based

on solution chemistry. The 4-n-oxide isomer, however, undergoes many reactions,

including oxygen atom and oxygen ion transfer and sulfur–oxygen exchange, that

involve the n-oxide oxygen. The increased reactivity of the oxygen is attributed to

increased charge density at the oxygen due to pi electron donation of the nitrene anion

in the para position.

The dissociation pathways of gas-phase amino acids with a canonical (non-

zwitterionic) α-amino acid moiety are studied by using mass spectrometry.

Page 18: Mass spectrometric characterization of remotely charged

xv

Investigation of the canonical amino acid moiety is possible because the ionized

amino acid has a charge center that is separated from the amino acid, and dissociation

occurs by charge-remote fragmentation. The negatively charged amino acid is found

to dissociate only by loss of NH3 upon collision-induced dissociation to form a

substituted α-lactone. The collision-induced dissociation spectrum of the positively

charged amino acid is complex with possibly 5 different fragmentation pathways. The

results on negatively charged amino acid, para-sulfonated phenylalanine (Phe*), show

that remote ionic groups can be used as mostly inert charge carriers to enable mass

spectrometry to be used to investigate the gas-phase physical and chemical properties

of different types of functional groups, including amino acids.

The b2 ion formed upon charge remote fragmentation of dipeptides, Phe*-

GlyOH and Gly-Phe*OH, is characterized using LCQ-Deca mass spectrometry.

Comparison with authentic samples confirmed the lack of diketopiperazine or

oxazolone on b2 ion and electronic structure calculation suggested that the b2 ion is

oxazolone-enol.

Page 19: Mass spectrometric characterization of remotely charged

1  

CHAPTER ONE: GUIDE TO DISSERTATION

1.1 Guide to Dissertation

Mass Spectrometer (MS) is a quantitative tool used to measure the mass-to-charge

ratio of ions. For more than a decades, our group has used a home-built flowing-

afterglow triple-quadrupole MS1,2 to characterize the electronic structure of nitrene

anions.3 Nitrenes are a fascinating because although isoelectronic to carbenes, they show

very different reactivity.4-6 Recently, our group has started to use a commercial LCQ-

Deca MS7 to investigate the charge remote fragmentation8,9 of amino acids and dipeptides.

10

This thesis describes the gas-phase characterization of pyridine n-oxide nitrene,

and charge-remote fragmentation of amino acids and dipeptides using mass spectrometers

(MS) as an analytical tool. Chapter 2 features a literature review on the gas-phase

structure of the α-amino acids. It addresses the interactions present on low energy

conformation of gaseous amino acids and how the different side chains on amino acids

affect those interactions.

All the experimental results in this thesis were obtained from a home-built

flowing-afterglow triple-quadrupole MS,2 or a commercial LCQ-Deca MS.7 Chapter 3

has the detail explanation on different components of these instruments.

Page 20: Mass spectrometric characterization of remotely charged

2  

In Chapter 4, we report the investigation on reactivity of 3- and 4-

pyridinylnitrene-n-oxide radical anions with O2, CO2, NO and CS2. We discovered that

these isomer react differently only with CS2 where N-oxide of 4-PNO. participates to

yield CS2O- product. The dissimilarity in reactivity is attributed to the resonance structure

of 4-PNO. where charge is in N-Oxide, which is not possible in 3-PNO. isomer.

Chapters 5 and 6 describe the positive and negative charge-remote fragmentation

of phenylalanine, PheSO3 and PheNMe3+, respectively. Techniques like isotope

labelling, comparison with authentic sample and electronic structure calculations are

implemented to characterize fragmentation ions and neutrals. We discovered PheSO3

has a single fragmentation pathway whereas PheNMe3+ has multiple fragmentation

pathways. The discrepancy in dissociation of PheSO3 and PheNMe3+ has been attributed

to the possibility on rearrangements of PheNMe3+ prior to dissociation.

Chapter 7 focuses on the characterization of the b2 ions formed upon dissociation

of dipeptide containing para-sulfonated phenylalanine and glycine. We discovered these

dipeptides rearrange to common structure before fragmentation. Comparison with

authentic samples confirmed the lack of diketopiperazine or oxazolone on b2 ion and

electronic structure calculation suggested that the b2 ion is oxazolone-enol.

Page 21: Mass spectrometric characterization of remotely charged

3  

1.2 References

  (1) Marinelli, P.; Paulino, J.; Sunderlin, L.; Wenthold, P.; Poutsma, J.; Squires, R. Int. J. Mass Spectrom. Ion Processes 1994, 130, 89.

(2) Graul, S.; Squires, R. Mass Spectrom. Rev. 1988, 7, 263.

(3) Rau, N.; Welles, E.; Wenthold, P. J. Am. Chem. Soc. 2013, 135, 683.

(4) Platz, M. Acc. Chem. Res. 1995, 28, 487.

(5) Horner, L.; Christma.A . Angew. Chem. Int. Ed. 1963, 75, 707.

(6) Belloli, R. J. Chem. Educ. 1971, 48, 422.

(7) Thermo Electron Corporation: San Jose, CA.

(8) Cheng, C.; Gross, M. Mass Spectrom. Rev. 2000, 19, 398.

(9) Adams, J. Mass Spectrom. Rev. 1990, 9, 141.

(10) Koirala, D.; Kodithuwakkuge, S.; Wenthold, P. J. Phys. Org. Chem. 2015, 28, 635.

Page 22: Mass spectrometric characterization of remotely charged

4  

  

CHAPTER TWO: REVIEW ON GAS-PHASE STRUCTURES OF AMINO ACIDS

2.1 Introduction

Amino acids constitute a class of organic molecules containing carboxylic acid (-

COOH) and amine termini (-NH2). One way to classify amino acids is according to the

location of core structural functional groups such as alpha- (α-), beta- (β-), gamma- (γ-)

or delta- (δ-) amino acids. The α-amino acids are of particular interest because they are

proteinogenic, meaning they constitute the building blocks of peptides and proteins.

Although amino acids are abundantly found in nature, proteins are constructed from a set

of 20 naturally occurring α-amino acids, Figure 2.1. The generic formula for an α-amino

acid is RCH(NH2)COOH where the amine group (-NH2) and variable side chain ( R ) are

attached to the α-carbon of the carboxylic acid group (-COOH). Proteinogenic amino

acids are indispensable agents of biological function since some proteins functions as

enzymes, some as antibodies, and others provide structural supports. In addition, these

amino acids may have non-protein functions as neuro-transmitters or as precursors of

important neuro-transmitters or hormones, Table 1.1. For these reasons, the intrinsic

properties of the 20 naturally occurring α-amino acids have attracted the interest of a

number of researchers over the past several decades. For simplicity, α-amino acid will

here with be referred to as AA.

Page 23: Mass spectrometric characterization of remotely charged

5  

  

 

 

Figure 2.1. Structure of naturally occurring 20 α-amino acids

 

Page 24: Mass spectrometric characterization of remotely charged

6  

  

Table 2.1: Significance of 20 α-amino acids

AA Significance

Gly Essential for the synthesis of nucleic acids, bile acids, and porphyrins.

Ala Involves in the metabolism of the vitamin pyridoxine.

Val Incorporate into poteins and enzymes to help dictate the tertiary structure of the macromolecules. Leu

Ile

Pro Acts as a major component of the protein collagen and the connective tissue.

Ser Play an important role in porphyrins metabolism. Thr

Cys Metabolisms of coenzyme A, heparin, biotin and lipoic acid.

Met Initiates translation of messenger RNA.

Lys Used at the active sites of enzymes. Arg

Asn

Present outside proteins and enzymes interacting with the surroundings. Asp Glu Gln

His Responsible for the biosynthesis of neurotransmitter namely histamine.

Phe Play key roles in the biosynthesis of neurotransmitters. Trp

Tyr Synthesis of thyroid hormones and melanin biological pigments.

The function of proteins and their multidimensional structures are highly

dependent upon the conformation of their constituent AAs. Therefore, the knowledge of

the shape and the conformation of AAs is of great interest in a variety of fields, ranging

from biology and chemistry to pharmacology and molecular modeling. AAs have been

Page 25: Mass spectrometric characterization of remotely charged

7  

  

studied extensively in the condensed phase 1-4 and in the solution 5-7 but less so in the gas

phase.8,9 AAs exist in equilibrium between a doubly-charged species of zwitterions (Z),

RCH(NH3+)COO-, and neutral canonical forms (N), RCH(NH2)COOH, Scheme 2.1.8,10-

13 They are predominately present as zwitterions in condensed phase due to the strong

electrostatic, intermolecular, and polarizing interactions with the environment. However,

the lack of these intermolecular interactions favors canonical forms of AAs in gas phase

8,9,13. Since studies in condensed phases are affected by intermolecular interactions,

conformational preferences can be biased by the environment.14 The advantage of gas-

phase studies is that the gas-phase data can be contrasted with the theoretical models and

the molecules can be studied in detail with an optimized interplay between experiment

and theory.15

Gas phase study of isolated AAs creates an opportunity to provide information on

the neutral canonical forms of AAs present in peptide side chains. In addition, the

inherent molecular properties are provided, free of the intermolecular interactions

occurring in the condensed phase where AAs are bipolar zwitterionic species. Although

the intrinsic properties of isolated AA are of great interest, understanding the role of

solvent is also equally important since biological molecules including proteins and AAs

Scheme 2.1

Page 26: Mass spectrometric characterization of remotely charged

8  

  

exist in solution in living organisms.16 Hence, AAs are studied in gas-phase either in

isolated form or with a controlled number of solvent molecules. In particular, because

AA-AA intermolecular interaction can be neglected, the effect of solvation on the

configuration of an AA can be compared between experiment and theory. Most of the

experimental and theoretical studies of these types are designed to determine AA

configuration in the presence of a known number of solvent molecules.8,17-21 The latest

review by Bouchoux10 covers the results regarding neutral AA configuration and basicity

in the gas phase. This review summarizes the experimental and theoretical findings on

low energy conformers of canonical AAs, as well as addressing the chemical

environment required for stability of zwitterionic structures.

2.2 Experimental Techniques

Several techniques have been implemented to study the conformations of

canonical AAs is gas phase. Samples are first sublimed into gas phase and then analyzed

using spectroscopic methods.

Conventionally, solid sample on a uniformly heated stainless steel surface parallel

to a high vacuum chamber where desired analysis can be performed. This heating method

has presented considerable difficulties on creating gaseous AAs due to the high melting

points and associated low vapor pressures of AAs. Nonetheless, conventional heating

method has been employed for the analysis of gaseous Gly11,22-24, Ala11,22,Val 11,25, Cys22,

Thr22, Met22, Pro11,25, Phe25, Ile25 and Leu25.

An alternative to conventional heating is laser ablation. In this method, the fast

desorption produced by the energy of a laser pulse prevents thermal decomposition

Page 27: Mass spectrometric characterization of remotely charged

9  

  

normally caused by conventional heating. An Nd-YAG laser is used to ablate solid

samples, which are then seeded in the carrier gas and are expanded supersonically in an

analysis chamber.

Until the introduction of laser-ablation techniques, a gas-phase spectrum was only

available for solids with sufficient vapor pressure and thermal stability. Hence, there was

a lack of information on the structures of most AAs and microsolvation clusters due to

the vaporization difficulties associated with the aforementioned high melting points and

thermal instability. Several AAs like Gly, Ala26, Val16, Ile27, Leu28, Ser14, Cys29, Phe30,

Pro31,32, Thr14, Asp33 ,Asn30, Glu34, His35,Trp36, have now been studied in gas phase by

this means.

Several spectroscopic methods have been implemented to study gaseous

aminoacids. Microwave spectroscopy has been implemented Gly, Ala22,26, Val16, Ile27,

Leu28, Ser14, Cys29, Phe30, Pro31,32, Thr14, Asp33, Asn30, Glu34, His35 and Trp36 in gas

phase. Infrared spectroscopy has been used to study AAs like Gly22, Ala22, Thr22, Cys22,

Met22, Val25, Pro25, Phe25, Ile25 and Leu25. Raman spectroscopy 37,38 and electron

diffraction spectroscopy 39 also are implemented to study amino acids in gas phase.

Page 28: Mass spectrometric characterization of remotely charged

10  

  

2.3 Computational Studies

Each experimental study is conducted along with a complementary

computational study. Computational studies are used to find the low energy conformers

of the AA and the desired parameter associated with them. For example, in case of

microwave spectroscopy, the rotational constants for several low energy conformers are

predicted computationally and compared with experimental results. Computational

studies assist on qualitative and quantitative analysis of conformations present in a

gaseous sample of the AA. 23,38,40,41

Although computational studies help to interpret the experimental findings, it can

easily mislead to a false conclusion. For instance, the results for gas-phase study of

canonical Met22,42 were interpreted based on the conformations that were later discovered

to be more than 10 kJ/mol above the most stable conformation. So the conclusions drawn

at that time are now invalid10.

On the other hand, a false conclusion could be drawn due to false or insufficient

experimental outcomes. For instance, the rotational constant obtained from the

microwave spectroscopy of Gly matched with the second most stable conformer of Gly

23,43 suggesting this is the stable conformation of Gly is gas phase. However, it was later

discovered that it was not possible to ascertain which conformer was the main one from

that investigation because the intensities of the microwave spectra strongly depend on the

magnitude of the dipole moment of the molecule and not on the abundance of the

conformer in the gaseous sample 12,44. But with some modifications to the instrument, a

convincing result between experimental and theoretical studies of gaseous Gly

Page 29: Mass spectrometric characterization of remotely charged

11  

  

conformers was drawn.45 Therefore, systematic experimental and theoretical studies are

essential to make a conclusive decision on preferred conformation of gaseous AAs.

2.4 Conformations of AAs in Gas Phase

AAs are very flexible molecules, and their conformational landscape is

characterized by a delicate balance of covalent and non-covalent interactions, which may

result in a large number of low-energy conformers.26 The most important interaction

present in gaseous AA is intramolecular hydrogen bonding. There are four common

intramolecular hydrogen bonding present in AA, Figure 2.2.10,16 They are: 1) between the

acidic hydrogen and the carbonyl oxygen, 2) between acidic hydrogen and neighboring

basic site, 3) between N-H hydrogen and carbonyl oxygen, and 4) N-H hydrogen and

hydroxy oxygen as shown in Figure 2.2. The combinations of these interactions create

three unique interactions namely types I, II and III as shown in Figure 2.2. These

notations are used throughout this review for classification of interaction present in AA.

The following section describes the structure of AAs based on the side chain.

Figure 2.2. The 4 kinds of possible intramolecular hydrogen bond in AAs

Page 30: Mass spectrometric characterization of remotely charged

12  

  

2.4.1 Group 1: AA with R = H i.e. (Gly)

Gly is the simplest of all AA where the side chain is hydrogen. Due to its

simplicity, the structure of Gly has been studied extensively to build a general framework

on understanding the structure of other complicated AAs.11,12,19,23,24,42,44-54 Gly has three

internal rotational degrees of freedom: the rotation of the hydroxy group around the C1-O

bond (α), the rotation around the Cα-C1 bond (β), and the rotation around Cα-N bond (γ)

as shown in Scheme 2.2.

The interactions shown in Figure 2.2 can be attributed to the rotation around C1-O

and Cα-C1 bonds. For C1-O bond rotation, conformations with O=C-O-H dihedral angles

closer to 0° ( also called cis) or 180° (also called trans) are significant since they favor

type 1 and type 2 hydrogen bonding shown in Figure 2.2, respectively. For Cα-C1 bond

rotation, conformations with N- Cα-C1-O dihedral angles closer to 0° or 180° are

significant since they favor type 4 and type 3 hydrogen bonding shown in Figure 2.2,

respectively.

The rotation around N-Cα determines the number of NH hydrogen involved in

type 3 or type 4 hydrogen bonding, which is defined by < Lp-N-Cα-C1 dihedral angle,

where Lp is the lone pair of electrons in nitrogen. When this dihedral angle is 180° (anti

or a), both hydrogens of amine can hydrogen bond whereas when this dihedral angle is

Scheme 2.2

Page 31: Mass spectrometric characterization of remotely charged

13  

  

+60° (gauche+ or g+) or -60° (gauche- or g-) only one hydrogen of amine can hydrogen

bond, Scheme 2.3.

The five most stable conformations of Gly are shown in Scheme 2.4.

9,12,41,43,44,46,49,55 All calculations predicts conformation Ia, conformation with type I

(Figure 2.2) hydrogen bonding and anti orientation of < Lp-N-Cα-C1 (Scheme 2.6), as the

most stable form by ~0.5kcal/mol than the next closest conformation. However, the

stability of the other conformations depends on the level of theory and the basis set used

in the calculations. The comparison of the relative energies calculated at the HF and MP2

levels shows the former method overestimates the energy gap between GlyIa and the

other conformations. DFT calculations of the structure and relative stabilities of the Gly

conformers produced results very similar to the results of the MP2 method.

Scheme 2.3

Page 32: Mass spectrometric characterization of remotely charged

14  

  

Scheme 2.4

Types I and III conformations prefer anti (a) configuration of Lp-N-Cα-C1 since

this configuration allows a bifurcated hydrogen bonding between amine and carboxylic

acid. However, type II conformation prefers gauche (g) orientation of Lp-N-Cα-C1 since

this configuration allows the hydrogen bonding between acidic hydrogen and basic amine.

Although the most stable conformation of Gly predicted computationally has been

consistent, the experimentally determined conformation has varied. Brown et al.23 and

Suenram & Lovas43 are first two groups to independently report similar findings on the

experimental results of the conformation of gaseous Gly in 1978. They both detected the

microwave spectrum of gaseous Gly in a specially constructed cell maintained at a

temperature at which Gly has an appreciable vapor pressure (160-200 °C) while

minimizing the effects of decomposition.23,43 Their findings suggest GlyIIg is the most

abundant conformation, which is contradictory to the computational findings.

Furthermore, they were not able to identify any lines attributing to the lowest energy

conformation, GlyIa.

Suenram & Lovas45 further investigated gaseous Gly with a number of

improvements to their spectrometer, which provided improved sensitivity. In this study,

Page 33: Mass spectrometric characterization of remotely charged

15  

  

they reported the good agreement on the experimental and theoretical rotational constants

for presence of GlyIa and GlyIIg. In addition, GlyIa was detected to be lower in energy,

by 490 cm-1, when compared to GlyIIg.

Although GlyIIg and GlyIIIa are predicted to have similar energies, there is no

sign on the presence of GlyIIIa in microwave spectroscopy experiment. Godfrey and co-

workers41 stated that GlyIIIa species relaxes to the lower energy GlyIa due to rotation

around C-O bond, augmenting the total concentration of GlyIa in those studies. This can

be understood as the relaxation of GlyIIIa to lowest energy conformer, GlyIa, via a low-

barrier pathway. Since no such low-barrier pathway is available for the relaxation of

GlyIIg, this conformation is observed in these experiments 41,55.

The joint analysis of electron diffraction data and rotational constants performed

by Iijima et. al12 concluded GlyIa to be the main (~76 %) conformer and the rest being a

mixture of GlyIIg and GlyIIIa, although there is no substantial experimental evidence.

Similarly, experiments performed using other techniques have claimed the observation of

GlyIIg and GlyIIIa along with GlyIa10. In addition, Balabin38 recently reported the

presence of new conformer, GlyIg, along with GlyIa and GlyIIe, in an experiment

performed with jet-cooled, high sensitive, Raman spectroscopy. This is the only

experimental incident where the presence of GlyIg is reported.

In summary, 75% of gaseous Gly exists as Ia and the rest as a mixture of IIg, IIIa,

and Ig while experiment is performed in between 200-300K. Almost 100% of Gly will

exist as the main conformer GlyIa at extremely low temperatures, as in cool interstellar

space.12

Page 34: Mass spectrometric characterization of remotely charged

16  

  

2.4.2 Group 2: AA with Aliphatic Side Chains

The five AAs in this category are Ala, Val, Leu, Ile and Pro. As name suggests,

AAs in this category have a hydrophobic aliphatic side chain where R contain just

carbon(s) and hydrogens. As in Gly, the conformational behavior of these AA can be

expected to be primarily governed by the intramolecular hydrogen bonding within the

AA backbone. However, the increase on size of the side chain complicates the

conformational landscape of the molecule considerably, resulting in a greater number of

plausible conformers27. For example, besides the free rotation discussed in Gly, rotation

around other C-C (for example Cβ- Cα) bonds would generate additional conformers

(Scheme 2.5).

Ala is the most studied AA in this group, which is mainly due to its simple

structure with R=Me. The five most stable conformations of free Ala determined

computationally are shown in Scheme 2.6.50,56-59 The experimental studies performed on

gaseous Ala confirmed the presence of only two conformers, Ala1 and Ala2.26,39-41,60,61

The failure to observe other low-energy conformers given by electronic structure

computations is not surprising since similar results were observed for Gly, which is

attributed to vibrational relaxation.

Scheme 2.5

Page 35: Mass spectrometric characterization of remotely charged

17  

  

Although the methyl group of Ala has no effect on the lower energy configuration

when compared to Gly, the presence of a bulky isopropyl group in Val (R = CH (CH3)2)

could generate a much steric demand, which could affect the conformation of the AA

back bone of the molecule.16 Theoretical calculations predict three unstrained staggered

orientations of the isopropyl group which are unique structures with different energies.

Considering rotations around Cβ-Cα bond, three unstrained staggered rotamers are

present in Val (Scheme 2.5).

Page 36: Mass spectrometric characterization of remotely charged

18  

  

The five most stable conformers of a free Val determined computationally are

shown in Scheme 2.9.10,16 The conformational behavior of Val is in closest agreement

with smaller aliphatic neutral α-AAs like Gly and Ala. Affixing an isopropyl side chain to

the α-AA backbone does not alter the relative energies of the lowest energy conformers.

This observation indicates the relative stabilities of the AA skeleton are controlled by

Scheme 2.6

Page 37: Mass spectrometric characterization of remotely charged

19  

  

intramolecular hydrogen bonding.16 The experimental finding is also comparable to Gly

and Ala. Only Val1 and Val2 are observed in experiments, whereas the absence of other

conformers can be attributed to facile relaxation to lower energy conformers due to

rotation around the C-C bonds.16

As the chain length of side chain grows, the number of C-C bond rotations to

consider increases as well. In the case of Ile and Leu, configuration associated with Cγ-

Cβ rotations as well as rotations in Val need to be considered. Due to this, the number of

possibly lower energy conformers appears in the PES of Leu and Ile will increase by a

factor of 3 compared to Val. The five lowest energy conformers of Leu and Ileu

determined computationally are shown in Scheme 2.6 as rotation around Cγ-Cβ bond. As

for smaller aliphatic AAs, the conformational behavior of Ile and Leu is primarily

governed by the intramolecular hydrogen bonding within the AA backbone. However,

the side chain orientation for the most stable conformation predicted computationally is

the same in Val and Ile but some differences appear on Leu. This can be attributed to a

minimization of steric repulsions within the side chain and the AA backbone since both

Val and Ile have two aliphatic groups connected to the beta carbon.28

Type I and type II configurations (Figure 2.2) are present in experimentally

studies of Leu and Ile with conformation with type I configuration being more stable than

that with type II configuration. This is consistent with the other aforementioned AAs. The

conformation with type I configuration is computationally predicted to be the lowest

energy conformation and is always the most abundant species in experimental results.

Moreover, the conformations with the type II configuration observed experimentally have

the same configuration due to rotation around Cβ-Cα and Cγ-Cβ bonds as for the lowest

Page 38: Mass spectrometric characterization of remotely charged

20  

  

energy conformation with type I configuration. Even though experimental and

computational results show these similarities for AA’s with aliphatic side chains, it is

different in the case for Leu. Leu2 conformation is computationally predicted to be the

most stable conformer with type II configuration, but Leu3 is observed experimentally. It

is unclear why Leu3 is observed despite its higher energy, but it is clear that in

experimental studies, the same configuration (due to rotation around Cβ-Cα and Cγ-Cβ

bond) as in lowest energy conformation with type I configuration is preferred for

conformation with type II configuration.27,28

Pro is the only α-AA containing a secondary amino group as part of a flexible

five-membered ring. Pro is predicted to have lower energy when the five membered

pyrrolidine ring adopts an envelope conformation.31,42 The envelope conformation of the

ring can be endo-like (en) and exo-like (ex) with respect to the carboxy moiety. All three

types (I, II and III) of hydrogen bonding are possible in Pro but there is only one N-H

hydrogen involving in the formation of intramolecular hydrogen bond with lone pairs of

carboxylic acid oxygen’s. The combination of en and ex with I, II and III give total of six

possible conformations of Pro (only 5 of them is shown in Scheme 2.9). In fact, there are

six lowest energy conformations of Pro and all of these configurations can be

independently detected in gaseous Pro experiment. Unlike the type I configuration

observed in lowest energy conformation of aforementioned aliphatic AAs, Pro exhibits

type II configuration as primary interactions on controlling its stability. As previously

mentioned, the most stable conformer with type I configuration of other AAs has

bifurcated H-bonding of NH2 with O=C, which is not possible in Pro, but this does not

explain variation in inversion of stability of I and II, as in the case of Pro. A comparison

Page 39: Mass spectrometric characterization of remotely charged

21  

  

can be drawn between Gly and Pro. The N-methylated Gly and N,N-dimethylated Gly

have same conformation preference as in Gly. However, the findings with Pro can be

attributed to a consequence of the torsional restrictions imposed by the five-membered

pyrrolidine ring inverting the stability. 10,31,42,62,63

2.4.3 Group 3: AAs with Oxygen or Sulfur on Side Chain

The four AAs in this category are Ser, Thr, Cys and Met. As discussed in previous

sections, the conformations of AA with non-polar side chains is mainly controlled by the

stabilization due to intramolecular hydrogen bonding between amino and carboxylic acid

groups present in the AA’s backbone. However, due to the polar side chains, the AAs in

this group have a new set of interactions present in them. Hydrogen on S-H or O-H can

interact with lone pair of nitrogen or oxygen of the backbone (e.g. S-H…O-H or

S…H…O=C), except for Met because it lacks a hydrogen on the sulfur. In addition, the

hydrogen on the acid and amine can interact with the loan pair of sulfur or oxygen of side

chain (e.g. O-H...S or N-H…S). These interactions will dramatically increase the number

of low-energy conformers. For example, there exists 51 conformational minima in PES of

Ser.64

The five low energy conformers of Ser, Thr, Cys and Met are shown in Scheme

2.7. It can be noted the low-energy conformations are mainly controlled by the

intramolecular hydrogen bonds established between the three polar groups (COOH, NH2

and S/O). In the case of Ser and Thr, the additional H-bonding between hydroxyl H and

amino group stabilizes the type I configuration. The carboxylic acid group stabilizes type

II and type III configurations. In the case of Cys, both type I and type II configurations

Page 40: Mass spectrometric characterization of remotely charged

22  

  

with preferred hydrogen bonding between the carboxylic acid and thiol. In addition, the

most stable conformer in Ser, Thr and Met have type I configuration whereas Cys has

type II configuration.

Microwave spectroscopy studies have detected six conformers of Ser56,64, six

conformers of Cys29,56, and seven conformations of Thr14, which is significantly more

than observed for AA’s with aliphatic side chains. Furthermore, conformations with type

III configuration not detected for previously discussed AA’s have been observed for these

AAs. The lack of evidence of type III configurations in experimental studies have been

explained based on its relaxation to lower energy conformer of type I configuration due

to the low-energy barrier. However, the conformers with type III configuration observed

in case of Ser, Cys and Thr have the lowest energy in their family (compared to type I

configuration when every other dihedral angles are kept constant). In addition, the

relaxation from type III configuration to type I configuration has a higher energy barrier,

which hinders the relaxation. For example, the energy barrier for relaxation of Cys4 to

Cys1 is calculated to be approximately 700 cm-1. Hence, it is possible to detect

conformation with type III configuration for this group of AAs. The results obtained from

experimental studies on gaseous Met22,42,63 are on conformations situated more than 10

kJ/mol above the most stable conformers according to the predicted PES. Therefore, the

experimental detection of Met is avoided.10

Page 41: Mass spectrometric characterization of remotely charged

23

 

  

 

23

Sch

eme

2.7

Page 42: Mass spectrometric characterization of remotely charged

24  

  

2.4.4 Group 4: AAs with Amine Substituted Side Chain

The two AAs in this category are Lys and Arg. Due to their polarity, several low-

energy conformer of these AA can be expected since new intramolecular hydrogen bonds

are possible. In addition, the long flexible hydrocarbon chain can bend to favor these

interactions.

There are five types of hydrogen-bonding configurations present in these AA:

(a) NH2…OCOH…NH2, (b) COOH…NH2…NH2, (c) H2N…HOCO…H2N,

(d) COOH…NH2 and (e) HOCO…H2N. However, due to the free rotation around all C-C

bonds, only the first two types are present for the five low energy conformers of Lys and

Arg shown in Scheme 2.8. It can be noted the lowest energy conformer of both Lys and

Arg has type II configuration, where –COOH is in the trans orientation. The first three

Scheme 2.8

Page 43: Mass spectrometric characterization of remotely charged

25  

  

lowest energy conformers of Lys have NH2…OCOH…NH2 and 4th and 5th conformers

have COOH…NH2…NH2 hydrogen bonding configurations where first NH2 is from the

AA backbone and second amine is from the side chain in both cases. In case of Arg, the

1st and 4th conformers have H2N…HOCO…H2N. The 2nd, 3rd and 5th conformers have

NH2…OCOH…N where the first NH2 is from AA backbone and second one is from side

chain in both cases and N is a nitrogen lacking hydrogen.65

2.4.5 Group 5: AAs with Carbonyl Substituted Side Chain

The four AAs in this category are Asp, Asn, Gln and Glu. Unlike previous

sections where polar side chains would interact with the polar backbone, AAs in this

section have a capacity to interact and form H-bonding within the side chain, which

increases the possible number of low-energy conformers on the PES.

The five low energy conformations of Asp, Glu, Asn and Gln are shown in

Scheme 2.9. All conformations of Asp and Glu have cis orientation of the COOH side

chain due to a 20 kcal/mol stabilization energy compared to the trans orientation. In

addition, all conformations have type II configuration of AA backbone and additional

hydrogen bonding with the side chain or they have type I configuration of AA backbone

for further stability.33 Since Asn and Gln have amine side chains, one hydrogen of amine

NH2 side chain form H-bond with acid of AA backbone and another with C=O of amide

in all cases. Interestingly, the most stable conformation of Asn adopts a ring like

structures due to presence of continuous hydrogen bonding.

Page 44: Mass spectrometric characterization of remotely charged

26  

  

NH

H

H

OO

H

Asp

NH

H

H

OO

H

H H

HHNH

H

H

OOH

HH

O OO

H

O

H

O

O H

NH

H H

OOH

HH

O

HO

NH

H H

OO

H

H

HOO

H

H

H

H

H

N

OO

H

O

OH

H

H

H

HH

H

N

O

O

H

O H

O

H

H

H

H

HH

N

O

O

H

H

H

OO

H

H

H

H

H

N

O

O H

OO

H

H

H

Glu

H

H

H

H

N

O

O H

H

H

O

O

H

Asn

NH

H

H

OO

H

HH

O

N H

NH

H

H

OOH

HH

NH

O

H

NH

H H

OOH

HH

N

HOH

NH

H H

OO

H

H

N HO

H

H

H

NH

H H

OOH

H

H

N

HOH

H

H

H

H

N

OO

H

O

NH

H

H

H

H

HH

O

H

H

H

H

N H

ON

H

O

Gln

H

NH H

NH

O

OH

H

H

O

H

H

H

HH

H

NH

O

O H

H

O

NH

HH

HH

H

NH

O

O H

H

ON

HH

1 2 3 4 5

2.4.6 Group 6: AAs with Aromatic Side Chain

The four AAs in this category are Tyr, Trp, Phe and His. These aromatic AAs

have ultraviolet radiation absorptions with large extinction coefficients. This feature is

often used as an ultraviolet marker for detection of these AAs, either separately, or

incorporated into proteins and enzymes, via ultraviolet spectrophotometry.

Scheme 2.9

Page 45: Mass spectrometric characterization of remotely charged

27  

  

The presence of an electron rich aromatic ring in the side chain stabilizes the

structure by utilizing interactions between the polarized hydrogen and the pi-electron.

The five low energy conformers of this group are shown in Scheme 2.10.10,66-71 The most

stable structure has trans COOH with O-H … NH2 hydrogen bond, which is further

stabilized by interaction of NH2 hydrogen with the pi-system of the aromatic ring. This

stability is consistent for all AAs in this group.

The conformation of Phe is of great interest to our research group since we have

stated to study charge-remote fragmentation of Phe in MS.72 The formation preferences

of negatively and positively charged Phe, PheSO3 and PheNMe3+, are shown in Figures

5.1 and 6.2 of this thesis, respectively. The low energy conformations of PheSO3 are

similar to Phe where Type II conformation is preferred whereas the low energy

conformations of PheNMe3+ prefer Types I and III conformations.

This concludes the structure of canonical AAs in gas phase. The next section is on

stabilization of zwitterionic AA in gas phase.

Scheme 2.10

Page 46: Mass spectrometric characterization of remotely charged

28  

  

2.5 Stabilization of Zwitterionic AAs in Gas Phase

Zwiterionic isomers of AA have been proved to be unstable in the gas phase.

However, they can be stabilized in gas-phase clusters through solvation with water17,73,

methanol17, metal cation17,53, zeolite18, chloride17 and ammonium ion19. Because the

zwitterion formation is important for the structure, function and activity of biological

systems involving peptides, proteins and nucleic acids, knowing what conditions stabilize

the zwitterion is essential. Several theoretical and experimental studies are performed to

understand the effects of solvation on the relative stability of canonical and zwitterionic

conformers.

The recent theoretical study by Pathak17 predicts two solvent (water and

methanol) molecules or a single ion can sufficiently stabilize the zwitterionic form of all

α-AAs. A combination of experimental and theoretical study on Trp by Polfer et.al73

reported that water preferentially binds on the carboxylic acid group and the water

molecules strongly affects the relative energetics of different structural motifs.

One of the very first ab initio studies examining how the successive addition of

water molecule affects the chemical dynamic of ZN equilibrium of the AA Gly

performed by Jensen and Gordon.8 They studied the effect of solvation (by 1 and 2 water

molecules) on the transition state energy and enthalpy of ZN proton transfer (PT)

reactions. Even though the zwitterionic form of Gly does not exist in the gas phase,

calculations were performed on the isolated structure because it provides valuable

information about the intrinsic chemistry of this species necessary for a direct evaluation

of the solvent effect. They discovered 1, 2 and 3 unique zwitterionic minima in the PES

of bare, mono-hydrated and di-hydrated Gly, respectively.

Page 47: Mass spectrometric characterization of remotely charged

29  

  

Since the only interaction present in Z0 (Gly Zwitterion with 0 water molecule) is

intramolecular hydrogen bonding, the proton transfer has to occur internally. The PT

reaction, Z0 N0, is calculated to be exothermic by 17 kcal/mol and the transition state

for this reaction is only 0.02 kcal/mol higher than Z0, indicating Z0 readily collapses to

N0. The destabilization of Z in the gas-phase has been attributed to the shorter (1.55 Å)

intramolecular hydrogen bond distance in Z0 than in N0 (1.97 Å) and to the fact the

proton affinity of COO anion is higher (by 129 kcal/mol) than the NH2.

The mono- and di-hydrated Gly can have either intra or intermolecular hydrogen

bonding which can assist the PT reaction. The Z1 conformers have a water molecule

bridging the COO and NH3+ groups. Intramolecular and intermolecular PT reactions are

calculated to be exothermic by 11.8 kcal/mol and 10.4 kcal/mol, respectively, with no

barrier for both mechanisms. It is apparent one water molecule is not enough to stabilize

the zwitterionic form of Gly. However, for the PT reaction of Z2 N2 is calculated to be

exothermic by 6.3 kcal/mol with a positive barrier of 1.9 kcal/mol. Furthermore, Z2

structure with intermolecular hydrogen bonding, and not with only intramolecular H-

bond, is the stable one. Hence, the study concludes two water molecules can stabilize Gly

zwitterionic structure in the gas-phase and the electrostatic interaction (H-bonding)

between water and the AA is essential for the stability of this structure.

Several theoretical and experimental studies have been performed since then to

determine how many water molecules are necessary to stabilize the Z conformer of

AAs.8,21,51,59,65,73-75 The zwitterionic structures of Arg, Asp acid, Asp, Cys, Phe, Ser, Thr

and Tyr can be stabilized by as few as one water molecule.17 The common feature among

them, with the exception of Phe, is a polar side chain. Interestingly, the other AAs with

Page 48: Mass spectrometric characterization of remotely charged

30  

  

polar side chains such as Glu, Gln and Lys, and with nonpolar side chain such as Ala, Gly,

His, Ile, Leu, Met, Pro, Typ and Val require two molecules of water to stabilize the

zwitterionic structure.17 The zwitterionic forms are stabilized by the formation of

hydrogen bonding interactions with water molecules. In general, the negative end of the

solvent dipole orients toward the basic site of AA and the positive end towards the acid

site of AA.17

Both theoretical and experimental studies carried out to examine the relative

stability of the canonical and zwitterionic forms as a function of the number of

microsolvating water molecules indicated that the effects of microsolvating water

molecules depend highly on the structure of AAs. Although most AAs require 4-5 water

to stabilize their zwitterionic conformers molecules those with strongly basic side chains

such as Arg and Lys may need fewer water molecules. 74,76Val is at the opposite side to

these latter AAs in the effects of side chain.20,53 The isopropyl side chain is highly

nonpolar, and is neither acidic nor basic. Val is a normal AA in its behavior towards the

effects of microsolvating water molecules, in contrast to other AAs such as Arg and Lys

whose zwitterions were predicted to become stable under the influence of fewer number

of water molecules due to side chains of strong basicity.20

In addition, calculations for the Arg-H2O system predicts the zwitterionic Arg be

thermodynamically more stable than the canonical form in the gas phase under the

influence of a single water molecule because of the strongly basic guanidine side chain.74

Canonical conformers of Arg-H2O are found to isomerize to the zwitterionic forms via a

small barrier (∼6 kcal/mol).74 In contrast to other AAs for which more water molecules

Page 49: Mass spectrometric characterization of remotely charged

31  

  

are required to stabilize the zwitterion, in case of Arg three water molecules are enough

because of the strongly basic side chain.

2.6 Conclusion This chapter summarizes conformational landscape of the 20 α-amino acids in gas

phase. The intramolecular hydrogen bonding mainly stabilizes the conformation of

canonical amino acid in gas phase. Zwiterionic isomers of AA have been proved to be

unstable in the gas phase, which exist predominately in the condensed phase. However,

they can be stabilized in gas-phase clusters through solvation with water, methanol, metal

cation, zeolite, chloride and ammonium anion.

Page 50: Mass spectrometric characterization of remotely charged

32  

  

2.6 References (1) Albrecht, G.; Corey, R. J. Am. Chem. Soc. 1939, 61, 1087.

(2) Marsh, R. Acta Crystallogr. 1958, 11, 654.

(3) Levy, H.; Corey, R. J. Am. Chem. Soc. 1941, 63, 2095.

(4) Donohue, J. J. Am. Chem. Soc. 1950, 72, 949.

(5) Gaffney, J.; Pierce, R.; Friedman, L. J. Am. Chem. Soc. 1977, 99, 4293.

(6) Cohn, E.; McMeekin, T.; Edsall, J.; Blanchard, M. J. Am. Chem. Soc. 1934, 56, 784.

(7) Ellzy, M.; Jensen, J.; Hameka, H.; Kay, J. Spectrochim. Acta, Part A 2003, 59, 2619.

(8) Jensen, J.; Gordon, M. J. Am. Chem. Soc. 1995, 117, 8159.

(9) Jensen, J.; MS, G. J. Am. Chem. Soc. 1991, 113, 7917.

(10) Bouchoux, G. Mass Spectrom. Rev. 2012, 31, 391.

(11) Lago, A.; Coutinho, L.; Marinho, R.; de Brito, A.; de Souza, G. Chem. Phys. 2004, 307, 9.

(12) Iijima, K.; Tanaka, K.; Onuma, S. J. Mol. Struct. 1991, 246, 257.

(13) Jarrold, M. Ann. Rev.Phys. Chem. 2000, 51, 179.

(14) Alonso, J.; Perez, C.; Sanz, M.; Lopez, J.; Blanco, S. Phys. Chem. Chem. Phys. 2009, 11, 617.

(15) Tia, M.; de Miranda, B.; Daly, S.; Gaie-Levrel, F.; Garcia, G.; Nahon, L.; Powis, I. J. Phys. Chem. A 2014, 118, 2765.

(16) Lesarri, A.; Cocinero, E.; Lopez, J.; Alonso, J. Angew. Chem. Int. Ed. 2004, 43, 605.

(17) Pathak, A. Chem. Phys. Lett. 2014, 610, 345.

(18) Yang, G.; Zhou, L.; Liu, C. J. Phys. Chem. B 2009, 113, 10399.

(19) Wu, R.; McMahon, T. J. Am. Chem. Soc. 2008, 130, 3065.

(20) Kim, J.; Won, G.; Lee, S. Bull. Korean Chem. Soc. 2012, 33, 3797.

Page 51: Mass spectrometric characterization of remotely charged

33  

  

(21) Vaquero, V.; Sanz, M.; Mata, I.; Cabezas, C.; Lopez, J.; Alonso, J. J. Phys. Chem. A 2014, 118, 2584.

(22) Linder, R.; Seefeld, K.; Vavra, A.; Kleinermanns, K. Chem. Phys. Lett. 2008, 453, 1.

(23) Brown, R.; Godfrey, P.; Storey, J.; Bassez, M. J. Chem. Soc., Chem. Commun. 1978, 547.

(24) McGlone, S.; Elmes, P.; Brown, R.; Godfrey, P. J. Mol. Struct. 1999, 485, 225.

(25) Linder, R.; Nispel, M.; Haber, T.; Kleinermanns, K. Chem. Phys. Lett. 2005, 409, 260.

(26) Blanco, S.; Lesarri, A.; Lopez, J.; Alonso, J. J. Am. Chem. Soc. 2004, 126, 11675.

(27) Lesarri, A.; Sanchez, R.; Cocinero, E.; Lopez, J.; Alonso, J. J. Am. Chem. Soc. 2005, 127, 12952.

(28) Cocinero, E.; Lesarri, A.; Grabow, J.; Lopez, J.; Alonso, J. ChemPhysChem 2007, 8, 599.

(29) Sanz, M.; Blanco, S.; Lopez, J.; Alonso, J. Angew. Chem. Int. Ed. 2008, 47, 6216.

(30) Cabezas, C.; Varela, M.; Pena, I.; Mata, S.; Lopez, J.; Alonso, J. Chem. Commun. 2012, 48, 5934.

(31) Lesarri, A.; Mata, S.; Cocinero, E.; Blanco, S.; Lopez, J.; Alonso, J. Angew. Chem. Int. Ed. 2002, 41, 4673.

(32) Mata, S.; Vaquero, V.; Cabezas, C.; Pena, I.; Perez, C.; Lopez, J.; Alonso, J. Phys. Chem. Chem. Phys. 2009, 11, 4141.

(33) Sanz, M.; Lopez, J.; Alonso, J. Phys Chem Chem Phys 2010, 12, 3573.

(34) Pena, I.; Sanz, M.; Lopez, J.; Alonso, J. J. Am. Chem. Soc. 2012, 134, 2305.

(35) Bermudez, C.; Mata, S.; Cabezas, C.; Alonso, J. Angew. Chem. Int. Ed. 2014, 53, 11015.

(36) Sanz, M.; Cabezas, C.; Mata, S.; Alonso, J. J. Chem. Phys. 2014, 140.

(37) Balabin, R. J. Chem. Phys. Letters 2010, 1, 20.

Page 52: Mass spectrometric characterization of remotely charged

34  

  

(38) Balabin, R. Phys. Chem. Chem. Phys. 2010, 12, 5980.

(39) Jaeger, H.; Schaefer, H.; Demaison, J.; Csaszar, A.; Allen, W. J. Chem. Theory Comput. 2010, 6, 3066.

(40) Godfrey, P.; Firth, S.; Hatherley, L.; Brown, R.; Pierlot, A. J. Am. Chem. Soc. 1993, 115, 9687.

(41) Godfrey, P.; Brown, R.; Rodgers, F. J. Mol. Struct. 1996, 376, 65.

(42) Plekan, O.; Feyer, V.; Richter, R.; Coreno, M.; de Simone, M.; Prince, K.; Carravetta, V. J. Electron. Spectrosc. Relat. Phenom. 2007, 155, 47.

(43) Suenram, R.; Lovas, F. J. Mol. Spectrosc. 1978, 72, 372.

(44) Schafer, L.; Sellers, H.; Lovas, F.; Suenram, R. J. Am. Chem. Soc. 1980, 102, 6566.

(45) Suenram, R.; Lovas, F. J. Am. Chem. Soc. 1980, 102, 7180.

(46) Tse, Y.; Newton, M.; Vishveshwara, S.; Pople, J. J. Am. Chem. Soc. 1978, 100, 4329.

(47) Brown, R.; Godfrey, P.; Storey, J.; Bassez, M.; Robinson, B.; Batchelor, R.; Mcculloch, M.; Rydbeck, O.; Hjalmarson, A. Mon. Not. R. Astron. Soc. 1979, 186, P5.

(48) Csaszar, A. J. Am. Chem. Soc. 1992, 114, 9568.

(49) Hu, C.; Shen, M.; Schaefer, H. J. Am. Chem. Soc. 1993, 115, 2923.

(50) Csaszar, A. J. Mol. Struct. 1995, 346, 141.

(51) Alonso, J.; Pena, I.; Sanz, M.; Vaquero, V.; Mata, S.; Cabezas, C.; Lopez, J. Chem. Commun. 2013, 49, 3443.

(52) Balta, B.; Basma, M.; Aviyente, V.; Zhu, C.; Lifshitz, C. Int. J. Mass

spectrum. 2000, 201, 69.

(53) Dunbar, R.; Polfer, N.; Oomens, J. J. Am. Chem. Soc. 2007, 129, 14562.

(54) Kasalova, V.; Allen, W.; Schaefer, H.; Czinki, E.; Csaszar, A. J. Comput. Chem. 2007, 28, 1373.

(55) Godfrey, P.; Brown, R. J. Am. Chem. Soc. 1995, 117, 2019.

(56) Gronert, S.; O'Hair, R. J. Am. Chem. Soc. 1995, 117, 2071.

Page 53: Mass spectrometric characterization of remotely charged

35  

  

(57) Cao, M.; Newton, S.; Pranata, J.; Schafer, L. THEOCHEM 1995, 332, 251.

(58) Csaszar, A. J. Phys. Chem. 1996, 100, 3541.

(59) Mullin, J.; Gordon, M. J. Phys. Chem. B 2009, 113, 8657.

(60) Stepanian, S.; Reva, I.; Radchenko, E.; Adamowicz, L. J. Phys. Chem. A 1998, 102, 4623.

(61) Hirata, Y.; Kubota, S.; Watanabe, S.; Momose, T.; Kawaguchi, K. J. Mol. Spectrosc. 2008, 251, 314.

(62) Plekan, O.; Feyer, V.; Richter, R.; Coreno, M.; de Simone, M.; Prince, K.; Carravetta, V. Chem. Phys. Lett. 2007, 442, 429.

(63) Plekan, O.; Feyer, V.; Richter, R.; Coreno, M.; de Simone, M.; Prince, K.; Carravetta, V. J. Phys. Chem. A 2007, 111, 10998.

(64) Blanco, S.; Sanz, M.; Lopez, J.; Alonso, J. Proceedings of the National Academy of Sciences of the United States of America 2007, 104, 20183.

(65) Leng, Y.; Zhang, M.; Song, C.; Chen, M.; Lin, Z. THEOCHEM 2008, 858, 52.

(66) Kaczor, A.; Reva, I.; Proniewicz, L.; Fausto, R. Journal of Physical Chemistry a 2006, 110, 2360.

(67) von Helden, G.; Compagnon, I.; Blom, M.; Frankowski, M.; Erlekam, U.; Oomens, J.; Brauer, B.; Gerber, R.; Meijer, G. Physical Chemistry Chemical Physics 2008, 10, 1248.

(68) Huang, Z.; Yu, W.; Lin, Z. Journal of Molecular Structure-Theochem 2006, 801, 7.

(69) Huang, Z.; Yu, W.; Lin, Z. Journal of Molecular Structure-Theochem 2006, 758, 195.

(70) Snoek, L.; Robertson, E.; Kroemer, R.; Simons, J. Chemical Physics Letters 2000, 321, 49.

(71) Perez, C.; Mata, S.; Blanco, S.; Lopez, J.; Alonso, J. Journal of Physical Chemistry a 2011, 115, 9653.

(72) Koirala, D.; Kodithuwakkuge, S.; Wenthold, P. J. Phys. Org. Chem. 2015, 28, 635.

Page 54: Mass spectrometric characterization of remotely charged

36  

  

(73) Blom, M.; Compagnon, I.; Polfer, N.; von Helden, G.; Meijer, G.; Suhai, S.; Paizs, B.; Oomens, J. J. Phys. Chem. A 2007, 111, 7309.

(74) Im, S.; Jang, S.; Lee, S.; Lee, Y.; Kim, B. J. Phys. Chem. A 2008, 112, 9767.

(75) Tian, S.; Sun, X.; Cao, R.; Yang, J. J. of Physical Chemistry a 2009, 113, 480.

(76) Hwang, T.; Eom, G.; Choi, M.; Jang, S.; Kim, J.; Lee, S.; Lee, Y.; Kim, B. J. Phys. Chem. B 2011, 115, 10147.

Page 55: Mass spectrometric characterization of remotely charged

37  

CHAPTER THREE: EXPERIMENTAL APPROACHES

3.1 Introduction

A general overview of the experimental setup and theoretical procedures used in

this thesis are presented in this chapter. Experimental results presented throughout this

thesis are obtained from two mass spectrometry (MS) instruments; namely, flowing-

afterglow triple-quadrupole and LCQ-Deca MS. The flowing afterglow,1 Figure 3.1, is a

homebuilt instrument and LCQ-Deca is a commercial instrument of Thermo Scientific.2

Both of them contain an ion source, mass analyzer and detector, which are common

features of all MS instruments.1,3-7 Chemical compounds are first ionized in the ion

source region. These ions are then transferred to the low-pressure chamber, containing

the mass analyzer, where they are manipulated and discriminated based on mass/charge

(m/z) ratios by electro-magnetic fields. Finally, they are detected and the generated signal

is processed to produce a mass spectrum.6 Our goal in this chapter is to address how these

features are adopted in each of our MS and to specify the methods used to generate data

presented in this thesis. More detailed descriptions of the flowing afterglow1,5 and LCQ-

Deca2 have been provided elsewhere.

Page 56: Mass spectrometric characterization of remotely charged

38  

3.2 The Flowing-Afterglow Triple-Quadrupole Mass Spectrometer

All experiments described in chapter four of this thesis were conducted using a

flowing afterglow triple-quadrupole mass spectrometer. This section is divided in three

sections; describing the ion source, the mass analyzer and the detector implemented in the

flowing afterglow triple-quadrupole MS.

Figure 3.1. The flowing-afterglow triple-quadrupole mass spectrometer.

3.2.1 The Ion Source

The flowing afterglow MS utilizes electron impact ionization method to generate

ions of interest. Historically, electron impact ionization, also known as electron ionization

(EI), was the most commonly used ionization method to create gas-phase ions for a wide

range of organic molecules.6,7 When an electron hits a neutral molecule, it may eject an

Page 57: Mass spectrometric characterization of remotely charged

39  

electron to produce positive ions or combine to form negative ions. Negative ions may

also be formed by dissociative attachment.1

This instrument uses an electron gun consisting of a rhenium metal filament to

emit free electrons when resistively heated. The electrons formed in an ion source are

attracted toward the metal mesh where 70 eV of biased potential is applied. In addition,

there is a continuous flow of dry, helium, carrier gas through an inlet situated adjacent to

the electron gun (Figure 3.1). The helium gas has two purposes: 1) to thermalize the

electrons and 2) to carry electrons downstream for electron ionization.1,5

Ions are synthesized in a one-meter-long flow tube in the gas phase (Figure 3.1).

The mixture of He and electrons are pumped downstream of the flow tube due to a

pressure gradient created by a roots blower vacuum pump. There are six inlet ports on the

flow tube through which vapors of compounds are introduced into the tube and ions are

generated by electron impact ionization. In addition, gas-phase ion-molecule reactions of

the ions residing in flowing helium can be studied by adding reagents like O2, CO2, NO

and CS2 from ports located further downstream on the flow tube.8,9 Furthermore, the flow

tube possesses kinetic rings to introduce a neutral reagent vapor at a calibrated flow rate

and then measuring the extent of reaction.

EI can be applied to volatile samples only because the sample has to be in the gas

phase before electron can ionize it. EI is an endothermic process and if the internal

energy of the compound is less than the energy deposited, the molecule fragments, which

obscures the m/z information of the analytes. Due to this reason, it is categorized as a

hard ionization method.6,7,10 The problem of fragmentation has been overcome by

chemical ionization (CI), a method using gas-phase chemical reactions to produce ions

Page 58: Mass spectrometric characterization of remotely charged

40  

with little to no fragmentation.8,11 All the ions studied in Chapter four are generated by EI.

However, the flowing afterglows the capability for CI is also limited to volatile and

thermally stable organic compounds, and needs to be performed in a low-pressure

atmosphere.

The first step in CI is to form reagent ions by EI, which are then reacted with a

molecule to produce ions of interest. The most useful reagent ions in a negative ion mode

include F- and OH-, which is mainly because these ions are very basic and can easily

deprotonate analytes to form anions of interest. The reagent ions, F- can be created by EI

of fluorine gas and OH- from a 2:1 mixture of methane and nitrous oxide, respectively

(equations 3.1 and 3.2).

(3.1)

(3.2)

3.2.2 The Mass Analyzer

The next region of the flowing afterglow instrument is a triple quadrupole mass

analyzer, which is differentially pumped at 10-6-10-7 Torr by two diffusion pumps. This

section of the instrument is separated from the flow tube by a molybdenum disk held at a

~1-2 V attracting potential. A small amount of gaseous ions produced in the flow tube are

sampled through a 1 mm orifice in a nosecone and are focused to the mass analyzer

Page 59: Mass spectrometric characterization of remotely charged

41  

region by a set of four lenses. The analyzer in the flowing afterglow is a beam type,

where the ions enter and pass through the analyzing field to the detector as a beam,

consisting of three axially aligned quadrupoles. In addition, there are lenses before

(entrance lens) and after (exit lens) the second quadrupole to help focus and guide ions in

and out. The first quadrupole (Q1) can be used to scan a range of m/z (scanning mode) or

to isolate a single m/z (single ion mode). When Q1 is used in scanning mode, the second

(q2) and third (Q3) quadrupoles are used to transfer all ions with no m/z discrimination.

If Q1 is used in the single ion mode, then the ions selected in Q1 can be fragmented in q2

by colliding it with an inert target gas (Ar) or reacted with reagent gases. The first

process is called collision-induced dissociation and the second process is referred to as q2

reactivity. During CID, Ar is dispenses at a pressure of ~150 uTorr with ~10 V of q2 pole

offset. Whereas, q2 reactivity is performed by dispensing reagent gases at a pressure of

~100 uTorr with q2 pole offset close to 0 V to prevent fragmentation. In these processes,

Q3 is used as the mass analyzer. The fragments formed from CID experiments help to

interpret structure of a molecule and q2 reactivity is used to verify the reaction products

observed in the flow tube.

3.2.3 Ion Detection Region

The detector in the flowing afterglow consists of a conversion dynode and an

electron multiplier. The negative ions coming out of Q3 are attracted by the conversion

dynode’s accelerating potential of +1500 V. These ions strike the inner walls of the

electron multiplier and eject electrons. The ejected electrons travel down the funnel

shaped electron multiplier and produce a cascade of electrons resulting in a measurable

Page 60: Mass spectrometric characterization of remotely charged

42  

current proportional to the intensity of the ion. Finally, the signal is collected using the

Q50 software where a mass spectrum is obtained.

3.3 The LCQ-Deca Mass Spectrometer

All experiments described in chapter four, five and six of this thesis were

conducted on LCQ-Deca Mass Spectrometer. This section is divided in three sections;

describing the ion source, the mass analyzer and the detector implemented in the LCQ-

Deca.

3.3.1 The Ion source

The LCQ-Deca uses electrospray ionization (ESI) to produce ions of interest. ESI

is one of the several methods, which ionizes non-volatile and thermally instable

compounds. It also creates little to no fragmentation and is categorized as a soft

ionization method.4,6,7,12-14 However, this method requires compounds to be soluble in

polar solvents like water, methanol, and acetonitrile.14 Generally, a dilute analyte solution

is injected through a hypodermic needle or stainless steel capillary. A very high voltage

(2–6 kV) is applied to the tip of the metal capillary, which causes the dispersion of the

sample solution into an aerosol of highly charged electrospray (ES) droplets. The droplets

are nebulized with a nonreactive gas (usually dry N2) to create a fine spray in which

solvent evaporation occurs as the droplets traverse the atmospheric interface, introducing

molecular ions into the analyzer.4,13 However, the actual mechanism for generating gas-

phase ions via ESI differs depending upon the analyte in question.

Page 61: Mass spectrometric characterization of remotely charged

43  

Ions in the LCQ-Deca are transferred from solution to the gas phase through an

ESI probe. The ESI probe includes sample tube, spray needle, sheath and auxiliary gas

inlets, and nozzle. Solution enters the ESI probe through the sample tube into the spray

needle. The sample tube could be a stainless steel or a fused-silica tube. In our instrument,

0.1 mm ID X 0.19 mm OD fused-silica sample tube is used. The sample tube is located

inside the metal spray needle, where a large negative or positive voltage (typically ±3 to

±5 kV) is applied. This allows the solution to be sprayed into a fine mist of charged

droplets. An ESI nozzle directs the flow of N2, as a sheath and an auxiliary gas, at the

droplets. Sheath gas is an inner coaxial nitrogen gas helping nebulize droplets into a fine

mist as it exits the nozzle. The auxiliary gas is the outer coaxial nitrogen gas assisting the

sheath gas in the nebulization and evaporation of sample solutions. For the experiments

reported in this thesis, the auxiliary gas is not used and a sheath gas flow rate is set in

between 20-30 units of instrumental parameter.

3.3.2 The Mass Analyzer

Located in between the ion source and mass analyzer are a heated capillary, a tube

lens and ion optics. The heated capillary is a heater-embedded metal tube with a hole

bored through the center of its long axis. Ions are drawn into the heated capillary in the

atmospheric pressure region and are transported to the capillary-skimmer region of the

vacuum manifold by a decreasing pressure gradient. Hence, the heated capillary assists in

further desolvating the ions. Ions exiting the heated capillary are focused and accelerated

Page 62: Mass spectrometric characterization of remotely charged

44  

toward the tube lens due to the offset voltage applied on the lens. The tube lens also

serves as a gate to stop the injection of ions into the mass analyzer during ion detection.

Ions from tube lens pass through the skimmer and enter the ion optics region. The

skimmer acts as a vacuum baffle between the higher pressure (1 Torr) and the lower

pressure ion optics regions (10-3 Torr). The ion optics consist a quadrupole, a lens and an

octapole. The quadrupole and octapole act as a transmission device to guide the ions to

the mass analyzer. The lens located in between assists in the focusing ions during the

injection and gating of ions during mass analysis. When a supplemental DC voltage is

added to the overall voltage gradient across the ion path, the non-selective ion

fragmentation occurs, which is called source CID. We have employed source CID in

several occasions to improve the intensity of the secondary ions, which can be studied

later in the mass analyzer.

The mass analyzer in the LCQ-Deca is a quadrupole ion trap, which utilizes a

three-dimensional field to trap ions.6,7,14 The quadrupole ion trap (QIT) is similar to the

quadrupole mass analyzer (QMA) present in flowing afterglow. Whereas a QMA has

electric fields in two dimensions (x and y) and the ions move perpendicular to the field

(the z direction), the QIT has the electric field in all three dimensions, which can result in

ions being trapped in the field.6,7 Ions with a stable trajectory through the quadrupole are

detected in QMA, whereas the ions with unstable trajectories get ejected from the trap

and are detected in QIT.15

The QIT in Deca includes three stainless steel electrodes: two of those are endcap

electrodes, known as the entrance and the exit, the third is a hyperbolic ring situated

centrally between the end-cap electrodes to form a cavity. The two quartz spacer rings

Page 63: Mass spectrometric characterization of remotely charged

45  

serving as electric insulators separate these electrodes. Ions are injected to the mass

analyzer cavity via a small opening in the entrance end cap electrode. Helium is used as a

dampening gas inside the ion trap, due to its ability to cool the ions without inducing

fragmentation. The ions are focused to the axis of the cavity due to dampening and are

trapped due to the RF voltage applied in the ring electrode. A resonance ejection RF

voltage is applied to the endcaps of the mass analyzer to minimize space charge effects in

the mass analyzer. When an ion is to be ejected from the mass analyzer cavity, it is

brought towards the frequency of the resonance ejection RF voltage. The application of

resonance ejection increases resolution because ions of a specific m/z leave the trap in a

more condensed manner.

An ion of interest can be isolated in QIT of LCQ-Deca by ejection of other ions

out of the trap. The ions with different m/z have different oscillatory frequencies in the

ion trap so can be separated. The ion isolation waveform voltage, which consists of a

distribution of resonance frequencies corresponding to unwanted ions, is applied to the

endcaps to eject all ions except those of a selected m/z. Thus isolated ions can be excited

by applying a resonance excitation RF voltage to the endcaps, which creates energetic

collisions of ions with the background helium gas and cause parent ions to fragment into

product ions. However, the resonance excitation RF voltage is not strong enough to eject

an ion from the mass analyzer. Therefore applied voltage can be tuned to get desired

fragmentation. Sequences of isolation followed by fragmentation then isolation again can

be performed for MSn experiment.

Page 64: Mass spectrometric characterization of remotely charged

46  

3.3.3 Ion Detection Region

The detector in the LCQ-Deca consists of a conversion dynode and an electron

multiplier. Ions are attracted toward the conversion dynode due to applied offset voltage

and produce secondary particles when ions hit the surface of the dynode. These

secondary particles are focused and are accelerated by a voltage gradient into the electron

multiplier. Secondary particles from the conversion dynode strike the inner walls of the

electron multiplier and eject electrons. The ejected electrons travel down the funnel

shaped electron multiplier and produce a cascade of electrons. It results in a measurable

current proportional to the intensity of the ion. Finally, the signal is collected using the

Xcalibur data system where a mass spectrum is obtained.

Page 65: Mass spectrometric characterization of remotely charged

47  

3.4 References

(1) Graul, S.; Squires, R. Mass Spectrom. Rev. 1988, 7, 263.

(2) LCQ Deca Mass Spectrometry, Manual

(3) Demartini., D. R. A Short Overview of the Components in Mass Spectrometry Instrumentation for Proteomics Analyses, 2013.

(4) Cooks, R.; Ouyang, Z.; Takats, Z.; Wiseman, J. Science 2006, 311, 1566.

(5) Marinelli, P.; Paulino, J.; Sunderlin, L.; Wenthold, P.; Poutsma, J.; Squires, R. Int. J. Mass Spectrom. Ion Processes 1994, 130, 89.

(6) Maher, S.; Jjunju, F.; Taylor, S. Rev Mod Phys 2015, 87, 113.

(7) Glish, G.; Vachet, R. Nat. Rev. Drug Discov. 2003, 2, 140.

(8) Rau, N.; Welles, E.; Wenthold, P. J. Am. Chem. Soc. 2013, 135, 683.

(9) Koirala, D.; Poole, J.; Wenthold, P. Int. J. Mass spectrom. 2015, 378, 69.

(10) WANNIER, G. Phys. Rev. 1953, 90, 817.

(11) Munson, M.; Field, F. J. Am. Chem. Soc. 1966, 88, 2621.

(12) Takats, Z.; Wiseman, J.; Gologan, B.; Cooks, R. Science 2004, 306, 471.

(13) Banerjee, S.; Mazumdar, S. Int. J. Anal. Chem. 2012.

(14) El-Aneed, A.; Cohen, A.; Banoub, J. Appl. Spectrosc. Rev 2009, 44, 210.

(15) Stafford, G.; Kelley, P.; Syka, J.; Reynolds, W.; Todd, J. Int. J. Mass Spectrom. Ion Processes 1984, 60, 85.

Page 66: Mass spectrometric characterization of remotely charged

48  

CHAPTER FOUR: THE FLOWING AFTERGLOW STUDY OF PYRIDINE N-OXIDE NITRENE RADICAL ANIONS

4.1 Introduction

Nitrenes are a fascinating class of electron-deficient and highly reactive

intermediates. They have a nitrogen motif with two bonding and four non–bonding

electrons in valence shell which are isoelectronic to their carbon analogues carbenes.

Nitrenes are nominally sp-hybridized with two electrons in the lower energy sp-orbital as

a pair, whereas the remaining two electrons can be oriented in two available p-orbitals in

four different ways as illustrated in Scheme 4.1. Although isoelectronic, nitrenes and

carbenes show very different reactivity.1-3 For example, both phenylcarbene (PhC) and

phenylnitrene (PhN) can be generated by photolysis of nitrogen containing precursors,

and are generated in the lowest energy singlet state. The initially formed closed–shell

singlet PhC can rapidly undergo intersystem crossing (ISC) to form a ground state triplet

which shows clean bimolecular reactivity. ISC is not favorable for PhN, so PhN

undergoes unimolecular reaction via lowest energy singlet resulting in forming of

polymeric tar when reacted with alkene, alcohol, and aromatic compounds.1,4

Page 67: Mass spectrometric characterization of remotely charged

49  

There are several studies carried out in the reactivity of nitrenes and their study

has led to many important mechanistic ideas, 1,4-6 however, only few studies of nitrene

radical anions have been reported. In addition, many of these studies are focused on the

use of the anions as precursors for photoelectron/photodetachment spectroscopy studies

of the corresponding nitrenes.7-11 Nitrene radical anions are formed when an electron is

added to nitrene. The PhN shown in Figure 4.1 can have two PhN¯ radical anions. Beside

sp-hybridezed electrons, PhN¯ consist of three electrons in two non-bonding molecular

orbitals, an in-plane σ orbitals and a π orbital (Scheme 4.2). The electronic structures of

phenylnitrene anions have been discussed in the context of photoelectron10,11 and

photodetachment7-9 spectroscopy studies, and have been examined by electronic structure

calculations. These studies have shown that the ground state of PhN is π2σ, with the σ2π

state lying approximately 0.5eV higher in energy.

Scheme 4.1

Page 68: Mass spectrometric characterization of remotely charged

50  

Benzoylnitrene radical anion, BzN, has been examined computationally and

experimentally.12-14 Like PhN, the ground state of BzN is predicted to be π2σ, but the

relative energy of the σ2π state (28 kcal/mol) is predicted to be much higher than that in

PhN, reflecting a greater preference for the carboxylate–like anion over having the

charge localized on a nitrogen σ orbital. Chemical reactivity of the BzN is consistent

with the open–shell structure like in PhN, in particular, reaction with NO results in

radical–radical coupling to give a net nitrogen/oxygen exchange, leading to the formation

of benzoate anion (eq 4.1).  

(4.1a)

(4.1b)

Scheme 4.2

Page 69: Mass spectrometric characterization of remotely charged

51  

Here, we investigate the reactivity of the 3– and 4–pyridinylnitrene–n–oxide

radical anions, 3PNO and 4PNO respectively.

These are interesting isomers to study because they contain the moieties of two

well characterized compounds: phenylnitrene anion (PN) and pyridine–N–oxide (PO).

PN contains the unusual hypovalent anionic nitrene, whereas PO has a persistent

zwitterionic N–oxide moiety. In addition, PO can nominally be viewed as an electron–

acceptor, but the adjacent oxide can act as a π donor, such that the electronic effect of the

N–oxide is dependent on the extent of stability afforded by electron donation or

acceptance (Figure 4.1). However, the resonance effect of the N–oxide is greatest at the

2– and 4–position, and not at the 3–position. Hence, 4PNO can be expected to be more

stabilized than 3PNO due to position of nitrene. Moreover, for 4PNO,  there are two

types of stabilization that can be envisioned to result from interaction between the nitrene

radical anion and the N–oxide, shown in Figure 4.1. In the π2σ state, the pyridinium can

serve as an electron pair acceptor, and the system can potentially be stabilized by

contribution from the quinone–like structure. Similarly, the σ2π state can be stabilized by

Page 70: Mass spectrometric characterization of remotely charged

52  

radical delocalization, creating a highly stable nitroxyl radical. Electron delocalization in

either state is such that it eliminates the formal charge separation of the n–oxides, which

could provide additional stabilization.

Figure 4.1. Resonance effect due to N-Oxide motif (Left) and the resonance structures of the π2α and α2π states of 4PNO (Right)

We have carried out a computational study of the electronic structures of 3PNO

and 4PNO, and an experimental investigation of their reactivity. Surprisingly, despite

the possible interactions between the nitrene radical anion and N–oxide in 4PNO, we do

not find significant differences in the computed electronic structure or reactivity of the

two isomers in the reaction with O2, CO2 and NO. However, a dramatic difference

between the ions is found in the reaction with CS2, as the 4–isomer undergoes reactivity

that is not observed for the 3–isomer, attributed to differences in accessibility of the oxide

end of the ion toward reaction. The results indicate that the resonance interaction

Page 71: Mass spectrometric characterization of remotely charged

53  

between the nitrene anion center and the n-oxide increases the nucleophilicity of the

oxygen in the n-oxide moiety.

4.2 Computational Detail

Structures and energies of the 2 and 2 states of 3PNO and 4PNO, NO, CS2

and possible reaction products were calculated at the B3LYP/6-31+G* level of theory.

Reaction energies correspond to differences in electronic energies between reactants and

products, and are not corrected for zero-point energies or thermal energy differences.

Calculations were carried out by using QChem,15 using the resources of the Center for

Computational Studies of Open-Shell and Electronically Excited Species

(iopenshell.usc.edu).

4.3 Results

The 4PNO- and 3PNO- radical anions are formed in high yields by dissociative

electron attachment to the corresponding azides (eq 4.2).

Page 72: Mass spectrometric characterization of remotely charged

54  

The azidopyridine-n-oxide precursors were prepared and supplied by our

collaborator James S. Poole from Ball State University. The synthesis and

characterization of these samples used in this work has been reported previously in

several occasions.6,16-19 The most significant impurity observed in these experiments

(with both isomers) is an ion with m/z 92, which is presumably the non-oxidized

pyridinylnitrene radical anion. The relative intensity of the m/z 92 ion is variable from

day to day, but decreases with time during the course of experiment run, indicating it

most likely results from ionization of a non-oxidized azidopyridine impurity.

Single point energies calculated at B3LYP /6–31+G* suggests that both isomers

have π2σ electronic ground states but 4PNO is more stable by 2.3 kcal/mol than 3PNO.

This electronic ground state is consistent to other aromatic nitrene radical anions like

benzoylnitrene radical anion. The σ2π states of 4PNO and 3PNO is 13.5 kcal/mol and

15.0 kcal/mol higher in energy than π2σ of 4PNO. The small difference in energy, 1.5

kcal/mol, between σ2π states of 4PNO and 3PNO reflects there is no significant

resonance effect, stabilization or destabilization, of the N–oxide in 4PNO as expected

(Figure 4.1).

 

4.3.1 Reactivity Studies

Reactions of 3PNO and 4PNO with O2, CO2, NO and CS2 are studied. When

reacted with O2 and CO2 both isomers produce a single peak at m/z 140 and at m/z 152

respectively. Upon CID, both of these product yield m/z 108 corresponding to the

molecular ion hence m/z 140 and m/z 152 are assigned to be O2 and CO2 adduct

Page 73: Mass spectrometric characterization of remotely charged

55  

respectively. Benzoylnitrene radical anion was found to be unreactive with CO2.13

Besides forming adducts and reacting differently with CO2 as compared to

benzoylnitrene, there is not much we could grasp on electronic structure from reactivity

of 3PNO and 4PNO with O2 and CO2.On the other hand, reactivity study of 3PNO

and 4PNO with NO and CS2 would help to interpret the electronic structure of

nitrenes.5,13 

The results of reactivity studies of ions 3PNO and 4PNO with NO and CS2 are

shown in Table 4.1. Overall, there are no measurable differences in the rates of the

reactions for the two isomers. The efficiencies of the reactions with NO (approximately

0.2) are much higher than those found for reaction with CS2, but similar to that reported

previously for the reaction of NO with BzN (eff = 0.15).13 Although there are no

differences in the reaction rates for the two isomers with these reagents, there are very

large differences in the observed products. In the sections below, we consider the

differences in the products that are formed.

Page 74: Mass spectrometric characterization of remotely charged

56  

Table 4.1: Reaction Efficiencies and Product Branching Ratios for Reactions of 3PNO and 4PNO with NO and CS2

reagent result Ion assignment 3PNO– 4PNO–

NO Effa 0.21 0.20 m/z 138 [M+NO] – <1% 18% m/z 110 [M+NO–N2] – 44% 82% m/z 82 [M+NO–N2–CO] – 21% Nb

m/z 80 [M+NO–N2–NO] – 20% Nb

m/z 61 [M+NO–C3H3–N2]– 14% Nb

m/z 92 [M+NO–NO2] – Yc Yc

CS2 Effa 0.03 0.02 m/z 186 [M+CS2] – 83% 77% m/z 64 S2

– 15% 15% m/z 32 S– <1% <1% m/z 58 SCN– 2% 2% m/z 124 [M+CS2–CSO] – Nb 7% m/z 92 [M-O] – Nb Yc

aReaction efficiency, which corresponds to kexp/kADO bNot observed for this isomer. cProduct is observed in Q2, but the flow tube branching ratio cannot be determined due to the presence of impurities

4.3.1.1Reaction with NO

With nitric oxide, both 3PNO and 4PNO are observed to undergo N-O

exchange, similar to what is observed with PhN and BzN.13 Many additional products

are observed for the reaction of 3PNO with NO, but, as shown Table 1, they are

assigned to secondary fragmentation of the phenoxide anion. Possible product structures

are shown in eq 4.3. A notable difference in the reactivity of 3PNO and 4PNO with

Page 75: Mass spectrometric characterization of remotely charged

57  

NO is that reaction with 4PNO leads to significant amount of adduct ion, whereas only a

trace is observed with 3PNO. NO addition was not reported for either PhN or BzN,13

and is generally associated with reactions of closed-shell anions.5,20 Unfortunately, the

structures of the NO adducts are not known. However, charge distribution calculations

reported below find that there is more charge on the oxygen in 4PNO than in 3PNO,

which raises the possibility that the difference in the extent of adduct formation is due

4PNO forming an adduct at the oxygen, as opposed to at the nitrogen.

Both n-oxide anions are found to react with NO by oxygen atom transfer to form

the pyridinylnitrene radical anion, as shown for 4PNO in eq 4.4. Although the

branching ratios for these reactions in the flow tube cannot be determined due to the

presence of impurity ions, they were verified as products by using the reaction of NO

with mass-selected ions in Q2.

Page 76: Mass spectrometric characterization of remotely charged

58  

If the NO BDEs in 3PNO and 4PNO are similar to that in pyridine-n-oxide

(approximately 72 kcal/mol),21 then the oxygen atom transfer reaction would be

essentially thermoneutral. Alternatively, given that the reaction is carried out in Q2,

albeit under very low energy conditions, the oxygen atom transfer could be slightly

endothermic and translationally driven.

4.3.1.2 Reaction with CS2

Although the rates at which 3PNO and 4PNO react with CS2 are similar, there

are significant differences between the reaction products. In particular, some products

observed with 4PNO involve loss of the oxygen atom, and these products are not

observed with 3PNO. One product involving the oxygen is observed at m/z 124, which

is 16 Daltons higher than the reactant. The most likely assignment for this product is that

it arises from addition of sulfur and loss of oxygen, i.e. a sulfur-oxygen exchange. The

resulting products would be the n-sulfide and carbonyl sulfide, OCS.

The ion signal at m/z 92 is also likely due to reaction involving the oxygen. As in

the reaction with NO, we are unable to quantify the yield of the m/z 92 product due to

background presence of the pyridinylnitrene radical anion. However, considering 4PyN

is also observed to react with CS2, it appears that m/z 92 is, in fact, formed in relatively

Page 77: Mass spectrometric characterization of remotely charged

59  

high yield (comparable to the yield of S2). This is also indicated by the results on “clean”

days, where the mass spectrum includes minimal amounts of the impurity. The formation

of m/z 92 in the reaction of 4PNO with CS2 was also confirmed by using mass-selected

ions in Q2.

The identity of the m/z 92 is not known unequivocally. Although the ion with m/z

92 could be the pyridinylnitrene radical anion, there is a second possibility. The ion

CS2O is isobaric with the pyridinylnitrene radical anion, and could, in principle be

formed by oxygen-anion transfer with 4PNO, as shown in eq 4.5.

 

Our experimental results suggest both structures are present. Evidence for the

formation of CS2O comes from the isotopic distribution of the product. Figure 4.2

shows a mass spectrum of 4PNO taken on an exceptionally clean day (with minimal m/z

92 contamination), with and without the addition of CS2. The background signal of m/z

92 in the spectrum is less than 10 kcps. However, upon addition of CS2, the signal

increases to nearly 100 kcps. Most importantly, the signal at m/z 94 also increases, to

nearly 8% of the m/z 92 signal, which agrees well with the value of 9% expected for

CS2O. Therefore, the M+2 isotope peak indicates that there is sulfur present in the m/z

92 ion.

Page 78: Mass spectrometric characterization of remotely charged

60  

However, reactivity evidence suggests the presence of 4PyN as well. Figure 4.3

shows the mass spectra for reaction of m/z 92 ion, formed by reaction of 4PNO with CS-

2, with NO. The formation of m/z 94 is clear evidence for the presence of a nitrene

radical anion, 4PyN.

Figure 4.2. The m/z 92 region of the mass spectra of 4PNO before (dashed) and after (solid) addition of CS2

Page 79: Mass spectrometric characterization of remotely charged

61  

 

Figure 4.3. The m/z 92 region of the mass spectra of 4PNO reaction with CS2 (dashed) and after addition of NO (solid)

Computationally, both reactions are computed to energetically favorable. At the

B3LYP/6-31+G* level of theory, the formation of CS2O and the triplet pyridinylnitrene

(eq 4.5) is computed to be exothermic by 72 kcal/mol, or more than 3 eV! For oxygen

atom transfer, there are multiple thermochemically accessible pathways. Direct transfer

to form CS2O (eq 4.6a) is computed to be exothermic by 7.7 kcal/mol. A second

pathway, shown in eq 4.6b, involves the formation of CO + triplet S2, and is computed to

be exothermic by 9.9 kcal/mol.

Page 80: Mass spectrometric characterization of remotely charged

62  

Deoxygenation of teritiary-amine-n-oxides in solution22-24 has been proposed to

occur by a mechanism such as that shown in Scheme 4.3, leading to the formation of the

dithiiranone as in eq 4.6a. In solution, the dithiiranone can be hydrolyzed to CO2 +

HSSH22 or can be utilized as a sulfur transfer reagent25 .

 

In the gas phase, the reaction of oxygen atom with CS2 gives CS + SO as the

major products.26 However, CO has also been observed in the reaction, 27 indicating that

CO + S2 is a possible decomposition pathway of CS2O. In fact, CO + S2 is energetically

Scheme 4.3

Page 81: Mass spectrometric characterization of remotely charged

63  

the most favored pathway26 but is slow because there is a slight (6.6 kcal/mol) barrier26

for the initial formation of CS2O. However, this barrier can be overcome in the reaction

of CS2 with 4PNO by the energy released upon formation of the initial ion/molecule

complex in combination with the oxygen atom transfer exothermicity. Therefore, oxygen

atom transfer to form CO and S2 products should be able to occur for 4PNO. Formation

of CS + SO is approximately 3 eV less favorable26 than formation of CO + S2 and is

therefore highly endothermic in the reaction of 4PNO with CS2.

As noted above, sulfur-oxygen exchange is also observed in the reaction of CS2

with 4PNO. The mechanism of sulfur-oxygen exchange likely involves a first step

similar to that for oxygen atom transfer, addition of the oxygen to the center carbon of

CS2, as shown in Scheme 4.4. However, instead of forming the disulfide bond as in

dithiiranone formation, a N-S bond is formed.

A small amount of NCS product is observed in the reaction of CS2 with both

3PNO and 4PNO. NCS has been observed previously in reactions of CS2 with closed-

shell, nitrogen-based anions,5,28,29but has not been reported previously for reactions of

open-shell anions. The mechanism is presumably similar to that shown in Scheme 4.4,

although occurring at the nitrogen. The other products observed with CS2 (S, S2 and

CS2 adduct) are similar to those observed previously with nitrene radical anions.13

Page 82: Mass spectrometric characterization of remotely charged

64  

4.3.1.3 Comparison of Isomers: Oxygen Nucleophilicity

There are significant differences in the reactions that occur between ions 3PNO

and 4PNO and CS2. Specifically, 4PNO undergoes sulfur-oxygen exchange and

oxygen-atom and/or oxygen-anion transfer reactions that are not observed with 3PNO.

The common feature of these reactions is that they all involve initial nucleophilic attack

of the oxygen in the anion at the carbon of the carbon disulfide. The fact that 4PNO

undergoes reaction at the oxygen whereas 3PNO does not reflects important differences

in their electronic structures.

The best example that illustrates the electronic structure differences between the

ions is the oxygen-transfer reaction with CS2. Although carbon disulfide is known to

deoxygenate tertiary-amine-n-oxides,22 the reaction is generally not observed with

aromatic-n-oxides. Therefore, the lack of oxygen transfer with 3PNO is consistent with

the expectation that the substituent in the meta position does not interact with the n-oxide

moiety. Similarly, the nitrene anion para to the n-oxide can serve as a -donor, as shown

Scheme 4.4

Page 83: Mass spectrometric characterization of remotely charged

65  

in Figure 4.1, which makes the oxygen in the 2 state more nucleophilic. Hammett

analysis of the deoxygenation of aniline-n-oxides in solution has shown22 that the

reaction is favored by strong electron donating substituents, which increase the

nucleophilicity of the oxygen. Apparently, the nitrene radical anion is such a strong -

electron donor that it even enables oxygen transfer in the normally inert aromatic-n-

oxides.

The difference in the reactivity cannot be attributed to differences in overall

thermochemistry for the reactions. All of the reactions observed for 4PNO are also

computed to be exothermic for 3PNO, although they don’t occur. For example, the

sulfur-oxygen exchange reaction observed for 4PNO (Scheme 4.4) is computed to be

exothermic by 45 kcal/mol. Similarly, the sulfur-oxygen exchange reaction for 3PNO is

computed to be exothermic by 37 kcal/mol, but does not occur at all. Therefore, the

difference in reactivity is more likely due to kinetic differences that result from

differences in electronic structure.

The increased nucleophilicity of the oxygen in 4PNO apparent in the reactivity

studies is consistent with the computed charge distributions in the 2 states. The

B3LYP/6-31+G* computed Mullikin charges at the heavy atoms in 3PNO and 4PNO

are shown in Figure 4.4.

Page 84: Mass spectrometric characterization of remotely charged

66  

Figure 4.4. Calculated Mullikin charge distributions in 3PNO, 4PNO and pyridine-n-oxide. Charges on hydrogen have been summed into the heavy atoms.

Although both ions have increased charge density at the oxygen than is found in

pyridine-n-oxide there are some differences in charge densities at the carbon atoms, the

most significant difference is for the oxygen, where the calculated charge is larger in in

4PNO than in 3PNO, which accounts for the increased reactivity at that site. Increased

reactivity at the oxygen accounts for the observation of either oxygen atom (eq 4.6) or

oxygen ion (eq 4.5) transfer with

As suggested above, the difference in the charge density at the oxygen may also

account for the difference in the extent of adduct formation with nitric oxide. This is

most likely the case if the adduct is an electrostatic complex. However, as noted, the

structure of the adduct is not known. Aside from the extent of adduct formation, there is

little difference in the reactivity of with nitric oxide. The differences in the observed

products can be attributed to differences in the ability of the resulting phenoxide ion to

fragment.

Page 85: Mass spectrometric characterization of remotely charged

67  

4.4 Conclusion

The reactivity of 3PNO and 4PNO, particularly with CS2, show that there are

significant differences in their electronic structures, which can be understood as resulting

from the resonance interaction between the nitrene anion and the n-oxide moiety in

4PNO, and the lack thereof in 3PNO. In the 2 state, the monovalent nitrogen anion

is a strong electron donor, which increases the charge density on the oxygen, as shown

in Figure 4.2. The result is consistent with the conclusions based on condensed-phase

studies that oxygen atom transfer is favored by -donors.22 However, the reaction has not

been observed previously for aromatic n-oxides. This work shows that oxygen atom

transfer for aromatic n-oxides can occur with sufficiently strong donors. The

differences in the electronic structures of 3PNO and 4PNO do not result in differences

in reactivity with NO, although there are differences in the stabilities of the resulting

products.

Page 86: Mass spectrometric characterization of remotely charged

68  

4.5 References

(1) Platz, M. Acc. Chem. Res. 1995, 28, 487.

(2) Horner, L.; Christma.A . Angew. Chem. Int. Ed. 1963, 75, 707.

(3) Belloli, R. J. Chem. Educ. 1971, 48, 422.

(4) Borden, W.; Gritsan, N.; Hadad, C.; Karney, W.; Kemnitz, C.; Platz, M. Acc. Chem. Res. 2000, 33, 765.

(5) Rau, N.; Welles, E.; Wenthold, P. Journal of the American Chemical Society 2013, 135, 683.

(6) Hostetler, K.; Crabtree, K.; Poole, J. J. Org. Chem. 2006, 71, 9023.

(7) Drzaic, P.; Brauman, J. J. Am. Chem. Soc. 1984, 106, 3443.

(8) Drzaic, P.; Brauman, J. J. Phys. Chem. 1984, 88, 5285.

(9) McDonald, R.; Davidson, S. J. Am. Chem. Soc. 1993, 115, 10857.

(10) Travers, M.; Cowles, D.; Clifford, E.; Ellison, G. J. Am. Chem. Soc. 1992, 114, 8699.

(11) Wijeratne, N.; Da Fonte, M.; Ronenius, A.; Wyss, P.; Tahmassebi, D.; Wenthold, P. J. Phys. Chem. A 2009, 113, 9467.

(12) Wijeratne, N.; Wenthold, P. J. Am. Soc. Mass. Spectrom. 2007, 18, 2014.

(13) Wijeratne, N.; Wenthold, P. J. Org. Chem. 2007, 72, 9518.

(14) Wijeratne, N.; Wenthold, P. J. Phys. Chem. A 2007, 111, 10712.

(15) Shao, Y.; Molnar, L.; Jung, Y.; Kussmann, J.; Ochsenfeld, C.; Brown, S.; Gilbert, A.; Slipchenko, L.; Levchenko, S.; O'Neill, D.; DiStasio, R.; Lochan, R.; Wang, T.; Beran, G.; Besley, N.; Herbert, J.; Lin, C.; Van Voorhis, T.; Chien, S.; Sodt, A.; Steele, R.; Rassolov, V.; Maslen, P.; Korambath, P.; Adamson, R.; Austin, B.; Baker, J.; Byrd, E.; Dachsel, H.; Doerksen, R.; Dreuw, A.; Dunietz, B.; Dutoi, A.; Furlani, T.; Gwaltney, S.; Heyden, A.; Hirata, S.; Hsu, C.; Kedziora, G.; Khalliulin, R.; Klunzinger, P.; Lee, A.; Lee, M.; Liang, W.; Lotan, I.; Nair, N.; Peters, B.; Proynov, E.; Pieniazek, P.; Rhee, Y.; Ritchie, J.; Rosta, E.; Sherrill, C.; Simmonett, A.; Subotnik, J.; Woodcock, H.; Zhang, W.; Bell, A.; Chakraborty, A.; Chipman, D.; Keil, F.; Warshel, A.; Hehre, W.; Schaefer, H.; Kong, J.; Krylov, A.; Gill, P.; Head-Gordon, M. Phys Chem Chem Phys. 2006, 8, 3172.

(16) Itai, T.; Okusa, G.; Sako, S. Chem. Pharm. Bull. 1963, 11, 1146.

Page 87: Mass spectrometric characterization of remotely charged

69  

(17) Itai, T.; Natsume, S. Chem. Pharm. Bull. 1964, 12, 228.

(18) Crabtree, K.; Hostetler, K.; Munsch, T.; Neuhaus, P.; Lahti, P.; Sander, W.; Poole, J. J. Org. Chem. 2008, 73, 3441.

(19) Sawanishi, H.; Tajima, K.; Tsuchiya, T. Chem. Pharm. Bull. 1987, 35, 4101.

(20) Chacko, S.; Wenthold, P. Mass Spectrom. Rev. 2006, 25, 112.

(21) Afeefy, H. Y.; Liebman, J. F.; Stein, S. E.; National Institute of Standards and Technology, Gaithersburg, MD 20899: retrived February 11, 2014.

(22) Yoshimura, T.; Asada, K.; Oae, S. Bull. Chem. Soc. Jpn. 1982, 55, 3000.

(23) Hamana, M.; Nakashima, S.; Umezawa, B. Chem. Pharm. Bull. 1962, 10, 969.

(24) Nakayama, J.; Ishii, A. Adv. Heterocycl. Chem. 2000, 77, 221.

(25) Zipplies, M.; Devos, M.; Bruice, T. J. Org. Chem. 1985, 50, 3228.

(26) Saheb, V. J. Phys. Chem. A 2011, 115, 4263.

(27) Hudgens, J.; Gleaves, J.; McDonald, J. J. Chem. Phys. 1976, 64, 2528.

(28) Barlow, S.; Bierbaum, V. J. Chem. Phys. 1990, 92, 3442.

(29) Depuy, C. Org. Mass Spectrom. 1985, 20, 556.

Page 88: Mass spectrometric characterization of remotely charged

70

CHAPTER FIVE: MASS SPECTROMETRIC STUDY OF THE DECOMPOSITION PATHWAYS OF CANONICAL AMINO ACIDS AND α-LACTONES IN THE GAS

PHASE

5.1 Introduction

The significance of amino acids to the world around us, and to life itself, cannot

be overstated, and, therefore, it is not surprising that their structures, properties, and

reactivity are exceptionally well-characterized, at least in solution. However, owing to

their general lack of volatility, much less is known about the structure and reactivity of

amino acids in the gaseous state. Gaseous amino acids are rare, but would be highly

significant if reports of interstellar glycine, 1,2 for example, can be confirmed. In a more

practical sense, the gas phase serves as a reasonable approximation for a low dielectric,

hydrophobic environment, and so investigation of the gas-phase (solvent free) properties

of amino acids provides insight into the structures of amino acids in lipid membranes.3

Finally, regardless of any biological significance, results of studies of gaseous amino

acids can be compared to those for substituted carboxylic acids to determine, inter alia,

the effect of the amino group on the gas-phase structure and reactivity.

One of the most important differences between gaseous and condensed-phase

amino acids is that, because charge separation is unfavorable in the gas phase, the

solvent-free molecules will not have zwitterionic structures, Z, like those in condensed

phase, but will have non-ionic, canonical (C) structures instead.4-7 In fact, zwitterionic

Page 89: Mass spectrometric characterization of remotely charged

71

structures of gas-phase amino acids are so unexpected that many mass spectrometric

studies have been carried out to try to elucidate those factors, such as solvation or ion-

pairing, that can lead to stable ionic structures in various amino acids.8

Recent advances in spectroscopic methods in the last decade have provided more

detailed insight into the gas-phase structures, including details of the modes of internal

hydrogen bonding.9-11 The challenges in these experiments are two-fold: on one hand,

because the spectroscopy experiments generally require a good chromophore for

ultraviolet absorption, they work best for amino acids with aromatic side chains.

However, those amino acids are among the less volatile, and therefore creating gaseous

samples is difficult. Smaller, aliphatic amino acids are more volatile, but lack a good

chromophore to facilitate resononace-2-photon absorption, which is generally required

for these types of experiments. It is a testament to modern spectroscopic technology that

we do have spectra for both aromatic and aliphatic amino acids.12-15 Gas-phase

photoionization of aliphatic amino acids has also been reported.16

Amino acids have been extensively studied using mass spectrometry. The

development of ionization techniques like electrospray ionization (ESI), matrix assisted

laser desorption ionization (MALDI), and desorption electrospray ionization (DESI) has

made the desorption and ionization of non–volatile substrate, like AA, possible allowing

characterization using mass spectrometry (MS). However, in these studies, the amino

Page 90: Mass spectrometric characterization of remotely charged

72

acids are typically ionized in some way, such as via protonation or metallation to create

positively-charged ions, or by deprotonation to form negatively-charged ions.

Consequently, the properties of ionized amino acids investigated by using mass

spectrometry are generally not comparable to those of the “neutral” (zwitterionic or

canonical) amino acids. Here we describe an investigation of canonical amino acids by

using “charge-remote fragmentation,” 17,18 where dissociation of an ion occurs, as the

name suggests, at locations remote from the charge site. Charge-remote fragmentation is

commonly used as an analytical tool to determine the structures of biological molecules.

Lipids17,18 are particularly amenable to this technique, as they fragment in the

hydrocarbon chains. However, charge-remote fragmentation has also been used to

examine structures of other biologically relevant molecules such as drugs and their

metabolites and peptides.17,18

Adams and Gross19 have proposed that charge-remote fragmentation, where the

charge does not participate in the fragmentation, is analogous to thermolysis of the gas-

phase neutral molecule. If this is true, then it may be possible to use charge-remote

fragmentation electrosprayed ions as an alternative to pyrolysis in order to study the

gaseous decomposition of otherwise non-volatile organic molecules. In this study, we

examined sulfonated phenylalanine, PheSO3, a sulfonate anion with a separate canonical

amino acid moiety.

Page 91: Mass spectrometric characterization of remotely charged

73

The gas-phase reactivity of amino acids is similarly poorly known. The lack of

volatility of amino acids has made even nominally simple experiments such as gas-phase

pyrolysis exceedingly difficult, and these experiments generally require generation of the

amino acid in situ by pyrolysis of appropriate volatile precursors. For example, Chuchani

and co-workers20 reported that they could examine the pyrolytic dissociation of amino

acids generated by decomposition of the corresponding ethyl ester derivatives (eq 5.1).

By using this approach, they found that N-substituted glycines, such as N,N-dimethyl-,20

N-phenyl-21 and N-benzylglycines,22 for example, dissociate by loss of CO2, as shown in

eq 5.1, in the gas phase. In contrast, Al-Awadi and co-workers found that gaseous N-

phenylalanine does not undergo decarboxylation upon pyrolysis, but dissociates into

aniline, CO, and acetaldehyde, as shown in eq 5.2,23 in a process that is similar to what is

commonly observed for α-substituted carboxylic acids (eq 5.3).24-36

Page 92: Mass spectrometric characterization of remotely charged

74

By using mass spectrometry, we have examined the decomposition of the gaseous

amino acid, PheSO3, which contains a sulfonate charge and a canononical -amino acid.

We show that PheSO3 dissociates initially by fragmentation of the amino acid by the

pathway shown in eq 5.3 to form the α-lactone, and direct decarboxylation (eq 5.1) is not

observed. The α-lactone product is found to fragment by decarbonylation, as expected

(eq 5.3) but is also found to undergo decarboxylation to form the styrene product.

Electronic structure calculations find that the presence of the sulfonate charge has little

effect on the energetics of these dissociation pathways, indicating that the results can be

fairly interpreted as representative of what would happen with the non-ionized derivative.

5.2 Experimental

Experiments were carried out using a commercial LCQ-DECA (Thermo Electron

Corporation, San Jose, CA) quadrupole ion trap mass spectrometer, equipped with

electrospray ionization (ESI). Ions were introduced by spraying dilute aqueous solutions

of commercially available substrates, except for the cis-cinnamic acid and 4-

(acryloyloxy)benzenesulfonate, which were synthesized as described below; solution

details are provided as Supporting Information. Electrospray and ion focusing conditions

were varied so to maximize the signal of the ion of interest. Dissociation of ions can be

carried out during ion injection by using high injection voltages, or can be carried out by

using MSn experiments with mass-selected ions in the cell, with the helium buffer serving

as the collision target. Reactant ions for CID were isolated at qz = 0.250 and with a mass-

width sufficient to avoid off-resonance excitation of the mass-selected ions. The energy

Page 93: Mass spectrometric characterization of remotely charged

75

of collision-induced dissociation (CID) in the cell is reflected in the “normalized collision

energy,” which ranges from 0 – 100%.

5.3 Synthesis

5.3.1 (Z)-4-(3-ethoxy-3-oxoprop-1-en-1-yl)benzenesulfonate (deprotonated ethyl 4-sulfo-cis-cinnamate)

A solution of ethyl diphenyl phosphoacetate (0.16 g, 0.50 mmol) in THF (30 ml)

was treated with Triton B (0.27 ml, 0.60 mmol) at -78 °C for 15 min. A solution of p-

sulfonated benzaldehyde (0.12 g, 0.60 mmol in 5 ml MeOH) was then added dropwise,

and the resulting mixture was stirred at -78 °C for 1 hr. The reaction was quenched with 5

ml of D.I water. ESI mass spectra of the crude product indicated a mixture of m/z 255

(the sulfonated ethyl cinnamate) and m/z 249, which is assigned to a (PhO)2PO2 by-

product. NMR of the crude mixture in D2O contained doublets at 6.0 and 6.98 ppm (J =

12.4 Hz), which could be assigned to the cis-cinnamate based on comparison with

previously reported cis-structures. In contrast, the clearly observed NMR peak for the

authentic trans isomer is a doublet at 6.49 ppm with J = 15.4. From the NMR spectrum

of the cis-sample, we estimate a cis/trans ratio of about 98:2, consistent with the

selectivities obtained previously using this procedure. The ethyl ester was introduced

into the mass spectrometer by electrospray and converted to the carboxylic acid by CID,

and elimination of ethylene.

Page 94: Mass spectrometric characterization of remotely charged

76

5.3.2 4-(acryloyloxy)benzenesulfonate (PASO3)

4-(acryloyloxy)benzenesulfonate was prepared by mixing 1 mL of 4-

hydroxybenzenesulfonic acid sodium salt dihydrate with 3 mL of acryloyl chloride

(CH2=CHCOCl). The neat mixture was stirred for about 15 minutes at 35º, and 1 mL of

(i-Pr)2Net was added. The solution was stirred approximately 15 minutes, after which

one drop of the crude mixture was added to 1 mL of a 50/50 mixture of water and

methanol and examined by ESI mass spec. The main products observed in the mass

spectrum of the crude mixture are the hydroxybenzenesulfonate, a product with m/z 253,

which is likely HO3SC6H4SO4, and a product at m/z 227, which is the phenyl acrylate.

5.4 Computational Methods

Amino acid conformations were surveyed by using the Monte Carlo Multiple

Minimum search method, with the MMFFs force field, as implemented in MacroModel

(version 9.6).37 Multiple starting conformations were utilized, including canonical and

zwitterionic structures, to improve the chances of finding the key low energy structures.

All of the unique conformations calculated with the MMFFs force field to be within 100

kJ/mol of the lowest energy structure were identified and used as starting points for

B3LYP/6-31+G* geometry optimizations.

Stationary points were confirmed to be minima or saddle points from the

computed vibrational frequencies. Reported energies are electronic energies, computed

at the MP2/6-311+G** level of theory at the B3LYP geometries, designated MP2/6-

311+G**//B3LYP/6-31+G*, unless noted, and are not corrected for zero-point energies

or thermal energy contributions. The larger basis set is used to provide a good

Page 95: Mass spectrometric characterization of remotely charged

77

description of the van der Waals interactions and polarization of charge. Calculations

were carried out using QCHEM version 4.038 and Gaussian.39

5.5 Results

The mass spectrum of an aqueous solution of PheSO3, shown in Figure 5.1a, is

very clean, with essentially only peaks corresponding to PheSO3. Considering the large

difference in a acidities of sulfonic and carboxylic acids,40 it is not surprising that the

sulfonate structure, depicted as PheSO3, is computed (B3LYP/6-31+G*) to be more

than 20 kcal/mol more stable than the carboxylate structure, similar to what has been

reported previously for sulfated peptides.41 Although mixtures of ion isomers have been

found for amino acid anions when the acidities of the different groups are similar,42,43 the

large difference in the acidities of the carboxylic and sulfonic acids should result in an

ion that consists of a sulfonate charge with separate amino acid moiety. Although the

extent of interaction between the charge and the reactive group is always a question in

these studies, the distance to the charge is too great for any significant interaction to

occur. In addition, the saturated carbons that separate the amino acid from the ring

prevent any conjugative interactions with the charge.

Page 96: Mass spectrometric characterization of remotely charged

78

Figure 5.1. a) ESI and b) CID mass spectra of PheSO3. The CID for spectrum was

carried out with a normalized collision energy of 25%.

Calculations have been carried out to determine the structure of PheSO3

in the

gas phase. A conformational search identified more than 20 unique low-energy

conformations (within 50 kJ/mol of minimum) of the ion, and all but one which have

canonical structures. Qualitative depictions of the 5 lowest energy canonical

conformations and the lowest energy zwitterionic structure are shown in Figure 5.2.

Figure 5.2 also shows the relative energies of each stationary point, optimized at the

B3LYP/6-31+G* level of theory. The most stable conformation has the carboxylic acid

moiety anti to the aromatic ring, with a hydrogen bond between the –OH proton and the

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

150 170 190 210 230 250 270 290

M

M-NH3

M-NH3-COM-NH3

-CO2

a

b

Page 97: Mass spectrometric characterization of remotely charged

79

amine nitrogen. Indeed, the anti-conformation, when possible, is generally preferred for

the structures examined in this work. The conformation with the –COOH and aromatic

ring gauche is higher in energy by about 2.5 kcal/mol. Alternate hydrogen bonding

arrangements, such as NH- or OH to carbonyl oxygen, or NH to hydroxyl hydrogen, are

higher in energy by more than 4 kcal/mol. The most important conclusion from this

computational survey is that the lowest energy zwitterionic structures are significantly

higher in energy than the lowest canonical forms. Figure 5.2 also shows that the relative

energies are similar whether using either DFT or ab initio methods.

Figure 5.2. Low-energy conformations of the amino acid moiety of PheSO3 and their

relative energies, in kcal/mol, computed at the B3LYP/6-31+G* and MP2/6-311+G** (in parentheses) level of theory.

5.5.1 Dissociation of PheSO3

The CID spectrum of PheSO3 is shown in Figure 5.1b. The major fragments

observed are readily assigned as resulting from dissociation of the canonical amino acid

backbone. In particular, the main dissociation processes all involve loss of NH3 to form

an ion with m/z 227, which subsequently fragments by loss of either CO, to form m/z

199, or CO2, to form m/z 183. That the m/z 183 and 199 ions come from the M-NH3

Page 98: Mass spectrometric characterization of remotely charged

80

anion is confirmed by two additional results. First, although m/z 183 and 199 are major

products observed at the energy used for CID in Figure 5.1b, the M-NH3 ion is the only

product observed at very low energies, consistent with a sequential dissociation. Second,

CID of the isolated M-NH3 anion, generated either by CID during the ESI injection stage,

or by CID in the course of an MS/MS/MS experiment, verifies that it dissociates by loss

of CO and CO2 in a nearly 1:1 ratio as observed in Figure 5.1b. Therefore, the loss of

NH3 is the only pathway that can be uniquely assigned to PheSO3. Direct loss of CO2,

which has been found previously for pyrolysis of gas-phase amino acids (eq 5. 1),20 is not

observed upon activation of PheSO3.

Although the structure of the M-NH3 product is not known, the most likely

products to consider are those that are stable products formed by minimal rearrangement.

Therefore, the most likely products are either a cis- or trans-cinnamic acid (CASO3), an

α-lactone (αLSO3) or the phenyl acrylate (PASO3

). Other products are possible but

would require more extensive rearrangement.

CID of deprotonated phenyl alanine is known to result in formation of cinnamic

acid.44-46 Similarly, dissociation of PheSO3 by elimination across the ethylene backbone

would result in the formation of a cinnamic acid, CASO3, as shown in eq 5. 4. However,

multiple lines of evidence argue against the cinnamic acid. First, ammonia loss from d3-

labeled PheSO3, formed by H/D exchange using a mixture of D2O/CD3CO2D as the

solvent, occurs by loss of ND3 only, and loss of ND2H, NDH2 or NH3 is not observed to

any detectable extent at any collision energy.

Page 99: Mass spectrometric characterization of remotely charged

81

This rules out the possibility rules out elimination across the backbone as shown

in eq 5.4. It is also evidence against other, more complicated rearrangement processes

involving non-amino acid portions of the molecule. However, cinnamic acid formation

from PheSO3 does not require dissociation across the backbone. An alternate

mechanism, as shown in eq 5.5, proceeds via a zwitterionic intermediate.

Page 100: Mass spectrometric characterization of remotely charged

82

Therefore, we have examined the CID behavior of authentic trans- and cis-

cinnamic acids to compare with the product obtained from PheSO3. For this

experiment, sulfonated trans-cinnamic acid was formed by ESI of an aqueous solution of

the commercially available sulfonyl chloride, and the cis-cinnamic acid was formed in

situ by CID of the corresponding ethyl ester. A comparison of the CID spectra of the m/z

227 product obtained from PheSO3 and the authentic cinnamic acids, measured with a

normalized collision energy of 34%, is shown in Figure 5.3. Whereas all ions lose CO2

as the major fragmentation pathway, significant differences are evident. For example, the

CASO3 ions do not lose CO to any appreciable extent, and lose only SO2 (to form m/z

163) in addition to losing CO2. Moreover, although the m/z 183 ion that corresponds to

PheSO3-NH3-CO2 is found to lose SO2 upon CID (vide infra), the CASO3

ions

undergo facile loss of CO2 and SO2 to form m/z 119 in much higher yield than is found

for the M-NH3 ion. Finally, although the authentic CASO3 ions dissociate by loss of

CO2, they do so only at energies much higher than those needed for dissociation of

PheSO3. The amounts of the m/z 163 and m/z 119 products observed in Figures 5.3a,

5.3b and 5.3c indicate that CASO3 constitutes at most about 5% of the M-NH3 product.

Page 101: Mass spectrometric characterization of remotely charged

83

Figure 5.3. CID spectra of m/z 227 ions obtained a) by CID of PheSO3, and from

sulfonated b) cis and c) trans cinnamic acid

As shown in eq 5.6, the zwitterionic intermediate could also rearrange through

nucleophilic attack at the ipso-position followed by a Grob-like fragmentation to give

PASO3.

Page 102: Mass spectrometric characterization of remotely charged

84

NH3

O

O

+

_

SO3_

SO3_

O

+ NH3 (5.6)

O

PASO3_

Although PASO3 is likely a relatively stable product, it can be ruled out for

multiple reasons. First, it is inconsistent with the CID results, in that it cannot readily

lose either CO or CO2. Second, formation of the phenyl acrylate would require a high-

energy, non-aromatic transition state (vide infra). Ultimately, the formation of PASO3

was ruled out by using an authentic sample. The primary dissociation pathway of

PASO3 is loss of the acryloyl group (CH2=CHCO) to form the phenoxyl radical, loss of

CO or CO2 does not occur at all. In contrast, formation of the α-lactone is consistent with

all of the data. Formation of the lactone, either by a concerted loss of NH2H or through

a zwitterionic intermediate (eq 5.7) involves loss of only the exchangeable protons, and is

therefore consistent with the results for the deuterated ions.

Page 103: Mass spectrometric characterization of remotely charged

85

Loss of CO like that observed here is a well-known decomposition pathway for α-

lactones,24-36 and therefore loss of CO from the M-NH3 anion is consistent with the

presence of the α-lactone. Although decarboxylation is generally not considered a

decomposition pathway for that type of structure, it is not unprecedented,47 and

computational studies described herein predict that it is expected to occur for αLSO3.

Based on these results, we conclude that the M-NH3 ion found upon CID of

PheSO3 is most likely the α-lactone, αLSO3

. Although α-lactones are nominally high-

energy reactive species, they have previously been proposed as products in the CID of -

substituted carboxylates in mass spectrometry experiments,48-51 which is essentially how

it is shown being formed via the zwitterion in eq 5.7. The difference between the

previous studies and this work is that the α-lactone formed upon charge remote

fragmentation of PheSO3 is ionic, and therefore detectable by mass spectrometry.

Page 104: Mass spectrometric characterization of remotely charged

86

Previous studies of halogenated carboxylates 48-51 have inferred α-lactone structures

based on the observation of halide products.

5.5.2 Dissociation of the α-Lactone, αLSO3

As described in the section above, the PheSO3 - NH3 product, assigned to be the

α-lactone (αLSO3), dissociates by loss of either CO or CO2. Loss of CO from α-

lactones is well-known and predicted to occur with a barrier of about 28-30 kcal/mol52 to

form the corresponding aldehyde or ketone (eq 5.3). The CO-loss product obtained upon

CID of αLSO3 in this work is found to undergo further dissociation by loss of 29 mass

units, which we assign to HCO as shown in eq 5.8.

Dissociation to form the benzyl radical product, BzSO3, is also observed upon

CID of control substrates, such as sulfonated phenylpropionic acid or phenethylamine (eq

5. 9), and so it is not surprising that it is also observed for the aldehyde.

Page 105: Mass spectrometric characterization of remotely charged

87

The loss of CO2 has not been reported previously for the gas-phase decomposition of α-

lactones, although it has been found to occur, along with decarbonylation, upon

photolysis of the bis-n-butyl-substituted α-lactone in solution,47 forming the olefin (eq

5.10).

By comparing CID spectra of the m/z 183 product with an authentic sample, we

have confirmed that the product formed upon decarboxylation of αLSO3 is similarly the

corresponding styrene (eq 5.11).

Page 106: Mass spectrometric characterization of remotely charged

88

In summary, we have found that the dissociation of the gas-phase amino acid,

PheSO3, results in formation of an α-lactone, αLSO3

, which dissociates by

decarbonylation to form the aldehyde, and by decarboxylation accompanied by hydrogen

shift to form the styrene.

5.6 Discussion

The results described in the previous section are the first studies of the

dissociation of a simple (non-N-substituted), canonical amino acid in the gas phase. The

dissociation pathway for the amino acid is similar to that observed previously for

substituted alanine23 and other substituted carboxylic acids,24-36 but is very different from

what has been reported for glycine derivatives.20-22 The difference in dissociation

pathways between glycine and alanine derivatives has been noted previously22 but has not

been explained satisfactorily.

That the decomposition of PheSO3 involves only the amino acid moiety and not

the SO3 group is evidence that the sulfonate charge does not participate in the

dissociation processes. Fragmentation of the SO3 group is not even observed for any of

the dissociation products, except for the sulfonated styrene shown in eq 5.11, which

dissociates by loss of SO2. Therefore, the experimental results indicate that the sulfonate

group is serving extensively as a charge carrier, and does not significantly affect the

possible dissociation pathways. Although the dissociation pathway, loss of NH3, is the

same as what is found for deprotonated phenylalanine,44-46 the product formed with that

ion is deprotonated cinnamic acid. Control experiments in this work show that cinnamic

acid is not formed upon dissociation of PheSO3. The mechanism proposed for

Page 107: Mass spectrometric characterization of remotely charged

89

deprotonated phenylalanine44-46 involves intramolecular proton transfer to the carboxylate

and subsequent proton transfer to the amine, and therefore the charge center is

extensively involved in the process. The lack of formation of cinnamic acid with

PheSO3 suggests the charge center is not participating in the dissociation process.

Deprotonated tyrosine also undergoes CID by loss of ammonia,53 possibly via the

phenoxide ion.

Electronic structure calculations predict that the sulfonate charge does not have a

large effect on the energies of the reaction pathways for PheSO3. A comparison of the

relative energies of observed and other possible products in the dissociation of PheSO3

and neutral phenylalanine, Phe, calculated at the MP2/6-311+G**//B3LYP/6-31+G*

level of theory, are shown in Table 5.1. Table 5.1 shows that the relative energies for all

of the possible dissociation products for PheSO3 and Phe are very similar, with

differences for the primary process of less than 3 kcal/mol.

Table 5.1. Computed Reaction Energies for Dissociation Pathways of PheSO3 and Phea

Relative Energies Products PheSO3

Phe Amino acid (reactant) 0.0 0.0 ArCH2C(CO)CH + NH3 (α-lactone, eq 5. 7) 47.0 44.6 ArCH=CHCO2H + NH3 (trans cinnamic acid, eq 5. 4)

17.7 16.4

ArCH2CH2NH2 + CO2 1.1 -4.6 ArCH2CHO + CO + NH3 34.9 31.9 ArCH=CH2 + CO2 + NH3 15.9 11.4 ArOC(O)CH=CH2 + NH2 39.4 39.0

aElectronic energy differences between products and the amino acid reactants, computed at the MP2/6-311+G**//B3LYP/6-31+G* level of theory.

Page 108: Mass spectrometric characterization of remotely charged

90

5.6.1 Amino Acid Dissociation

A potential energy surface for the dissociation of PheSO3, shown in Figure 5.4,

provides insight into the product selectivity. Because CID is a kinetic process, the most

important features of the potential energy surface are the activation barriers. Therefore,

we have calculated the barriers for NH3 and CO2 loss from PheSO3 and Phe at the

MP2/6-311+G**//B3LYP/6-31+G* level of theory. Despite multiple attempts, we could

not find any transition states that directly connect the canonical structure of the amino

acid with the expected deamination or decarboxylation products. However, insight into

the mechanism was obtained from multidimensional potential energy surface

calculations. We find that loss of NH3 or CO2 occurs effectively by dissociation of the

zwitterionic structure, in that proton transfer from the carboxylic acid to the amino-group

occurs very early in the process at energies well below those required for dissociation.

Page 109: Mass spectrometric characterization of remotely charged

91

Figure 5.4. Calculated energetics for the dissociation of PheSO3

(X = SO3) and Phe (X

= H). Values correspond to electronic energies and geometries calculated at the MP2/6-311+G**//B3LYP/6-31+G* level of theory a) not a stable geometry; corresponds to an

amino acid geometry the same as that for zwitterionic PheSO3; b) computed utilizing a 3

kcal/mol barrier for the reverse reaction, obtained from a relaxed surface scan; see Supporting Information.

Loss of NH3 from Phe to form the α-lactone is estimated to have a barrier of about

3 kcal/mol higher than the reaction energy, about 46 kcal/mol. The mechanism for

ammonia elimination involves an intramolecular SN2 reaction (eq 5.12), where the

carboxylate displaces the amine leaving group in the zwitterionic structure. Therefore,

the dissociation of PheSO3 to form the α-lactone is analogous to the elimination

reactions of α-halocarboxylates that have been reported previously.50,54-60 In an ab initio

Page 110: Mass spectrometric characterization of remotely charged

92

study, Davidson and co-workers61 found that there is no barrier in excess of the

dissociation energy for formation of the α-lactone from chloroacetate, whereas a small

barrier was calculated for the zwitterionic structure of Phe in this work.

The barrier for elimination of NH3 to form the cinnamic acid, not shown on Figure 5.4, is

calculated to be 68 kcal/mol, which accounts for why that process does not occur.

However, given that charge remote fragmentation is analogous to pyrolysis, that

cinnamic acid is not formed is not surprising considering that pyrolysis of alkylated

amines does not occur by concerted elimination of NH3, but results in the formation of

imines.62 Similarly, we have calculated the barrier for formation of the phenyl acrylate,

PASO3, as shown in eq 5. 6. Whereas for the neutral phenyl alanine we find a concerted

transition state for rearrangement accompanied by ammonia loss, the transition state for

the sulfonated version has the ammonia completely dissociated, and is part of a step-wise

process. In either case, the barrier for phenyl acrylate formation is very high, computed

to be about 80 kcal/mol from the zwitterion, and 90 - 100 kcal/mol from the canonical

structure.

Although loss of CO2 from the amino acid is not observed in this work, it has

been reported previously, and therefore we have carried out calculations to determine

Page 111: Mass spectrometric characterization of remotely charged

93

why it does not occur for PheSO3. Decarboxylation of the zwitterionic structure of Phe

is calculated to occur without a barrier in excess of the dissociation energy, such that the

net barrier for the reaction is determined by the energies of the dissociation products.

Therefore, loss of CO2 has a barrier of more than 60 kcal/mol. As shown in Figure 5.4,

loss of CO2 would necessarily lead to formation of the zwitterionic product (the

ammonium-ylide) and not the phenethyl amine. Although formation of the amine is

exothermic, Alexandrova and Jorgensen63 have noted that the barrier for proton shift in

the ylide is formally symmetry forbidden and is therefore expected to have a very high

barrier. Consequently, the formation of the amine from glycine is essentially a

sequential, two-step process consisting of decarboxylation followed by proton transfer,

with a proton transfer barrier of approximately 45 kcal/mol for the ylide in aqueous

solution.63 The barrier for proton shift in the ammonium-ylide formed upon

decarboxylation of Phe in the gas phase is calculated to be 25.6 kcal/mol, such that the

overall barrier for formation of the amine upon decarboxylation of Phe is calculated to be

86.4 kcal/mol. The higher barrier for proton shift computed for glycine is likely due to

solvent stabilization of the ammonium-ylide. Interestingly, Alexandrova and Jorgensen

also calculated a slight (~8 kcal/mol) barrier in excess of the dissociation energy for the

decarboxylation step. However, the presence of the excess barrier is also likely a

consequence of explicitly including solvation in the QM/MM approach, and would not be

expected in a gas-phase calculation.64 The barriers for decarboxylation obtained for

glycine by Alexandrova and Jorgensen63 and for Phe in this work are significantly higher

than what has been reported previously for the decarboxylation of canonical N,N-

dimethyl- or N-phenylglycine.20,21,65 However, the ~42 kcal/mol barriers suggested in the

Page 112: Mass spectrometric characterization of remotely charged

94

previously studies do not seem reasonable because they are lower than the energies

required for decarboxylation, which we calculate to be 52.3 kcal/mol and 59.6 kcal/mol

for dimethylglycine and N-phenylglycine, respectively. Considering that decarboxylation

leads initially to the ammonium-ylide, and that rearrangement of the ylide to the amine

occurs in a second step and is formally symmetry forbidden, decarboxylation should not

occur with an activation energy less than that required for formation of the ammonium-

ylide.

In summary, although loss of CO2 from the amino acid can occur to form the

ammonium-ylide, formation of the α-lactone is predicted to be energetically preferred by

about 20 kcal/mol. Consequently, that is the only process that is observed for PheSO3.

5.6.2 Dissociation of the α-Lactone

In this study, we have also been able to investigate the dissociation of the α-

lactone, αLSO3. α-Lactones are highly reactive molecules that have been proposed as

short lived intermediates in a variety of chemical reactions, often involving α-substituted

carboxylates50,54-60 or α,β-unsaturated acids,66,67 in the photolysis of cyclic peroxides47,68-

71 and α-halocarboxylic acids,72 in the oxidation of ketenes,73 or in the gas-phase

dissociation of α-substituted carboxylic24-29 or in the decomposition of amino acids.23,74

They have also been proposed as intermediates in atmospheric oxidation reactions75,76 in

mass spectrometry,77-80 and as products of the reaction of carbenes with CO2.81-85 Despite

being highly reactive, α-lactones have been investigated spectroscopically by using

matrix isolation,82-84 gas-phase86 and solution-phase time-resolved IR.85 There has long

been discussion regarding the electronic structure of α-lactones, and whether they are best

Page 113: Mass spectrometric characterization of remotely charged

95

considered to be cyclic, canonical structures56,58-60 or zwitterionic,55 although those

discussions generally relate to the structure in solution. The calculated structures of gas-

phase α-lactones have exceptionally long (~1.55 Å) and weak87 Cα-O bonds, and even in

the gas phase there is considerable ionic character.87

α-Lactones are highly reactive, and react rapidly by polymerization, by

nucleophilic addition, or by decomposition.88 Although an early photolysis study

reported dissociation by loss of CO and CO2,47 more recent studies of α-substituted

carboxylic acid pyrolyses that proceed through α-lactone intermediates24,25,30-36 have

reported dissociation only by loss of CO. In this work, we observe loss of both CO and

CO2 from the α-lactone, in similar amounts (if anything, CO2 loss is slightly favored). As

described above, loss of CO occurs to form the aldehyde, as expected, whereas loss of

CO2 results in the formation of the styrene (eq 5.10).

Computed potential energy surfaces for the decomposition channels are shown in

Figure 5.5. The barriers for CO loss from the α-lactone derived from PheSO3 and Phe

are predicted to be about 30 kcal/mol, similar to that predicted for alkyl-substituted α-

lactones.52 Determining the barrier for styrene formation is more difficult. The

experimental observation that CO and CO2 are formed in similar indicates that the

barriers for the two processes are similar. Therefore, a stepwise mechanism consisting of

decarboxylation to form the carbene followed by hydrogen shift likely has a barrier that is

too high to compete with CO loss (see Figure 5.5). Considering that decarboxylation of

β-lactones results directly in the formation of the olefin and occurs with moderate energy

barriers89,90 we explored the possibility of a mechanism involving the dyotropic shift91-94

to the β-lactone95 followed by decarboxylation (eq 5. 13a). However, the search for the

Page 114: Mass spectrometric characterization of remotely charged

96

transition state for α- to β-lactone rearrangement did not give a structure for a dyotropic

process, as shown in 10a, but instead gave a structure that consists of 1,2-hydrogen shift

without assistance of the carboxylate anion, as shown in eq 5.13b. The structure shown

in eq 5.13b was confirmed to be a saddle point by vibrational analysis, and an intrinsic

reaction coordinate calculation (IRC)96,97 finds that it is a transition state that connects the

α-lactone with the styrene and CO2.

Figure 5.5. Calculated energetics for the dissociation of αLSO3 (X = SO3

) and the non-sulfonated derivative (X = H). Values correspond to electronic energies and geometries calculated at the MP2/6-311+G**//B3LYP/6-31+G* level of theory, unless indicated a)

not a stable geometry; corresponds to the geometry of the lactone with the CO2 removed.

Page 115: Mass spectrometric characterization of remotely charged

97

The 1,2-hydrogen shift mechanism shown in eq 5.13b is facilitated by the high

degree of polarization expected for the α-lactone (eq 5.14),87 and by the fact that the β-

cationic carboxylate (β(+)-CO2) that would result from hydride transfer is unstable with

respect to decarboxylation in the gas phase. The coupling of hydrogen migration and

leaving-group loss is analogous to what has been proposed for the Schmidt reaction,94

and has been observed previously in reactions such as the elimination of protonated

ethers98,99 and acetates.100

The computed barriers for decarboxylation of the α-lactone are 32.1 kcal/mol for

αLSO3, and 42.1 kcal/mol for the non-ionic system, indicating that the sulfonate is

Page 116: Mass spectrometric characterization of remotely charged

98

having a large effect on the stability of the transition state. Because the charge of the

sulfonate group does not delocalize into the π-system of the ring via conjugation, the

stabilization of the transition state is likely an inductive effect that stabilizes the

formation of the benzylic cation. The effect would be expected to be stronger for an

ionic group that is conjugated with the aromatic ring, such as an O group of a phenoxide

(eq 5.15), which accounts for why the phenoxide ion obtained from deprotonating

tyrosine loses NH3 and CO2 as the main dissociation channel, and an M-NH3-CO ion is

not observed at all.53

OOH

H H

OOH

H H+

_

(5.15)

HH

CO2

_

H

O O_ _

O

The calculated barriers for decarboxylation and decarbonylation of the ionic and

non-ionic substrates are consistent with the experimentally observed results. The large

difference in barrier heights for the non-ionic lactone accounts for why loss of CO is

generally the only observed channel in pyrolysis experiments. However, for αLSO3, the

computed barriers for loss of CO and CO2 are very similar, which accounts for why the

products are observed in nearly equal amounts. As shown in eq 5.15, the resonant

interaction between the charge site and the benzylic site in deprotonated tyrosine is

expected to strongly favor the CO2 loss channel. Finally, preliminary results with para-

Page 117: Mass spectrometric characterization of remotely charged

99

trimethylammonium-substituted Phe find that the corresponding α-lactone dissociates

only by loss of CO, consistent with electrostatic destabilization of the benzylic cation.

5.7 Conclusions

By using an ion with the charged-moiety isolated from an amino acid, we have

been able to utilize mass spectrometry to investigate the chemical properties of a gas-

phase amino acid. In this work, we have characterized the decomposition pathways for a

phenylalanine derivative. The only observed pathway is loss of ammonia, to form an α-

lactone. This reaction is similar to what has been observed previously for pyrolysis of N-

substituted alanines23 and other α-substituted carboxylic acids,24-36 but is inconsistent

with what has been reported for pyrolysis of glycine derivatives.20-22 The main difference

between the studies that find deamination23 (including this work) and those that find

decarboxylation to occur20-22 is our work and that of Al-Awadi et al.23 involves a direct

investigation of the gaseous amino acid, whereas Chuchani and co-workers have tried to

generate the amino acid in situ by pyrolysis of the corresponding ethyl esters.

Considering that the computational results in this work and by others63 have shown that

decarboxylation of gaseous amino acids to form the amine has a prohibitively high

barrier, the products that correspond to decarboxylation of the amine are likely formed

via a pathway not involving the amino acid.

The α-lactone that is formed from the amino acid in this work undergoes

dissociation by loss of CO and by loss of CO2. Although loss of CO is expected, the loss

of CO2 has generally not been reported in reactions involving α-lactones. However,

given the calculated relative barriers for CO and CO2 loss, decarboxylation should occur,

Page 118: Mass spectrometric characterization of remotely charged

100

to some extent, in α-lactones containing a β-hydrogen that can shift during lactone ring

opening, similar to the reaction shown in eq 5.13b. It may be that the formation of the

benzylic cation may facilitate the hydride shift reaction, although decarboxylation has

been observed for a completely aliphatic derivative (eq 5.9).47 The results of this work

and that reported previously for deprotonated tyrosine show that stabilization of the β-

cation can affect the branching ratio for CO vs CO2 loss. Therefore, just as anionic

stabilization of the transition state favors CO2, the presence of a cationic group would be

expected to disfavor decarboxylation, and favor loss of CO, which is what is observed in

preliminary studies.

Finally, the results of this study show that mass spectrometry can be used to

investigate the properties of canonical amino acids. Whereas this is a new approach for

the investigation of gas-phase amino acids, the use of an inert, remote charge to

investigate neutral chemistry is not new and is similar, in principle, to the distonic ion

approach used by Kenttämaa and co-workers101 to examine the reactivity of aromatic

radicals. In the same way, it should be possible to use mass spectrometry to investigate

further the reactivity of gas-phase amino acids, and investigate the effects of structure and

solvation.

Page 119: Mass spectrometric characterization of remotely charged

101

5.8 References

(1) Kuan, Y.-J.; Charnley, S. B.; Huang, H.-C.; Tseng, W.-L.; Kisiel, Z. Astrophys. J. 2003, 593, 848.

(2) Snyder, L. E.; Lovas, F. J.; Hollis, J. M.; Friedel, D. N.; Jewell, P. R.; Remijan, A.; Ilyushin, V. V.; Alekseev, E. A.; Dyubko, S. F. Astrophys. J. 2005, 619, 914.

(3) Johansson, A. C. V.; Lindahl, E. J. Chem. Phys. 2009, 130, 185101.

(4) Chapo, C. J.; Paul, J. B.; Provencal, R. A.; Roth, K.; Saykally, R. J. J. Am. Chem. Soc 1998, 120, 12956.

(5) Julian, R. R.; Jarrold, M. F. J. Phys. Chem. A 2004, 108, 10861.

(6) Jensen, J. H.; Gordon, M. S. J. Am. Chem. Soc. 1995, 117, 8159.

(7) Kassab, E.; Langlet, J.; Evleth, E.; Akacem, Y. J. Mol. Struct. 2000, 531, 267.

(8) Jarrold, M. F. Annu. Rev. Phys. Chem. 2000, 51, 179.

(9) Zwier, T. S. J. Phys. Chem. A 2006, 110, 4133.

(10) Bakker, J. M.; Aleese, L. M.; Meijer, G.; Helden, G. v. Phys. Rev. Lett. 2003, 91, 203003.

(11) Schermann, J.-P. Spectroscopy and Modelling of the Biomolecular Building Blocks; Elsevier: Netherland, 2008.

(12) Hu, Y.; Bernstein, E. R. J. Chem. Phys. 2008, 128, 164311/1.

(13) Hu, Y.; Bernstein, E. R. J. Phys. Chem. A 2009, 113, 8454.

(14) Antoine, R.; Dugourd, P. Phys. Chem. Chem. Phys. 2011, 13, 16494.

(15) Gao, Y. K.; Traeger, F.; Kotsis, K.; Staemmler, V. Phys. Chem. Chem. Phys. 2011, 13, 10709.

(16) Jochims, H.-W.; Schwell, M.; Chotin, J.-L.; Clemino, M.; Dulieu, F.; Baumgartel, H.; Leach, S. Chem. Phys. 2004, 298, 279.

(17) Adams, J. Mass Spectrom. Rev. 1990, 9, 141.

(18) Cheng, C.; Gross, M. L. Mass Spectrom. Rev. 2000, 19, 398.

(19) Adams, J.; Gross, M. L. J. Am. Chem. Soc. 1989, 111, 435.

Page 120: Mass spectrometric characterization of remotely charged

102

(20) Ensuncho, A.; Lafont, M. J.; Rotinov, A.; Dominguez, R. M.; Herize, A.; Quijano, J.; Chuchani, G. Int. J. Chem. Kinet. 2001, 33, 465.

(21) Dominguez, R. M.; Tosta, M.; Chuchani, G. J. Phys. Org. Chem. 2003, 16, 869.

(22) Tosta, M.; Oliveros, J. C.; Mora, J. R.; Cordova, T.; Chuchani, G. J. Phys. Chem. A 2010, 114, 2483.

(23) Al-Awadi, S. A.; Abdallah, M. R.; Hasan, M. A.; Al-Awadi, N. A. Tetrahedron 2004, 60, 3045.

(24) Al-Awadi, N. A.; Kumar, A.; Chuchani, G.; Herize, A. Int. J. Chem. Kinet. 2001, 33, 612.

(25) Chuchani, G.; Dominguez, R. M.; Rotinov, A.; Martin, I. J. Phys. Org. Chem. 1999, 12, 612.

(26) Domingo, L. R.; Picher, M. T.; Andres, J.; Safont, V. S.; Chuchani, G. Chem. Phys. Lett. 1997, 274, 422.

(27) Domingo, L. R.; Picher, M. T.; Safont, V. S.; Andres, J.; Chuchani, G. J. Phys. Chem. A 1999, 103, 3935.

(28) Rotinov, A.; Chuchani, G.; Andre, J.; Domingo, L. R.; Safont, V. S. Chem. Phys. 1999, 246, 1.

(29) Safont, V. S.; Moliner, V.; Andres, J.; Domingo, L. R. J. Phys. Chem. A 1997, 101, 1859.

(30) Chuchani, G.; Dominguez, R. M. Int. J. Chem. Kinet. 1995, 27, 85.

(31) Chuchani, G.; Dominguez, R. M. Int. J. Chem. Kinet. 1999, 31, 725.

(32) Chuchani, G.; Dominguez, R. M.; Rotinov, A. Int. J. Chem. Kinet. 1991, 23, 779.

(33) Chuchani, G.; Martin, I.; Rotinov, A. Int. J. Chem. Kinet. 1995, 27, 849.

(34) Chuchani, G.; Martin, I.; Rotinov, A.; Dominguez, R. M. J. Phys. Org. Chem. 1993, 6, 54.

(35) Chuchani, G.; Rotinov, A. Int. J. Chem. Kinet. 1989, 21, 367.

(36) Chuchani, G.; Rotinov, A.; Dominguez, R. M. J. Phys. Org. Chem. 1996, 9, 787.

Page 121: Mass spectrometric characterization of remotely charged

103

(37) Version 9.6 ed.; Schrödinger, LLC: New York, NY, 2009.

(38) Kong, J.; White, C. A.; Krylov, A. I.; Sherrill, D.; Adamson, R. D.; Furlani, T. R.; Lee, M. S.; Lee, A. M.; Gwaltney, S. R.; Adams, T. R.; Ochsenfeld, C.; Gilbert, A. T. B.; Kedziora, G. S.; Rassolov, V. A.; Maurice, D. R.; Nair, N.; Shao, Y.; Besley, N. A.; Maslen, P. E.; Dombroski, J. P.; Daschel, H.; Zhang, W.; Korambath, P. P.; Baker, J.; Byrd, E. F. C.; Voorhis, T. V.; Oumi, M.; Hirata, S.; Hsu, C.-P.; N.Ishikawa; Florian, J.; A.Warshel; Johnson, B. G.; Gill, P. M. W.; M.Head-Gordon; Pople, J. A. J. Comput. Chem. 2000, 21, 1532.

(39) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; G. E. Scuseria; Robb, M. A.; Cheeseman, J. R.; J. A. Montgomery, J.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.; Gaussian, Inc.: Pittsburgh, PA, 2003.

(40) Bartmess, J. E. 2012.

(41) Nha Tran, T. T.; Wang, T.; Hack, S.; Bowie, J. H. Rapid Commun. Mass Spectrom. 2013, 27, 1135.

(42) Tian, Z.; Pawlow, A.; Poutsma, J. C.; Kass, S. R. J. Am. Chem. Soc 2007, 129, 5403.

(43) Tian, Z.; Wang, X.-B.; Wang, L.-S.; Kass, S. R. J. Am. Chem. Soc. 2009, 131, 1174.

(44) Eckersley, M.; Bowie, J. H.; Hayes, R. N. Int. J. Mass Spectrom. Ion Processes 1989, 93, 199.

(45) Choi, S.-S.; Kim, O.-B. Int. J. Mass Spectrom. 2013, 338, 17.

(46) Couldwell, A. M.; Thomas, M. C.; Mitchell, T. W.; Hulbert, A. J.; Blanksby, S. J. Rapid Commun. Mass Spectrom. 2005, 19, 2295.

(47) Adam, W.; Rucktaeschel, R. J. Org. Chem. 1978, 43, 3886.

(48) Baker, M.; Gabryelski, W. Int. J. Mass Spectrom. 2007, 262, 128.

Page 122: Mass spectrometric characterization of remotely charged

104

(49) Choi, T. S.; Ko, J. Y.; Heo, S. W.; Ko, Y. H.; Kim, K.; Kim, H. I. J. Am. Soc. Mass Spectrom. 2012, 23, 1786.

(50) Graul, S. T.; Squires, R. R. Int. J. Mass Spectrom. Ion Processes 1990, 100, 785.

(51) Rijs, N. J.; O'Hair, R. A. J. Dalton Trans. 2012, 41, 3395.

(52) Domingo, L. R.; Andres, J.; Moliner, V.; Safont, V. S. J. Am. Chem. Soc. 1997, 119, 6415.

(53) Tian, Z.; Kass, S. R. J. Am. Chem. Soc. 2008, 130, 10842.

(54) Bean, C. M.; Kenyon, J.; Phillips, H. J. Chem. Soc. 1936, 303.

(55) Cowdrey, W. A.; Hughes, E. D.; Ingold, C. K.; Masterman, S.; Scott, A. D. J. Chem. Soc. 1937, 1252.

(56) Grunwald, E.; Winstein, S. J. Am. Chem. Soc. 1948, 70, 841.

(57) Strijtveen, B.; Kellogg, R. M. Recl. Trav. Chim. Pays-Bas 1987, 106, 539.

(58) Winstein, S. J. Am. Chem. Soc. 1939, 61, 1635.

(59) Winstein, S.; Henderson, R. B. J. Am. Chem. Soc. 1943, 65, 2196.

(60) Winstein, S.; Lucas, H. J. J. Am. Chem. Soc. 1939, 61, 1576.

(61) Antolovic, D.; Shiner, V. J.; Davidson, E. R. J. Am. Chem. Soc. 1988, 110, 1375.

(62) Hamada, Y.; Takeo, H. Appl. Spectrosc. Rev. 1992, 27, 289.

(63) Alexandrova, A. N.; Jorgensen, W. L. J. Phys. Chem. B 2011, 115, 13624.

(64) Alexandrova, A. N., personal communication.

(65) Perez, J. R.; Lorono, M.; Dominguez, R. M.; Cordova, T.; Chuchani, G. J. Phys. Org. Chem. 2008, 21, 402.

(66) Pirinccioglu, N.; Robinson, J. J.; Mahon, M. F.; Buchanan, J. G.; Williams, I. H. Org. Biomol. Chem. 2007, 5, 4001.

(67) Niwayama, S.; Noguchi, H.; Ohno, M.; Kobayashi, S. Tetrahedron Lett. 1993, 34, 665.

(68) Adam, W.; Liu, J.-C.; Rodriguez, O. J. Org. Chem. 1973, 38, 2269.

Page 123: Mass spectrometric characterization of remotely charged

105

(69) Adam, W.; Rucktaeschel, R. J. Amer. Chem. Soc. 1971, 93, 557.

(70) Chapman, O. L.; Wojtkowski, P. W.; Adam, W.; Rodriguez, O.; Rucktaeschel, R. J. Amer. Chem. Soc. 1972, 94, 1365.

(71) Adam, W.; Blancafort, L. J. Org. Chem. 1996, 61, 8432.

(72) Nishino, S.; Nakata, M. J. Mol. Struct. 2008, 875, 520.

(73) Schmittel, M.; von, S. H. Liebigs Ann. 1995, 1815.

(74) Kenyon, J.; Phillips, H. Trans. Faraday Soc. 1930, 26, 451.

(75) Carr, S. A.; Glowacki, D. R.; Liang, C.-H.; Baeza-Romero, M. T.; Blitz, M. A.; Pilling, M. J.; Seakins, P. W. J. Phys. Chem. A 2011, 115, 1069.

(76) Kjaergaard, H. G.; Knap, H. C.; Oernsoe, K. B.; Joergensen, S.; Crounse, J. D.; Paulot, F.; Wennberg, P. O. J. Phys. Chem. A 2012, 116, 5763.

(77) Schroder, D.; Goldberg, N.; Zummack, W.; Schwarz, H.; Poutsma, J. C.; Squires, r. R. Int. J. Mass Spectrom. Ion Processes 1997, 165/166, 71.

(78) Lam, A. K. Y.; O'Hair, R. A. J. Rapid Commun. Mass Spectrom. 2010, 24, 1779.

(79) Wyer, J. A.; Feketeova, L.; Brondsted, N. S.; O'Hair, R. A. J. Phys. Chem. Chem. Phys. 2009, 11, 8752.

(80) Armentrout, P. B.; Ye, S. J.; Gabriel, A.; Moision, R. M. J. Phys. Chem. B 2010, 114, 3938.

(81) Kistiakowsky, G. B.; Sauer, K. J. Am. Chem. Soc. 1958, 80, 1066.

(82) Milligan, D. E.; Jacox, M. E. J. Chem. Phys. 1962, 36, 2911.

(83) Wierlacher, S.; Sander, W.; Liu, M. T. H. J. Org. Chem. 1992, 57, 1051.

(84) Sander, W. W. J. Org. Chem. 1989, 54, 4265.

(85) Showalter, B. M.; Toscano, J. P. J. Phys. Org. Chem. 2004, 17, 743.

(86) Chen, S.-Y.; Lee, Y.-P. J. Chem. Phys. 2010, 132, 114303/1.

(87) Ruggiero, G. D.; Williams, I. H. J. Chem. Soc., Perkin Trans. 2 2001, 733.

(88) L'Abbé, G. Angew. Chem., Int. Ed. Eng. 1980, 19, 276.

Page 124: Mass spectrometric characterization of remotely charged

106

(89) Morao, I.; Lecea, B.; Arrieta, A.; Cossio, F. P. J. Am. Chem. Soc. 1997, 119, 816.

(90) Ocampo, R.; Dolbier, W. R., Jr.; Bartberger, M. D.; Paredes, R. J. Org. Chem. 1997, 62, 109.

(91) Reetz, M. T. Angew. Chem., Int. Ed. Engl. 1972, 11, 130.

(92) Reetz, M. T. Angew. Chem., Int. Ed. Engl. 1972, 11, 129.

(93) Fernandez, I.; Cossio, F. P.; Sierra, M. A. Chem. Rev. (Washington, DC, U. S.) 2009, 109, 6687.

(94) Gutierrez, O.; Tantillo, D. J. J. Org. Chem. 2012, 77, 8845.

(95) Buchanan, J. G.; Ruggiero, G. D.; Williams, I. H. Org. Biomol. Chem. 2008, 6, 66.

(96) Gonzalez, C.; Schlegel, H. B. J. Chem. Phys. 1989, 90, 2154.

(97) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

(98) Ben, A. J.; Karni, M.; Apeloig, Y.; Mandelbaum, A. Int. J. Mass Spectrom. 2003, 228, 297.

(99) Morlender-Vais, N.; Mandelbaum, A. J. Mass Spectrom. 1997, 32, 1124.

(100) Kuzmenkov, I.; Etinger, A.; Mandelbaum, A. J. Mass Spectrom. 1999, 34, 797.

(101) Smith, R. L.; Thoen, K. K.; Nousiainen, J. J.; Nelson, E. D.; Kenttamaa, H. I. J. Am. Chem. Soc. 1996, 118, 8669.

Page 125: Mass spectrometric characterization of remotely charged

107  

CHAPTER SIX: EFFECT OF CHARGE POLARITY ON AMINO ACID DISSOCIATION

6.1 Introduction

The structures, properties, and reactivity of amino acids are well-characterized in

condense phase 1-4 and in solution. 5-7 However, owing to their lack of volatility, much

less is known about the structure and reactivity of amino acids in the gaseous state. 8,9

Mass spectrometry has been used to study the gas phase properties of amino acids.10-13

However, in these studies, the amino acids are typically ionized in some way, such as via

protonation 11,12 or metallation to create positively-charged ions, or by deprotonation 11 to

form negatively-charged ions. Consequently, the properties of ionized amino acids

investigated by using mass spectrometry are generally not comparable to those of the

“neutral” (zwitterionic or canonical) amino acids.13

Recently, we reported the charge-remote fragmentation14,15 of para-sulfonated

phenylalanine anion, PheSO3, using mass spectrometry.13 We discovered PheSO3

dissociates only by loss of NH3 to make a product assigned to be an α-lactone. Here, we

report the findings on charge-remote fragmentation of N,N,N-trimethyl ammonium

phenylalanine, PheNMe3+, and compared with PheSO3 . Same experimental and

computational methods as explained in Chapter 5 for PheSO3 were used in this study.

Page 126: Mass spectrometric characterization of remotely charged

108  

6.2 Synthesis

All N,N,N-trimethyl ammonium derivatives were synthesized by methylation of

aniline. Boc protected samples were used as the starting materials. For example,

PhNMe3+Cl was synthesized using Boc-4-aminophenylalanine as a starting material.

0.06 g of Boc-4-aminophenylalanine and 3 equivalent of 2, 6-lutidine were dissolved in 2

ml of DMF, and 2.5 ml of CH3I was added. The solution was stirred overnight in room

temperature. Excess solvent and CH3I were then evaporated under vacuum. The

precipitated salt was filtered and the Boc group was removed by treating with an excess,

~2.5 ml, of trifluoroacetic acid (TFA). Excess TFA was removed by evaporation under

vacuum and the product was studied in mass spectrometry.

6.3 Results and Discussion

Calculations have been carried out on the structure of PheNMe3+ in the gas phase.

Figure 6.1 shows the 7 lowest energy canonical conformation of PheNMe3+ and the

relative energies of each stationary point optimized at the B3LYP/6-31+G* level of

theory. The conformation preferences in PheNMe3+ are based on the orientation of

carboxylic acid, where cis orientation of COOH is more favorable than trans orientation.

Page 127: Mass spectrometric characterization of remotely charged

109  

However, the conformation preferences in PheSO3 are based on the relative position of

carboxylic acid with aromatic ring, where anti is more favorable than gauche (Chapter 5).

The conformational preference is not expected to affect the gas-phase properties of these

charge-remote amino acids since all structures are within 5 kcal/mol.

Figure 6.1: The 7 low energy conformations of PheNMe3+. Relative energies are shown in kcal/mol

6.3.1 Dissociation of PheNMe3+

Upon collision-induced dissociation, the PheNMe3+ with m/z 206 losses the

neutrals of masses 17, 44, 45, 59, 61, 74, 76, 87, 88 and 89 Dalton (Da), as shown in

Figure 6.2a. Other than the 17 and 45 Da losses, the other fragmentation pathways were

not observed on fragmentation of PheSO3 (Chapter 5).13 The techniques like H/D

0 0.2 0.9 1 A B C D

1 1 2 E F G

Page 128: Mass spectrometric characterization of remotely charged

110  

exchange of labile protons (d-labeled compounds), authentic sample comparison and

tandem MS are implemented to help determine the structures of the neutrals lost and the

product ions formed.

Figure 6.2. CID spectra of a) PheNMe3+ with m/z 223, b) d3- PheNMe3+ with m/z 226

m/z

130 140 150 160 170 180 190 200 210 220 2300.0

0.2

0.4

0.6

0.8

1.0

1.2

130 140 150 160 170 180 190 200 210 220 230

Rel

Int

0.0

0.2

0.4

0.6

0.8

1.0

1.2

X 10

X 10

206

223

179

178

164

162

149

147

136

134 135

206

226

182

178

167

162

149

147

137

134 138

207

a

b

Page 129: Mass spectrometric characterization of remotely charged

111  

6.3.1.1 Neutral Loss of Mass 17 Da and Formation of m/z 206 ion

Two neutrals, NH3 and OH, are possible fragments of mass 17 Da that can be lost

from PheNMe3+. The CID spectrum of the d3-PheNMe3+ with m/z 226 shown in Figure

6.2b does not have a peak at m/z 208 accounting for the loss of the OD radical. Therefore

the loss of hydroxy radical can be ruled out confirming the loss of ammonia. The

ammonia loss is a common fragment present in the CID of protonated,11,12 deprotonated11

and charge-remote13 phenylalanine.

Although the structure of the deamination product is not known, the most likely

products to be formed are either the cinnamic acid or an α-lactone derivatives (Chapter

5.4.1). The deamination product observed on dissociation of PheSO3 was deduced to be

the α-lactone based on its dissociation behavior and by ruling out possible alternatives.13

However, the recent ion spectroscopy shows the product obtained upon dissociation of

the sulfonated phenylalanine is an α-lactone confirming the assignment made previously

based on indirect evidence.

Upon CID, the [PheNMe3 –NH3]+ ion with m/z 206 losses neutrals of masses 15,

42, 44 and 46 Da. The molecular formulas of CH3, CO2 and CO2H2 can be attributed to

the masses 15, 44 and 46 Da respectively but that for neutral of 42 Da could not be

assigned. However, the neutral CO loss, a well-known decomposition pathway for α-

lactone, is not observed in the CID of the deamination product 13,16-20 suggesting the

deamination product is most likely the cinnamic acid derivative (1), Figure 6.7 (a).

A comparison of the CID spectra of [PheNMe3 –NH3]+ and the authentic 1 with

m/z 206, measured with a normalized collision energy of 35%, are shown in Figure 6.3.

The CID of the authentic 1 forms a single product ion at m/z 191, which probably is due

Page 130: Mass spectrometric characterization of remotely charged

112  

to loss of CH3. Even though the other peaks present in CID spectrum of [PheNMe3 –

NH3]+ are not observed on the CID spectrum of 1, the fact both spectra in Figure 6.3 have

m/z 191 as the base peak suggest the similarity on the structures of the [PheNMe3 –

NH3]+ and 1.

Figure 6.3. CID spectra of m/z 206 obtained from a) CID of PheNMe3+, b) authentic cinnamic acid (1)

206

191

191

206

164

162160

a

b

Page 131: Mass spectrometric characterization of remotely charged

113  

This result is also consistent with the deuterated experiment. Upon CID, the d3-

PheNMe3+ with m/z 226 losses neutrals of 19 and 20 Da which are possibly ND2H and

ND3 12 loss with a proton coming from the ethylene backbone or the carboxylic acid

respectively13 to form 1. Formation of the cinnamic acid has been reported as a minor

product for pyrolysis21,22 and thermolysis23 of phenylalanine in the condensed phase.

6.3.1.2 Neutral Loss of Mass 44 Da and Formation of m/z 179 ion

Two neutral molecules, CO2 and NMe2, are possible fragments with mass 44 Da

that can be lost from PheNMe3+. The CID spectrum of the [PheNMe3 – 44]+ matches

with the authentic phenylethanamine, 2, (Figure 6.4 a and b). This result confirms CO2 is

lost upon CID of PheNMe3+ and [PheNMe3 –CO2]+ ion with m/z 179 is 2, Figure 6.7 (b).

This result is also consistent with d-labeled experiment since CID of d3-PheNMe3+ loses

neutral of mass 44 Da as well, Figure 6.2 b).

The product ions formed upon CID of m/z 2 are also present in CID spectrum of

PheNMe3+. Thus, the CID spectrum of 2 and d2-2 shown in Figure 6.4 b and c are

compared to get further insight on fragmentation pathways. The product ions with m/z

164(166), 162(162), 135(136), and 134(134) formed upon CID of 2 (d2-2) with m/z

179(181) are due to the neutral loss of mass 15(15), 17(19), 44(45), and 47(45) Da

(number in bracket for deuterated sample).

Page 132: Mass spectrometric characterization of remotely charged

114  

Figure 6.4. CID spectra of a) m/z 179 obtained from CID of PheNMe3+, b) 2 with m/z 179, c) d2-2 with m/z 181

179

179

181

166

136

134

164

164

135

134

135

134

a

b

c

Page 133: Mass spectrometric characterization of remotely charged

115  

The neutral loss of 15(15) Da is likely CH3 radical from the ammonium group to

form [2 – CH3]+ with m/z 164(166). The CID of m/z 164(166) losses neutrals with masses

29(30) and 30(32) Da, probably to form [2 – CH3 - NHCH2]+ with m/z 135(136) and [2 –

CH3 - NH2CH2]+ with m/z 134(134), respectively. However, the CID spectrum of

[PheNMe3 – 59]+ with m/z 164 has product ions at m/z 149, 147, 135, 134, 136, 121 and

120, which are more than CID spectrum of [2-CH3] with m/z 164, Figure 6.5. The

common base peaks in both spectra suggest that the m/z 164 is mostly [PheNMe3 –CO2 -

CH3]+ with some impurities of other isobaric ion.

Figure 6.5. CID spectra of m/z 164 obtained from a) 2, b) PheNMe3+  

m/z

120 130 140 150 160 170

Rel

Int

0.0

0.2

0.4

0.6

0.8

1.0

0.0

0.2

0.4

0.6

0.8

1.0 a

b

164

164

149147

136134

135

135

120

121

a

b

Page 134: Mass spectrometric characterization of remotely charged

116  

Therefore, the m/z 135 and 134 seen in CID spectrum of PheNMe3+ are probably

[PheNMe3 –CO2 – CH3 - NHCH2]+ and [PheNMe3 –CO2 – CH3 - NH2CH2]+, Figure 6.7

(b.2.1) and (b.2.2), respectively.

The neutral of 17(19) Da with mass difference of 2 indicates both labile protons

are lost in neutral, probably as NH3 to produce [2 – NH3]+ with m/z 162(162). The CID of

[2 –NH3]+ ion loses neutral of mass 15 Da, probably CH3 neutral from ammonium site, to

produce [2 –NH3 – CH3]+ with m/z 149. Therefore, the m/z 164 and 149 seen in CID

spectrum of PheNMe3+ are probably [PheNMe3 –CO2 – NH3]+ and [PheNMe3 – CO2 –

NH3 – CH3]+, Figure 6.7 (b.1) and (b.1.1), respectively.

The decarboxylation is the main fragmentation pathway of PheNMe3+ since the

base peak at m/z 179 in CID spectrum of PheNMe3+ is due to decarboxylation. In

addition, decarboxylation leads to several other product ions like m/z 164, 162, 147, 135,

and 134 present in CID spectrum of PheNMe3+. Decarboxylation has been reported as a

major product for pyrolysis21,22 and thermolysis23 of phenylalanine in the condensed

phase as well. However, CO2 fragmentation pathway is not observed in collision-induced

fragmentation of protonated, deprotonated or charge-remote phenylalanine.

Since CID is a kinetic process, comparing the barrier a fragmentation pathways

might provide further insight on why this fragmentation is observed in one case and not

in other. Figure 6.6 shows the potential energy surface for the decarboxylation of

PheSO3, PheNMe3+ and Phe. The decarboxylation to form zwitterionic products (the

ammonium-ylide) is slightly favorable for PheNMe3+ but the experimental evidence

suggests decarboxylation leads to phenylethyl amine. Formation of the amine is

exothermic, but the barrier for proton shift in the ylide is symmetry forbidden and is

Page 135: Mass spectrometric characterization of remotely charged

117  

therefore expected to have a very high barrier. The barrier for proton shift in the

ammonium-ylide formed upon decarboxylation of Phe in the gas phase is calculated to be

25.6 kcal/mol, allowing the overall barrier for formation of the amine upon

decarboxylation of Phe to be calculated at 86.4 kcal/mol. The potential energy diagram

does not provide any testiment on why decarboxylation is preferred only in PheNMe3+.

X

CO2H

NH2

X

NH3+

_

X

NH2

X

NH3+

_

(86.4)[75.8]

X

NH2

H

+ CO2

+ CO2

+ CO2

0.02.2

(-3.6)

8.9

(13.5a)

[10.4]

63.7(60.8)[49.8]

Values in kcal/molX = SO3(X = H)[X=NMe3

+]

_

O

O

Figure 6.6. PES for decarboxylation of PheNMe3+, PheSO3 and Phe. The barriers and the molecular energy were calculated at the MP2/6-311+G**//B3LYP/6-31+G* level of

theory.

Page 136: Mass spectrometric characterization of remotely charged

118  

6.3.1.3 Neutral Loss of Mass 45 Da and Formation of m/z 178 ion

Two neutral molecules, CO2H and CONH3, are the possible fragments with mass

45 Da that can be lost from PheNMe3+. The CID spectrum of d3-PheNMe3+ shown in

Figure 6.2 b) depicts all three labile protons are lost to yield m/z 178 suggesting CONH3

is the neutral lost. The neutral loss of 45 Da was observed for the CID of PheSO3 as

well, which was loss of CO from deamination product. Since the CID [PheNMe3 - NH3] +

with did not lose CO, Figure 6.3 a), CONH3 is probably lost as a single fragment to

produce [PheNMe3 – CONH3]+ with m/z 178, Figure 6.7 (c).

6.3.1.4 Neutral Loss of Mass 74 Da and Formation of m/z 149 ion

C2O2NH4 is the only possible fragment with mass 74 Da that can be lost form

PheNMe3+. This neutral fragment is likely CH(NH2)COOH that result in formation of

the benzyl radical ion, [PheNMe3 – CH(NH2)COOH]+, with m/z 149, Figure 6.7 (d).

This result is consistent with CID of d3-PheNMe3+, where all labile protons are lost to

form m/z 149 product, Figure 6.2 b). Upon CID, [PheNMe3 – CH(NH2)COOH]+ with m/z

149 losses neutral of mass 15 Da, which probably is loss of CH3. Therefore, the m/z 134

seen in CID spectrum of PheNMe3+ is assigned as [PheNMe3 – CH(NH2)COOH – CH3]+,

Figure 6.7 (d.1).

6.3.1.5 Neutral Loss of Mass 87 Da and Formation of m/z 136 ion

The neutral with mass 87 Da lost from PheNMe3+ is likely C3O2NH5. The ionic

product has a formula that matches trimethylanilinum, [PheNMe3 – Ala + H]+ with m/z

136, Figure 6.7 (e). The CID spectrum of d3-PheNMe3+ with m/z 226 contains peaks at

Page 137: Mass spectrometric characterization of remotely charged

119  

m/z 136 and 137, Figure 6.2 b), suggesting the proton donated by alanine could be the

labile or not labile hydrogen.

6.4 Conclusion

Upon collison induded dissociation, PheNMe3+ m/z 223 produces the fragment

ions with m/z 206, 179, 178, 164, 162, 149, 147, 136, 135 and 134. The proposed

fragmentation pathways of PheNMe3+ are summarized in Figure 6.7.

The product ions with m/z 206, 179, 178, 149 and 136 are identified as primary

fragmentation pathways. The product ions with m/z 164 and 162 are secondary

fragmentation pathways and those with m/z 147, 135 and 134 are tertiary fragmentation

pathways. Far more fragmentation pathways are found for PheNMe3+ than for PheSO3-.

In PheSO3, deammination was the only primary fragmentation pathway, whereas

PheNMe3+ has 4 additional pathways. The neutrals loss of mass 45 and 74 Da proposed

as the primary fragmentation pathway on this study were observed as the secondary and

the tertiary fragmentation pathways in PheSO3, respectively.

Loss of CO2 (44 Da) and C3H5NO2 (87 Da) are observed only for the positively

charged ion. The comparisons of PES for decarboxylation pathway of PheNMe3+ and

PheSO3 do not account the descripencies in fragmentation of these charge-remote Phe.

Even when they both lose some common neutral fragments, the structures of the product

ions are different. Thus, although both PheNMe3+ and PheSO3 are charge-remote Phe,

they behave differently upon collison-induced dissociation.

Page 138: Mass spectrometric characterization of remotely charged

120

 

Fig

ure

6.7.

Pro

pose

d F

ragm

enta

tion

Pat

hway

s of

Phe

NM

e 3+

120

Page 139: Mass spectrometric characterization of remotely charged

121  

In order to explain the extensive fragmentation on PheNMe3+, we have

considered the possibility of rearrangement of the amino acid prior to dissociation.

Assuming that the benzene ring and the N,N,N-trimethyl ammonium groups (R) do not

participate in rearrangement, 19 stuctures are possible upon rearrangement of PheNMe3+

(Figure 6.8).

RNH

O

OH

(-13)

NH

O

OH

(-13)

RR

NH

O

O

(-12)

RNH

O

OH

(-5)

R

NH2 O

OH

(-1)

RNH2

O

OH

(0)

RO

O NH2

(0)

RN

OH

H O

(+1)

RN

O

O

(+2)

R NH

RNH OO

O O

(+3) (+3)

R NH O

OH

(+4)

R O

ONH2

(+5)

R

NH2

O

O

(+7)

R

O

ONH2

(+32)

R ON R

O

ONHO H

(+35)(+35)

R N OH

O

(+3)

R O

O

NH

(+7)

Ia Ib Ic Id

IIa IIb IIc IId

IIIa IIIb IIIc IIId

IIIe IIIf IIIg IIIh

IVa IVb IVc

R =

N+

Figure 6.8. Possible rearrangement products of PheNMe3+ and their relative energies in kcal/mol calculated at the B3LYP/6-31+G* level of theory .

Page 140: Mass spectrometric characterization of remotely charged

122  

Unfortunately, most of the structures in Figure 6.8 are too synthetically

challenging to examine the authentic ions. One authentic ion that has been investigated is

the carbamic acid isomer,Ia, which was formed in situ by CID of the boc-protected

phenylethylamine (eq 6.1).

 

The base peak at m/z 178 observed on CID spectrum of Ia, Figure 6.9, is due to

decarboxylation and this product ion is confirmed to be 2 by using CID (eq 6.2). The

results confirm that amino acid isomers could account for products observed upon CID of

PheNMe3+ .

Page 141: Mass spectrometric characterization of remotely charged

123  

Figure 6.9: CID spectrum of Ia with m/z 223

We propose that rearrangement of PheNMe3+ to Ia is a stepwise process, which

occurs in two steps as shown in eq 6.3. The first process involves a proton shift from an

amine to carbonyl oxygen and the attack of the electron-rich amine to electron-deficient

carbonyl to form a three-membered ring. The second process involves breaking a C-C

bond, a proton transfer from the hydroxyl group to the α-carbon and the regeneration of

carbonyl.

Although a rearrangement similar to that in eq 6.3 is possible for PheSO3, it

appparently does not occur, indicatig that the rearrangement barrier is prohibitive for

negative charged ion.

m/z

120 130 140 150 160 170 180 190 200 210 220 230

Rel

Int

0.0

0.2

0.4

0.6

0.8

1.0

223

179

Page 142: Mass spectrometric characterization of remotely charged

124  

6.5 References

(1) Albrecht, G.; Corey, R. J. Am. Chem. Soc. 1939, 61, 1087.

(2) Marsh, R. Acta Crystallogr. 1958, 11, 654.

(3) Levy, H.; Corey, R. J. Am. Chem. Soc. 1941, 63, 2095.

(4) Donohue, J. J. Am. Chem. Soc. 1950, 72, 949.

(5) Gaffney, J.; Pierce, R.; Friedman, L. J. Am. Chem. Soc. 1977, 99, 4293.

(6) Cohn, E.; McMeekin, T.; Edsall, J.; Blanchard, M. J. Am. Chem. Soc. 1934, 56, 784.

(7) Ellzy, M.; Jensen, J.; Hameka, H.; Kay, J. Spectrochim. Acta, Part A 2003, 59, 2619.

(8) Jensen, J.; MS, G. J. Am. Chem. Soc. 1991, 113, 7917.

(9) Jensen, J.; Gordon, M. J. Am. Chem. Soc. 1995, 117, 8159.

(10) Banerjee, S.; Mazumdar, S. Int. J. Anal. Chem. 2012.

(11) Piraud, M.; Vianey-Saban, C.; Petritis, K.; Elfakir, C.; Steghens, J.; Morla, A.; Bouchu, D. Rapid Commun. Mass Spectrom. 2003, 17, 1297.

(12) El Aribi, H.; Orlova, G.; Hopkinson, A.; Siu, K. J. Phys. Chem. A 2004, 108, 3844.

(13) Koirala, D.; Kodithuwakkuge, S.; Wenthold, P. J. Phys. Org. Chem. 2015, 28, 635.

(14) Cheng, C.; Gross, M. Mass Spectrom. Rev. 2000, 19, 398.

(15) Adams, J. Mass Spectrom. Rev. 1990, 9, 141.

(16) Chuchani, G.; Dominguez, R.; Rotinov, A. Int. J. Chem. Kinet. 1991, 23, 779.

(17) Chuchani, G.; Dominguez, R. Int. J. Chem. Kinet. 1995, 27, 85.

(18) Domingo, L.; Picher, M.; Andres, J.; Safont, V.; Chuchani, G. Chem. Phys. Lett. 1997, 274, 422.

(19) Chuchani, G.; Dominguez, R. Int. J. Chem. Kinet. 1999, 31, 725.

Page 143: Mass spectrometric characterization of remotely charged

125  

(20) Domingo, L.; Picher, M.; Safont, V.; Andres, J.; Chuchani, G. J. Phys. Chem. A 1999, 103, 3935.

(21) Wang, S.; Liu, B.; Sun, K.; Su, Q. J. Chromatogr. A 2004, 1025, 255.

(22) Patterso.JM; Haider, N.; Papadopo.EP; Smith, W. J. Org. Chem. 1973, 38, 663.

(23) Li, J.; Liu, Y.; Shi, J.; Wang, Z.; Hu, L.; Yang, X.; Wang, C. Thermochim. Acta 2008, 467, 20.

Page 144: Mass spectrometric characterization of remotely charged

126  

  

CHAPTER SEVEN: MOBILE C-H PROTONS IN A PROTON DEFICIENT PEPTIDE

7.1Introduction The significance of amino acids and peptides to the world around us, and to life

itself, cannot be overstated, and, therefore, it is not surprising that their structures,

properties, and reactivity are exceptionally well-characterized, largely due to the

capabilities of mass spectrometry and electrospray ionization.1 The ability for rapid,

accurate analysis of proteins and peptides by mass spectrometry has had a major impact

on the field of proteomics, and allows for the study of the structures and functions of

large proteins in complex biological system.

One of the most important features of using mass spectrometry to identify

proteins and peptides in proteomics is the ability to use tandem mass spectrometry (MSn)

to differentiate isobaric ions and for sequence determination. Therefore, whereas mass

spectral libraries can be used to identify known proteins, amino acid sequences in

unknown proteins can be determined by analysis of ionic fragments. In particular,

fragments observed in peptide dissociation, especially b- and y-type sequence ions, 2

resulting from fragmentation of amide bonds, can be used to determine the sequence from

the C- and N-terminus.2,3

Products obtained upon dissociation of protonated peptides have generally been

explained in terms of the “mobile proton theory”.2,4-6 The model assumes that protons,

Page 145: Mass spectrometric characterization of remotely charged

127  

  

initially localized on the most basic sites (N-terminus and the side chains of basic amino

acid residues) of protonated peptides, can be transferred to the less basic of sites upon

activation including the various peptide linkages, thereby producing a heterogeneous

population of protonated peptides.2,4,5 The mobile proton model provides a quantitative

framework such that if the peptide sequence and number of protons are known, one can

predict the general appearance of a fragmentation spectrum.4 Although the mobile proton

model is not intended to model the full peptide fragmentation spectrum quantitatively, it

successfully accounts for common dissociation, including aforementioned b- and y-type

ions.

Because peptide b ions are common, stable, species in peptide fragmentation,7 a

better understanding of how they are formed can improve peptide sequencing algorithms

and the current models for peptide fragmentation.2 Consequently several theoretical and

experimental studies have been carried in recent years to better understand the structures

of b ions and the mechanism of their formation.7-27 In this work, we use dipeptides to

investigate the formation of b2 ions.

Many structures have been proposed for b2 ions.4 Initially, an acylium ion was

thought to be the major b2+ ion but it was later discovered to be thermodynamically less

stable than its isomeric cyclic structures.26 Of the structures proposed for b2 ions, the

most common are protonated diketopiperazines and protonated oxazolones.3,13,18,22,27-39

Six-membered cyclic diketopiperazine is formed when the N-terminal amino group act as

a nucleophile and attack at the carbonyl carbon of second amino acid residue. However,

five-membered cyclic oxazolone is formed when the N-terminal carbonyl oxygen act as a

nucleophile and attacks at the carbonyl carbon of second amino acid residue. Which ion

Page 146: Mass spectrometric characterization of remotely charged

128  

  

forms is mainly determined by the structure of the peptide: A cis conformation of the first

amide bond favors diketopiperazine because it facilitates the nucleophilic attack by

amino group of N-terminal amino acid, Figure 7.1.32,35,39-41 Similarly, trans conformation

of the first amide bond favors oxazolone because it facilitates the nucleophilic attack by

carbonyl oxygen of N-terminal amino acid, Figure 7.1. Although the diketopiperazine is

lower in energy, oxazolone structures result for trans amide because the barrier for the

formation is lower than the barrier for cis-trans isomerization.

 

COOX

N R2H2N

O

R1

R2NH

COOX

H

OR1

NH2

cis Amide Bond

R2

NH

O

HN

OR1

R2N

O

H2NO

R1

-HXO

-HXO

Oxazolone b2 ion

Diketopiperazine b2 ion

1st residue 2nd residue

1st residue 2nd residue

trans Amide Bond

Figure 7.1. Formation of diketopiperazine from dipeptide with cis amide bond (top) and formation of oxazolone from dipeptide with trans amide bond (bottom)

The nature of the charge on peptides also plays an important role in the

dissociation mechanism. Mass spectrometry studies of peptide and peptide’s dissociation

usually involve the use of protonated42 or metallated 3,43 peptides, often multiply charged

depending on the number of basic sites. Therefore, the excess positive charge allows for

facile intramolecular proton transfer that does not require, inter alia, formation of salt-

Page 147: Mass spectrometric characterization of remotely charged

129  

  

bridges or other forms of charge separation. In a recent study, we reported the formation

and dissociation of sulfonated phenylalanine, PheSO3-, an anionic amino acid where the

charge is isolated from the α-amino acid moiety by a phenyl spacer. The ion dissociates

only by loss of NH3 to make a product assigned to be an α-lactone (eq 7.1). The proposed

mechanism involves proton transfer within the canonical amino acid to form a

zwitterionic intermediate, which dissociates by intramolecular substitution to make the

lactone (eq 7.1), and therefore nominally illustrates the potential of mobile electron

within anionic systems as well. However, an important feature of all these reactions is

that although energetically unfavorable proton transfer can occur in these reactions upon

activation, it is still generally limited to heteroatom positions, such as amines, amide

nitrogen or carboxyl groups.

In this work, we have used mass spectrometry to examine the dissociation of

dipeptides that include PheSO3¯, namely Gly-Phe*OH and Phe*-GlyOH, where Phe*

refers to PheSO3¯. Unlike protonated and metallated ions, these dipeptides are anionic

and hence proton deficient. However, they still have a canonical dipeptide moiety. In this

Page 148: Mass spectrometric characterization of remotely charged

130  

  

respect they are similar to protonated dipeptides containing His 29,38-40,44,45 or Arg,

13,24,25,46 which contain basic side chains that sequester positive charge.

We find that the sulfonated dipeptides dissociates similar to what has been

reported for Arg-Gly but the presence of anionic substrate must dissociate that involve

mobile C-H protons due to less number of mobile proton. We find Gly-Phe*OH and

Phe*-GlyOH both dissociates to form the same b2 ions after rearranging to a common

structure of the canonical amino acid.

7.2 Experimental

Experiments were carried out using a commercial LCQ-DECA (Thermo Electron

Corporation, San Jose, CA) quadrupole ion trap mass spectrometer, equipped with

electrospray ionization (ESI) source. Samples were dissolved in methanol: water (1:1)

and introduced into the source at a flow rate of 10 µl/min. Electrospray and ion focusing

conditions were varied so to maximize the signal of the ion of interest. Dissociation of

ions was carried out by using MSn experiments with mass-selected ions in the cell, with

the helium buffer serving as the collision target. Reactant ions for CID were isolated at

qz = 0.250 and with a mass-width sufficient to avoid off-resonance excitation of the

Page 149: Mass spectrometric characterization of remotely charged

131  

  

mass-selected ions. The energy of collision-induced dissociation (CID) in the cell is

reflected in the “normalized collision energy,” which ranges from 0 – 100%.

7.3 Synthesis

7.3.1 Synthesis of Amino Acid Esters

A 50 mg sample of the amino acid was dissolved in 10 ml of alcohol and cooled

to 0 C. Esterification was initiation by adding 3 ml of SOCl2, dropwise, into the cold

amino acid solution. After the solution warmed to room temperature, it was refluxed

overnight to get desired amino acid ester.

7.3.2 Synthesis of Gly-Phe*OH and Gly-Phe*OMe Dipeptides

0.09 mmol of Boc-GlyOH was dissolved in 15 ml of dichloromethane (DCM) and

0.09 mmol of coupling reagent (PyBop) was added. The solution was cooled down to

0 °C and 0.14 mmol of base (DIEA) was then added dropwise. After stirring for 20 min,

the solution was brought to a room temperature then 0.09 mmol of Phe*OH or

Phe*OMe was added and stirred for additional 40 min. Excess DCM was evaporated in

vacuum and formation of desired product was confirmed by mass spectrometry. The use

of N-Boc glycine prevents the coupling reaction in reverse order of amino acids for the

formation of Xxx-Gly dipeptide since the N is protected, where Xxx is amino acid, where

Xxx is Phe*OH in this case. Hence, the peak observed at desired m/z is Gly-Xxx and

none of Xxx-Gly. The Boc group was then removed by dissolving product in excess

trifluoroacetic acid (TFA) for 20 min and evaporating TFA in vacuum. In addition,

Page 150: Mass spectrometric characterization of remotely charged

132  

  

compound ii can be hydrolyzed to corresponding acid by treating their aqueous solution

with concentrated sulfuric acid.

7.3.3 Synthesis of Phe*-GlyOH and Phe*-GlyOMe Dipeptides

We used similar procedure in this section as explained above for Gly-Phe*OH

with some modifications. 0.09 mmole of Gly-OMe and 0.09 mmole of Phe*OH were

dissolved in 15 ml of DCM and 0.09 mmol of PyBop was added. The solution was cooled

down to 0°C and 0.14 mmol of DIEA was then added dropwise. The solution was stirred

at 0°C for 20 min and at room temperature for additional 40 min. Unlike reaction 1, we

did this reaction in single step because we want to reduce the self-coupling of Phe*OH to

form Phe*-Phe*OH dipeptide. This self-coupling reaction could be completely stopped

by using Boc-Phe*OH but Boc-Phe*OH is not commercially available and we did not

have any success on synthesizing it. To obtain the desired dipeptide, we used unmodified

Phe*OH with glycine ester. The benefit of using Gly-OMe is that it prevents Gly-Xxx

formation, which makes mass spectrum unambiguous. In addition, Phe*-GlyOMe can be

hydrolyzed to Phe*-GlyOH by treating its aqueous solution with concentrated sulfuric

acid.

7.3.4 Synthesis of Diketopiperazine

Synthesis of diketopiperazine (I) has been explained previously.47,48 Briefly, the

solution containing 50 mg of Boc-Gly-Phe*OMe and 10 ml formic acid (98%) was

stirred at room temperature for 2 hr. Excess formic acid was removed in vaccuo and the

crude product was dissolved in 15 ml of sec-butyl alcohol and 5 ml of toluene. The

Page 151: Mass spectrometric characterization of remotely charged

133  

  

solution was boiled for 3 hr and the solvent level maintained by addition of fresh butanol.

After concentrating the solution to 5-10 ml and cooling to 0C the products were filtered

off and studied on mass spectrometry.

7.3.5 Synthesis of Gly-Phe* Oxazolone

Synthesis of oxazolone has been explained previously. 48 Briefly, the solution

containing 10 umol of Boc-Gly-Phe*OH and 15 umol of 1, 3,-Dicyclohexylcarbodiimide

(DCC) were dissolved in 10 ml of dichloromethane. The solution was stirred at room

temperature for 3 hr. Excess solvent was removed in vaccuo and the resulting crude

product was dissolved in 3 ml TFA to produce Gly-Phe*-oxazolone (II).

 

Page 152: Mass spectrometric characterization of remotely charged

134  

  

7.4 Results and Discussion

7.4.1 Analysis of Gly-Phe*OH and Phe*-GlyOH

The CID (MS2) mass spectra of Gly-Phe*OH and Phe*-GlyOH (m/z 301) are

shown in Figure 7.2. Both isomers give the same products with the same intensities.

Primary product ions include m/z 284 (M-NH3), m/z 283(M-H2O), m/z 257(M-CO2), and

m/z 244 (M-C2ONH3, net loss of glycine). The fact that both Gly-Phe*OH and Phe*-

GlyOH give the same spectra indicates that they rearrange to a common structure before

dissociation. Rearrangement of dipeptides upon low-energy CID has been reported

previously by O’Hair and co-workers for protonated Arg-Gly/Gly-Arg.25 As described in

the introduction, protonated Arg-Gly/Gly-Arg is similar to Phe*-GlyOH/Gly-Phe*OH

because it is a canonical dipeptide with the charge on the side chain. Therefore, the

rearrangement of Gly-Phe*OH and Phe*-GlyOH dipeptides is likely similar to what

was proposed previous by O’Hair and co-workers for protonated Arg-Gly and Gly-Arg

dipeptides, involving proton transfer within the dipeptide to form zwitterionic structures,

which can rearrange to cyclic structure and eventually rearrange to form the common

acid anhydride intermediate (Figure 7.3). Consequently, any of the structures in Figure

7.3 can be the common intermediate leading to the dissociation product(s).

Page 153: Mass spectrometric characterization of remotely charged

135  

  

 

Figure 7.2. Mass Spectra: a) CID of Gly-Phe*OH and b) CID of Phe-Gly*OH

Page 154: Mass spectrometric characterization of remotely charged

136  

  

 

Figure 7.3. Proposed mechanism for formation of common structure from Phe*-GlyOH and Gly-Phe*OH

7.4.2 Analysis of Gly-Phe*OMe and Phe*-GlyOMe:

Support for the mechanism shown in Figure 7.3 comes from the experiments

involving methyl esters. O’Hair and co-workers25 have shown previously that

methylation of the C-terminus carboxylic acid inhibits the rearrangement reaction (Figure

7.3) of protonated Arg-Gly and Gly-Arg, presumably by blocking the initial proton

transfer step. Similarly, we have found that esterification of Gly-Phe*OH and Phe*-

Page 155: Mass spectrometric characterization of remotely charged

137  

  

GlyOH inhibits their rearrangement as well. The CID spectra of Gly-Phe*OMe and

Phe*-GlyOMe, shown in Figure 7.4, are significantly different in both the identities of

the products and the intensities. Although both ions dissociate to form products m/z 283

(M-MeOH), m/z 266 (M-MeOH-NH3) and m/z 170, the relative intensities are very

different. In particular, whereas the b2 ion (M-CH3OH @ m/z 283) is one of the dominate

peaks upon CID of Phe*-GlyOMe, it is very weak upon CID of Gly-Phe*OMe. In

addition, peaks at m/z 298, m/z 240, m/z 226, m/z 198, m/z 185 and m/z 171 are observed

solely on CID of Phe*-GlyOMe spectra, whereas m/z 266, m/z 241, and m/z 197 are

observed solely on CID of Gly-Phe*OMe spectra.

Page 156: Mass spectrometric characterization of remotely charged

138  

  

 Figure 7.4. Mass Spectra: a) CID of Gly-Phe*OMe and b) CID of Phe*-GlyOMe

The M-CH3OH ions observed upon CID of Gly-Phe*OMe and Phe*-GlyOMe

are the b2 ions. The fact that b2 ions can be observed with the dipeptides (Gly-Phe*OH

and Phe*-GlyOH) and the methyl esters illustrates why they are significant in standard

peptide sequencing: they are not dependent on the substitution at the carboxyl end. The

MS3 spectra of b2 ions obtained from Gly-Phe*OMe and Phe*-GlyOMe shown in

Figure 7.5a and 7.5b, are also different, consistent with different ions giving different

Page 157: Mass spectrometric characterization of remotely charged

139  

  

products. The CID spectra of the b2 ions obtained from Phe*-GlyOH and Gly-Phe*OH

agree with that from that obtained from Phe*-GlyOMe.

The CID spectra of the b2 ions can be compared to those for authentic ions. Figure

7.5c shows the CID spectrum of 1, whereas Figure 7.5d is for that of 2. There is a

marginal similarity between CID of the b2 ions obtained from Gly-Phe*OMe (Figure

7.5a) and for the CID 1 (Figure 7.5c). Peaks in common include m/z 255 and m/z 170.

However, the relative intensities do not match, and there are peaks in the CID of b2 ion

that are not found for CID 1. Therefore, although the CID spectrum is partially consistent

with 1, not all of the b2 ions can have that structure and there would need to be an

isobaric mixture.  

 

 

 

 

Page 158: Mass spectrometric characterization of remotely charged

140  

  

 

Figure 7.5. Mass Spectra: CID of b2 ions a) MS3 spectra of Gly-Phe*OMe and b) MS3 spectra of Phe*-GlyOMe c) CID of 1 d) CID of 2

Page 159: Mass spectrometric characterization of remotely charged

141  

  

The fact that Phe*-GlyOMe, which has no acidic proton, forms same b2 ion as

that of Phe*-GlyOH and Gly-Phe*OH indicates that the formation of b2 ions does not

require the rearrangement of precursor ions to a common structure, as shown in Figure

7.2. In addition, the CID spectrum of this b2 ion does not match the CID spectra of 1 or

that of 2. We have used isotopic labelling experiment to investigate further the formation

and structure of b2 ion from Phe*-GlyOMe.

The oxazolone structures for the b2 ions that would be obtained from Gly-

Phe*OMe and Phe*-GlyOMe are 2 and 3 respectively. An important difference between

the two structures involves the origin of the carbonyl in the oxazolone structure, which

originates from the C-terminus. Therefore, the carbonyl in 2 would come from the Phe*

residue, whereas that in 3 comes from the Gly residue. If the b2 ion obtained from Phe*-

GlyOMe has an oxazolone structure, then CID process that involves loss of the carbonyl

should be losing the glycine carbonyl. To test for this possibility, we examined Phe*-

GlyOMe with glycine labelled with 13C at the carbonyl (Figure 7.6). Upon CID, the CO

and CO2 losses both contain the 13C label and loss of non-labelled CO and CO2 is not

obtained. These results are consistent with the structure of 3 for the b2 ion.

Page 160: Mass spectrometric characterization of remotely charged

142  

  

Figure 7.6. Mass Spectrum: CID of Phe*-Gly (1-13C) OMe 

7.4.3 Deuterium Labelled Phe*-GlyOMe

We have also used deuterated labelling to investigate the mechanism of formation

of the b2 ions. Because the b2 ion formed upon dissociation of Phe*-GlyOMe is same as

that from Phe*-GlyOH/Gly-Phe*OH, we have focused on CID of Phe*-GlyOMe

because it does not undergo the rearrangement shown in Figure 7.3.

In order to determine which proton is involved in methanol loss, we examined

three deuterated labelled compounds d3-Phe*-GlyOMe, Phe*-Gly (2, 2-d2) OMe, and

d5- Phe*-GlyOMe. H/D exchange of labile protons was carried out in 50:50 D2O and

MeOD solution. The CID spectra of these compounds are shown in Figure 7.7. Upon

CID, d3-Phe*-GlyOMe loses MeOD and MeOH in a ratio of about 2:1 (Figure 7.7a).

Page 161: Mass spectrometric characterization of remotely charged

143  

  

The loss of MeOD is consistent with what is expected for b2 ion formation, 3, with loss of

methoxy from the C-terminus and loss of a mobile proton from one of the nitrogen

position. Loss of CH3OH is surprising, especially as the major product, because it

requires loss of hydrogen from a carbon position, which would not normally be

considered a “mobile” proton. Although the deuterium labelling shows those protons on

carbon are involved in the dissociation, it does not specify which carbons, whether they

are α-carbon on Gly or Phe* or from the aliphatic or aromatic region of the side-chain.

Addition deuterium labelling experiments, however, provide further insight. The

spectrum for Phe*-GlyOMe, where the Gly is labelled with deuterium in the α -positon

(2, 2-d2), is shown in Figure 7.7b. Again, the ion dissociates by loss of MeOH and

MeOD, but in this case, the intensities are reversed, with loss of MeOD being the major

b2 product.

Loss of CH3OD as the major pathway for b2 ion formation confirms that carbon-

based hydrogens are involved, and, in particular, it is the proton on the α-position of

glycine. However, it does not rule out participation of other protons, such as that at the a-

position of Phe* or those on the side chain. To test for the possibility, we also deuterated

the nitrogen positions. The CID spectrum of the d5-Phe*-GlyOMe is shown in Figure

7.7c. In this case, the b2 ion is formed almost exclusively (>98%) by loss of CH3OD.

Therefore, although b2 ion can be formed by loss of proton from nitrogen, the majority is

formed by loss of proton from the α-position of glycine.

A mechanism that occurs for the observed products is shown in Figure 7.8.

Nucleophilic attack of the Phe* carbonyl oxygen at the Gly carbonyl carbon

accompanied by proton transfer leads to the hemiacetal-like structure, X. Loss of alcohol,

Page 162: Mass spectrometric characterization of remotely charged

144  

  

with proton loss from the OH in X, leads to the standard oxazolone, accounting for loss

of the proton from N. For this pathway, the carbonyl in the oxazolone originates from the

Gly, consistent with the 13C labelling experiments. Alternatively, loss of alcohol with the

proton from the ring forms the hydroxyl-oxazole, the enol structure of the oxazolone.

Page 163: Mass spectrometric characterization of remotely charged

145  

  

 

Figure 7.7. Mass Spectra: CID of deuterated sample a) d3-Phe*-GlyOMe b) Phe*-Gly (2, 2-d2) OMe c) d5- Phe*-Gly (2, 2-d2) OMe

Page 164: Mass spectrometric characterization of remotely charged

146  

  

 

Figure 7.8. Proposed fragmentation pathway of Phe*-GlyOR’

7.4.4 Computational Results

Unlike phenols, hydroxyoxazoles are typically not more stable than the

corresponding oxazolones. At the MP2/6-31+G*//B3LYP/6-31+G* level of theory, the

simple oxazolone is computed to be 15.4 kcal/mol lower in energy than the

Page 165: Mass spectrometric characterization of remotely charged

147  

  

hydroxyoxazole, Table 7.1. However, interaction between the hydroxy group and the

sulfonate charge in the b2 is predicted to stabilize the enol structure.

The relative energies of isomeric b2 ion structures are shown in Table 7.1. The

most stable product is the diketopiperazine (1), and the 2 and 3 oxazolone b2 ions are

calculated to be 14 and 20 kcal/mol higher in energy, respectively. The relative energies

of the diketopiperizine and oxazolones are consistent with what has been found in

previous studies. However, at the MP2/6-31+G* level of theory, the lowest energy

structure, aside from the diketopiperizine, is predicted to be the hydroxyoxazole obtained

from Phe*-GlyOH, Figure 7.9. In contrast, the hydroxyoxazole obtained from Gly-

Phe*OH is significantly higher in energy. The stability of the enol structure for the

Phe*-GlyOH b2 can be attributed to a favorable hydrogen bond interaction between the

OH and the sulfonate group, as shown in Figure 7.9, and accounts for the preference for

losing proton from carbon, as opposed to from nitrogen. The oxazolone structure

obtained from Gly-Phe* also has a hydrogen bond, between the N-H and sulfonate,

which accounts for its stability. The other structures are not capable of having

interactions with the sulfonate. The optimized geometries are shown in Figure 7.9.

Table 7.1. Calculated Relative Energies of b2 Ionsa

Ion MP2/6-31+G* Diketopiperizine 1 0.0 Phe*-Gly Oxazolone 3 20 Phe*-Gly Oxazolone enol 4 12 Gly-Phe* Oxazolone 2 14 Gly-Phe* Oxazolone enol 5 26

aRelative electronic energies (not corrected) in kcal/mol at B3LYP/6-31+G* optimized geometries

Page 166: Mass spectrometric characterization of remotely charged

148  

  

Figure 7.9: Lowest energy structures of 1, 3, 4, 2, and 5 respectively

7.5 Conclusion

The charge-remote fragmentation of dipeptide is reported. We discovered that

Phe*-GlyOH and Gly-Phe*OH rearranges to common structure before dissociation. The

rearrangement is blocked by methylation. The b2 ion formed from dissociation of Phe*-

GlyOMe is same as that from Phe*-GlyOH and Gly-Phe*OH, which provided further

inside on mechanism of dissociation since first step does not need to be proton transfer.

Combination of several experimental and theoretical results suggest Phe*-Gly oxazolone

and Phe*-Gly oxazolone-enol are predominate structures of the b2 ions for the model

compound we studied. Most importantly, we have shown that C-H protons are involved

in the dissociation, effectively behaving as “mobile protons” in the proton-deficient

peptide.

Page 167: Mass spectrometric characterization of remotely charged

149  

  

7.6 References

(1) Banerjee, S.; Mazumdar, S. Int. J. Anal. Chem. 2012.

(2) Wysocki, V.; Tsaprailis, G.; Smith, L.; Breci, L. J. Mass Spectrom. 2000, 35, 1399.

(3) Lee, V.; Li, H.; Lau, T.; Siu, K. J. Am. Chem. Soc. 1998, 120, 7302.

(4) Wysocki; Cheng; Zhang; Hermann; Beardsley; Hilderbrand. In Principles of Mass spectrometry Applied to Biomolecules 2006.

(5) Boyd, R.; Somogyi, A. J. Am. Soc. Mass. Spectrom. 2010, 21, 1275.

(6) Dongre, A.; Jones, J.; Somogyi, A.; Wysocki, V. J. Am. Chem. Soc. 1996, 118, 8365.

(7) Yalcin, T.; Khouw, C.; Csizmadia, I.; Peterson, M.; Harrison, A. J. Am. Soc. Mass. Spectrom. 1995, 6, 1165.

(8) Rodriquez, C.; Shoeib, T.; Chu, I.; Siu, K.; Hopkinson, A. J. Phys. Chem. A 2000, 104, 5335.

(9) Harrison, A. Mass Spectrom. Rev. 2009, 28, 640.

(10) Ambihapathy, K.; Yalcin, T.; Leung, H.; Harrison, A. J. Mass Spectrom. 1997, 32, 209.

(11) Bythell, B.; Knapp-Mohammady, M.; Paizs, B.; Harrison, A. J. Am. Soc. Mass. Spectrom. 2010, 21, 1352.

(12) Bowie, J.; Brinkworth, C.; Dua, S. Mass Spectrom. Rev. 2002, 21, 87.

(13) Zou, S.; Oomens, J.; Polfer, N. Int. J. Mass spectrom. 2012, 316, 12.

(14) Chen, X.; Tirado, M.; Steill, J.; Oomens, J.; Polfer, N. J. Mass Spectrom. 2011, 46, 1011.

(15) Chen, X.; Yu, L.; Steill, J.; Oomens, J.; Polfer, N. J. Am. Chem. Soc. 2009, 131, 18272.

(16) Chu, I.; Shoeib, T.; Guo, X.; Rodriquez, C.; Lan, T.; Hopkinson, A.; Siu, K. J. Am. Soc. Mass. Spectrom. 2001, 12, 163.

(17) Bythell, B.; Csonka, I.; Suhai, S.; Barofsky, D.; Paizs, B. J. Phys. Chem. B 2010, 114, 15092.

Page 168: Mass spectrometric characterization of remotely charged

150  

  

(18) Farrugia, J.; O'Hair, R.; Reid, G. Int. J. Mass spectrom. 2001, 210, 71.

(19) Grewal, R.; El Aribi, H.; Harrison, A.; Siu, K.; Hopkinson, A. J. Phys. Chem. B 2004, 108, 4899.

(20) Harrison, A. J. Am. Soc. Mass. Spectrom. 2009, 20, 2248.

(21) Knapp-Mohammady, M.; Young, A.; Paizs, B.; Harrison, A. J. Am. Soc. Mass. Spectrom. 2009, 20, 2135.

(22) Yoon, S.; Chamot-Rooke, J.; Perkins, B.; Hilderbrand, A.; Poutsma, J.; Wysocki, V. J. Am. Chem. Soc. 2008, 130, 17644.

(23) Wysocki, V.; Resing, K.; Zhang, Q.; Cheng, G. Methods 2005, 35, 211.

(24) Forbes, M.; Jockusch, R.; Young, A.; Harrison, A. J. Am. Soc. Mass. Spectrom. 2007, 18, 1959.

(25) Farrugia, J.; O'Hair, R. Int. J. Mass spectrom. 2003, 222, 229.

(26) Nold, M.; Wesdemiotis, C.; Yalcin, T.; Harrison, A. Int. J. Mass Spectrom. Ion Processes 1997, 164, 137.

(27) Wang, D.; Gulyuz, K.; Stedwell, C.; Polfer, N. J. Am. Soc. Mass. Spectrom. 2011, 22, 1197.

(28) Polce, M.; Ren, D.; Wesdemiotis, C. J. Mass Spectrom. 2000, 35, 1391.

(29) Perkins, B.; Chamot-Rooke, J.; Yoon, S.; Gucinski, A.; Somogyi, A.; Wysocki, V. J. Am. Chem. Soc. 2009, 131, 17528.

(30) Paizs, B.; Suhai, S. J. Am. Soc. Mass. Spectrom. 2004, 15, 103.

(31) Paizs, B.; Lendvay, G.; Vekey, K.; Suhai, S. Rapid Commun. Mass Spectrom. 1999, 13, 525.

(32) Morrison, L.; Chamot-Rooke, J.; Wysocki, V. Analyst 2014, 139, 2137.

(33) Eckart, K.; Holthausen, M.; Koch, W.; Spiess, J. J. Am. Soc. Mass. Spectrom 1998, 9, 1002.

(34) Bythell, B.; Somogyi, A.; Paizs, B. J. Am. Soc. Mass. Spectrom. 2009, 20, 618.

(35) Bythell, B.; Suhai, S.; Somogyi, A.; Paizs, B. J. Am. Chem. Soc. 2009, 131, 14057.

Page 169: Mass spectrometric characterization of remotely charged

151  

  

(36) Sinha, R.; Erlekam, U.; Bythell, B.; Paizs, B.; Maitre, P. J. Am. Soc. Mass. Spectrom. 2011, 22, 1645.

(37) Paizs, B.; Suhai, S. J. Am. Soc. Mass. Spectrom. 2003, 14, 1454.

(38) Farrugia, J.; Taverner, T.; O'Hair, R. Int. J. Mass spectrom. 2001, 209, 99.

(39) Armentrout, P.; Clark, A. Int. J. Mass spectrum. 2012, 316, 182.

(40) Gucinski, A.; Chamot-Rooke, J.; Nicol, E.; Somogyi, A.; Wysocki, V. J. Phys. Chem. A 2012, 116, 4296.

(41) Paizs, B.; Suhai, S. Rapid Commun. Mass Spectrom. 2001, 15, 2307.

(42) Paizs, B.; Suhai, S. Mass Spectrom. Rev. 2005, 24, 508.

(43) Dunbar, R.; Polfer, N.; Berden, G.; Oomens, J. Int. J. Mass spectrum. 2012, 330, 71.

(44) Dunbar, R.; Steill, J.; Polfer, N.; Oomens, J. Int. J. Mass spectrum. 2009, 283, 77.

(45) Tsaprailis, G.; Nair, H.; Zhong, W.; Kuppannan, K.; Futrell, J.; Wysocki, V. Anal. Chem. 2004, 76, 2083.

(46) Tsaprailis, G.; Somogyi, A.; Nikolaev, E.; Wysocki, V. Int. J. Mass spectrum. 2000, 195, 467.

(47) Nitecki, D.; Halpern, B.; Westley, J. J. Org. Chem. 1968, 33, 864.

(48) Uraguchi, D.; Asai, Y.; Ooi, T. Angew. Chem. Int. Ed. 2009, 48, 733.

Page 170: Mass spectrometric characterization of remotely charged

VITA

Page 171: Mass spectrometric characterization of remotely charged

152  

VITA Damodar Koirala was born in November 1985 to Bishnu Prasad and Radhika

Koirala in Pokhara Nepal. He is youngest three children, and is loved, cared, and

supported by bother Kamal Koirala and sister Kamala Bhattarai throughout his career.

Upon graduating from high school from VS Niketan College, Kathmandu, he joined

University of Wisconsin-Superior for undergraduate degree in Aug 2005. He started with

Biology major but later switched to Chemistry and Mathematic majors. He graduated

with Bachelor of Science degree on Dec 2010. He worked for a year at UW-Superior

after graduation and joined Purdue University to pursue his PhD degree in Chemistry on

Aug 2011. He joined Dr. Paul Weldhold’s lab on Nov 2011 and soon started expertizing

on application of mass spectrometer.

On May 2013, Damodar got married to his beautiful wife Binita K. Koirala. They

then had a baby girl, Erina Koirala, on April of 2014. Both, Binita and Erina, supported

him in all expects since they arrived to his life. Currently, they live in Lafayette, IN, with

Damodar’s brother Kamal, sister-in-law Goma, 3 years old nephew Kritan and 2 and half

months old niece Keva Koirala. They are a big happy family.

Damodar will graduate with a Ph.D. in Chemistry in August of 2016. Eventually,

he desires to become an industrial chemist or a Chemistry teacher in a small

undergraduate University.

Page 172: Mass spectrometric characterization of remotely charged

PUBLICATIONS

Page 173: Mass spectrometric characterization of remotely charged

Mass spectrometric study of the decompositionpathways of canonical amino acids andα-lactones in the gas phaseDamodar Koiralaa, Sampath Ranasinghe Kodithuwakkugea‡ andPaul G. Wentholda*

The dissociation pathways of a gas-phase amino acid with a canonical (non-zwitterionic) α-amino acid moiety are studied byusing mass spectrometry. Investigation of the canonical amino acid moiety is possible because the ionized amino acid, asulfonated phenylalanine, has a charge center that is separated from the amino acid, and dissociation occurs by charge-remote fragmentation. The amino acid is found to dissociate only by loss of NH3 upon collision-induced dissociation to form asubstituted α-lactone. The dissociation is consistent with what has been observed previously upon pyrolysis of other α-substitutedcarboxylic acids. Decarboxylation, which has also been reported previously for amino acid pyrolysis, is not observed, likely becausethe product would be a high-energy, ammonium ylide. The resulting α-lactone is found to undergo dissociation by decarbonylationto give an aldehyde, and by loss of CO2. Decarboxylation is calculated to occur through a transition state involving hydride shiftcoupled with lactone ring-opening. The transition state is found to be stabilized by the negative charge, and therefore, decar-boxylation is more favorable for anions. The results show that remote ionic groups can be used as mostly inert charge carriersto enable mass spectrometry to be used to investigate the gas-phase physical and chemical properties of different types offunctional groups, including amino acids. Copyright © 2015 John Wiley & Sons, Ltd.

Keywords: amino acid dissociation; gas-phase amino acids; mass spectrometry

INTRODUCTION

The significance of amino acids to the world around us, and tolife itself, cannot be overstated, and, therefore, it is not surprisingthat their structures, properties, and reactivity are exceptionallywell-characterized, at least in solution. However, owing to theirgeneral lack of volatility, much less is known about the structureand reactivity of amino acids in the gaseous state. Gaseousamino acids are rare, but would be highly significant if reportsof interstellar glycine,[1,2] for example, can be confirmed. In amore practical sense, the gas phase serves as a reasonable ap-proximation for a low dielectric, hydrophobic environment, andso investigation of the gas-phase (solvent free) properties ofamino acids provides insight into the structures of amino acidsin lipid membranes.[3] Finally, regardless of any biological signif-icance, results of studies of gaseous amino acids can be com-pared with those for substituted carboxylic acids to determine,inter alia, the effect of the amino group on the gas-phase struc-ture and reactivity.One of the most important differences between gaseous and

condensed phase amino acids is that, because charge separationis unfavorable in the gas phase, the solvent-free molecules willnot have zwitterionic structures, Z, like those in the condensedphase, but will have non-ionic, canonical (C) structures instead.[4–7]

In fact, zwitterionic structures of gas-phase amino acids are so un-expected that many mass

spectrometric studies have been carried out to try to elucidatethose factors, such as solvation or ion pairing that can lead tostable ionic structures in various amino acids.[8]

Recent advances in spectroscopic methods in the last decadehave provided more detailed insight into the gas-phase struc-tures, including details of the modes of internal hydrogen bond-ing.[9–11] The challenges in these experiments are twofold: on theone hand, because the spectroscopy experiments generally re-quire a good chromophore for ultraviolet absorption, they workbest for amino acids with aromatic side chains. However, thoseamino acids are among the less volatile, and therefore, creatinggaseous samples is difficult. Smaller, aliphatic amino acids aremore volatile, but lack a good chromophore to facilitateresonance-2-photon absorption, which is generally required forthese types of experiments. It is a testament to modern spectro-scopic technology that we do have spectra for both aromaticand aliphatic amino acids.[12–15] Gas-phase photoionization of al-iphatic amino acids has also been reported.[16]

Amino acids have been extensively studied using mass spec-trometry. However, in these studies, the amino acids are typicallyionized in some way, such as via protonation or metallation tocreate positively charged ions or by deprotonation to form neg-

* Correspondence to: P. G. Wenthold, Department of Chemistry, Purdue University,560 Oval Drive, West Lafayette, IN 47907, USA.E-mail [email protected]

‡ Deceased, 2012

a D. Koirala, S. R. Kodithuwakkuge, P. G. WentholdDepartment of Chemistry, Purdue University, 560 Oval Drive, West Lafayette,IN, 47906, USA

Research Article

Received: 14 January 2015, Revised: 05 May 2015, Accepted: 11 May 2015, Published online in Wiley Online Library

(wileyonlinelibrary.com) DOI: 10.1002/poc.3464

J. Phys. Org. Chem. 2015 Copyright © 2015 John Wiley & Sons, Ltd.

153

Page 174: Mass spectrometric characterization of remotely charged

atively charged ions. Consequently, the properties of ionizedamino acids investigated by usingmass spectrometry are generallynot comparable with those of the “neutral” (zwitterionic or canon-ical) amino acids. Here, we describe an investigation of canonicalamino acids by using “charge-remote fragmentation,”[17,18] wheredissociation of an ion occurs, as the name suggests, at locationsremote from the charge site. Charge-remote fragmentation is com-monly used as an analytical tool to determine the structures ofbiological molecules. Lipids[17,18] are particularly amenable to thistechnique, as they fragment in the hydrocarbon chains. However,charge-remote fragmentation has also been used to examinestructures of other biologically relevant molecules such as drugsand their metabolites and peptides.[17,18]

Adams and Gross[19] have proposed that charge-remote frag-mentation, where the charge does not participate in the frag-mentation, is analogous to thermolysis of the gas-phase neutralmolecule. If this is true, then it may be possible to use charge-remote fragmentation

electrosprayed ions as an alternative to pyrolysis in order to studythe gaseous decomposition of otherwise non-volatile organicmolecules. In this study, we examined sulfonated phenylalanine,PheSO3

�, a sulfonate anion with a separate canonical amino acidmoiety. The gas-phase reactivity of amino acids is similarly poorlyknown. The lack of volatility of amino acids has made even nomi-nally simple experiments such as gas-phase pyrolysis exceedinglydifficult, and these experiments generally require generation ofthe amino acid in situ by pyrolysis of appropriate volatile precur-sors. For example, Chuchani and co-workers[20] reported that theycould examine the pyrolytic dissociation of amino acids generatedby decomposition of the corresponding ethyl ester derivatives(Eqn 1). By using this approach, they found that N-substitutedglycines, such as N,N-dimethylglycine,[20] N-phenylglycine,[21]

and N-benzylglycine,[22] for example, dissociate by loss of CO2,as shown in Eqn 1, in the gas phase. In contrast, Al-Awadi andco-workers found that gaseous N-phenylalanine does not un-dergo decarboxylation upon pyrolysis, but dissociates into ani-line, CO, and acetaldehyde, as shown in Eqn 2,[23] in a processthat is similar to what is commonly observed for α-substitutedcarboxylic acids (Eqn 3).[24–36]

By using mass spectrometry, we have examined the decomposi-tion of the gaseous amino acid, PheSO3

�, which contains a sulfo-nate charge and a canonical α-amino acid. We show thatPheSO3

� dissociates initially by fragmentation of the amino acidby the pathway shown in Eqn 3 to form the α-lactone, and direct de-carboxylation (Eqn 1) is not observed. The α-lactone product isfound to fragment by decarbonylation, as expected (Eqn 3) but isalso found to undergo decarboxylation to form the styrene product.Electronic structure calculations find that the presence of the sulfo-nate charge has little effect on the energetics of these dissociationpathways, indicating that the results can be fairly interpreted as rep-resentative of what would happen with the non-ionized derivative.

EXPERIMENTAL

Experiments were carried out using a commercial LCQ-DECA(Thermo Electron Corporation, San Jose, CA, USA) quadrupole iontrap mass spectrometer, equipped with electrospray ionization(ESI). Ions were introduced by spraying dilute aqueous solutionsof commercially available substrates, except for the cis-cinnamicacid and 4-(acryloyloxy)benzenesulfonate, which were synthesizedas described in the succeeding text; solution details are providedas Supporting Information. Electrospray and ion focusing condi-tions were varied so to maximize the signal of the ion of interest.Dissociation of ions can be carried out during ion injection by usinghigh injection voltages, or can be carried out by using MSn exper-iments with mass-selected ions in the cell, with the helium bufferserving as the collision target. Reactant ions for CID were isolatedat qz = 0.250 and with a mass width sufficient to avoid off-resonance excitation of the mass-selected ions. The energy ofcollision-induced dissociation (CID) in the cell is reflected in the“normalized collision energy,” which ranges from 0% to 100%.

Synthesis of (Z)-4-(3-ethoxy-3-oxoprop-1-en-1-yl)benzenesulfonate (deprotonated ethyl 4′-sulfo-cis-cinnamate)

A solution of ethyl diphenyl phosphonoacetate (0.16g, 0.50mmol) intetrahydrofuran (30ml)was treatedwith Triton B (0.27ml, 0.60mmol)at �78 °C for 15min. A solution of p-sulfonated benzaldehyde(0.12g, 0.60mmol in 5ml MeOH) was then added dropwise, andthe resulting mixture was stirred at �78 °C for 1h. The reaction wasquenchedwith 5ml of deionizedwater. ESImass spectra of the crudeproduct indicated a mixture of m/z 255 (the sulfonated ethylcinnamate) and m/z 249, which is assigned to a (PhO)2PO2

� by-product. Nuclearmagnetic resonance spectroscopy of the crudemix-ture in D2O contained doublets at 6.0 and 6.98ppm (J=12.4Hz),which could be assigned to the cis-cinnamate based on comparisonwith the previously reported cis-structures. In contrast, the clearlyobserved NMR peak for the authentic trans isomer is a doublet at6.49ppm with J=15.4. From the NMR spectrum of the cis-sample,we estimate a cis/trans ratio of about 98:2, consistent with the selec-tivities obtained previously using this procedure. The ethyl ester wasintroduced into the mass spectrometer by electrospray andconverted to the carboxylic acid by CID, and elimination of ethylene.

Synthesis of 4-(acryloyloxy)benzenesulfonate (PASO3�)

4-(Acryloyloxy)benzenesulfonate was prepared by mixing 1ml of4-hydroxybenzenesulfonic acid sodium salt dihydrate with 3mlof acryloyl chloride (CH2¼CHCOCl). The neat mixture was stirredfor about 15min at 35°, and 1ml of (i-Pr)2Net was added. The so-lution was stirred approximately 15min, after which one drop of

(1)

(3)

(2)

D. KOIRALA, S. R. KODITHUWAKKUGE AND P. G. WENTHOLD

wileyonlinelibrary.com/journal/poc Copyright © 2015 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2015

154

Page 175: Mass spectrometric characterization of remotely charged

the crude mixture was added to 1ml of a 50/50 mixture of waterand methanol and examined by ESI mass spec. The main prod-ucts observed in the mass spectrum of the crude mixture arethe hydroxybenzenesulfonate, a product with m/z 253, which islikely HO3SC6H4SO4

�, and a product at m/z 227, which is thephenyl acrylate.

Computational methods

Amino acid conformations were surveyed by using the MonteCarlo Multiple Minimum search method, with the Merck MolecularForce Field (MMFFs) as implemented in MacroModel (version9.6).[37] Multiple starting conformations were utilized, includingcanonical and zwitterionic structures, to improve the chances offinding the key low-energy structures. All of the unique conforma-tions calculated with the MMFFs force field to be within 100 kJ/molof the lowest energy structure were identified and used as startingpoints for B3LYP/6-31+G* geometry optimizations.Stationary points were confirmed to be minima or saddle

points from the computed vibrational frequencies. Reported en-ergies are electronic energies, computed at the MP2/6-311+G**level of theory at the B3LYP geometries, designated MP2/6-311 +G**//B3LYP/6-31 +G*, unless noted, and are not correctedfor zero-point energies or thermal energy contributions. Thelarger basis set is used to provide a good description of the vander Waals interactions and polarization of charge. Calculationswere carried out using QCHEM version 4.0[38] and Gaussian.[39]

RESULTS

The mass spectrum of an aqueous solution of PheSO3�, shown in

Fig. 1(a), is very clean, with essentially only peaks correspondingto PheSO3

�. Considering the large difference in acidities ofsulfonic and carboxylic acids,[40] it is not surprising that thesulfonate structure, depicted as PheSO3

�, is computed (B3LYP/6-31+G*) to be more than 20 kcal/mol more stable than the carbox-ylate structure, similar to what has been reported previously forsulfated peptides.[41] Although mixtures of ion isomers have beenfound for amino acid anions when the acidities of the differentgroups are similar,[42,43] the large difference in the acidities of thecarboxylic and sulfonic acids should result in an ion that consistsof a sulfonate charge with separate amino acid moiety. Althoughthe extent of interaction between the charge and the reactivegroup is always a question in these studies, the distance to thecharge is too great for any significant interaction to occur. In addi-tion, the saturated carbons that separate the amino acid from thering prevent any conjugative interactions with the charge.Calculations have been carried out to determine the structure of

PheSO3� in the gas phase. A conformational search identified

more than 20 unique low-energy conformations (within50 kJ/mol of minimum) of the ion, and all but one which have ca-nonical structures. Qualitative depictions of the five lowest energycanonical conformations and the lowest energy zwitterionic struc-ture are shown in Fig. 2. Figure 2 also shows the relative energies ofeach stationary point, optimized at the B3LYP/6-31+G* level oftheory. The most stable conformation has the carboxylic acid moi-ety anti to the aromatic ring, with a hydrogen bond between the –OH proton and the amine nitrogen. Indeed, the anti-conformation,when possible, is generally preferred for the structures examinedin this work. The conformation with the –COOH and aromatic ringgauche is higher in energy by about 2.5 kcal/mol. Alternate hydro-gen bonding arrangements, such as NH– or OH to carbonyl

oxygen, or NH to hydroxyl hydrogen, are higher in energy by morethan 4 kcal/mol. The most important conclusion from this compu-tational survey is that the lowest energy zwitterionic structures aresignificantly higher in energy than the lowest canonical forms.Figure 2 also shows that the relative energies are similar whetherusing either DFT or ab initio methods.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

150 170 190 210 230 250 270 290

M

M-NH3

M-NH3-COM-NH3-CO2

a

b

Figure 1. (a) ESI and (b) CID mass spectra of PheSO3�. The CID for spec-

trum was carried out with a normalized collision energy of 25%

Figure 2. Low-energy conformations of the amino acid moiety ofPheSO3

� and their relative energies, in kcal/mol, computed at theB3LYP/6-31 + G* and MP2/6-311 + G** (in parentheses) level of theory

CID OF CANONICAL AMINO ACIDS AND LACTONES

J. Phys. Org. Chem. 2015 Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc

155

Page 176: Mass spectrometric characterization of remotely charged

Dissociation of PheSO3�

The CID spectrum of PheSO3� is shown in Fig. 1(b). The major frag-

ments observed are readily assigned as resulting from dissociationof the canonical amino acid backbone. In particular, the main dis-sociation processes all involve loss of NH3 to form an ion withm/z 227, which subsequently fragments by loss of either CO, toform m/z 199, or CO2, to form m/z 183. That the m/z 183 and199 ions that come from the M–NH3 anion are confirmed by twoadditional results. First, although m/z 183 and 199 are major prod-ucts observed at the energy used for CID in Fig. 1(b), the M–NH3

ion is the only product observed at very low energies, consistentwith a sequential dissociation. Second, CID of the isolated M–NH3

anion, generated either by CID during the ESI injection stage orby CID in the course of an MS/MS/MS experiment, verifies that itdissociates by loss of CO and CO2 in a nearly 1:1 ratio as observedin Fig. 1(b). Therefore, the loss of NH3 is the only pathway that canbe uniquely assigned to PheSO3

�. Direct loss of CO2, which hasbeen found previously for pyrolysis of gas-phase amino acids(Eqn 1),[20] is not observed upon activation of PheSO3

�.Although the structure of the M–NH3 product is not known,

the most likely products to consider are those that are stableproducts formed by minimal rearrangement. Therefore, the mostlikely products are either a cis-cinnamic acid or trans-cinnamicacid (CASO3

�), an α-lactone (αLSO3�), or the phenyl acrylate

(PASO3�). Other products are possible but would require more

extensive rearrangement.Collision-induced dissociation of deprotonated phenyl alanine

is known to result in formation of cinnamic acid.[44–46] Similarly,dissociation of PheSO3

� by elimination across the ethylene back-bone would result in the formation of a cinnamic acid, CASO3

�,as shown in Eqn 4. However, multiple lines of evidence argueagainst the cinnamic acid. First, ammonia loss from d3-labeledPheSO3

�, formed by H/D exchange using a mixture of D2O/CD3CO2D as the solvent, occurs by loss of ND3 only, and loss ofND2H, NDH2, or NH3 is not observed to any detectable extentat any collision energy.

This rules out the possibility of elimination across the back-bone as shown in Eqn 4. It is also evidence against other morecomplicated rearrangement processes involving non-amino acidportions of the molecule.

However, cinnamic acid formation from PheSO3� does not re-

quire dissociation across the backbone. An alternate mechanism,as shown in Eqn 5, proceeds via a zwitterionic intermediate.

Therefore, we have examined the CID behavior of authentictrans-cinnamic acid and cis-cinnamic acid to compare with theproduct obtained from PheSO3

�. For this experiment, sulfonatedtrans-cinnamic acid was formed by ESI of an aqueous solution ofthe commercially available sulfonyl chloride, and the cis-cinnamicacid was formed in situ by CID of the corresponding ethyl ester. Acomparison of the CID spectra of the m/z 227 product obtainedfrom PheSO3

� and the authentic cinnamic acids, measured witha normalized collision energy of 34%, is shown in Fig. 3. Whereas

Figure 3. CID spectra of m/z 227 ions obtained (a) by CID of PheSO3�,

and from sulfonated (b) cis and (c) trans-cinnamic acid

(4)

(5)

D. KOIRALA, S. R. KODITHUWAKKUGE AND P. G. WENTHOLD

wileyonlinelibrary.com/journal/poc Copyright © 2015 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2015

156

Page 177: Mass spectrometric characterization of remotely charged

all ions lose CO2 as the major fragmentation pathway, significantdifferences are evident. For example, the CASO3

� ions do not loseCO to any appreciable extent, and lose only SO2 (to formm/z 163)in addition to losing CO2. Moreover, although the m/z 183 ionthat corresponds to PheSO3

�–NH3–CO2 is found to lose SO2 uponCID (vide infra), the CASO3

� ions undergo facile loss of CO2 andSO2 to form m/z 119 in much higher yield than is found for theM–NH3 ion. Finally, although the authentic CASO3

� ions dissoci-ate by loss of CO2, they do so only at energies much higher thanthose needed for dissociation of PheSO3

�. The amounts of them/z 163 and m/z 119 products observed in Fig. 3(a), 3(b), and 3(c) indicate that CASO3

� constitutes at most about 5% of theM–NH3 product.As shown in Eqn 6, the zwitterionic intermediate could also re-

arrange through nucleophilic attack at the ipso-position followedby a Grob-like fragmentation to give PASO3

�.

Although PASO3� is likely a relatively stable product, it can be

ruled out for multiple reasons. First, it is inconsistent with the CIDresults, in that it cannot readily lose either CO or CO2. Second,formation of the phenyl acrylate would require a high-energy,non-aromatic transition state (vide infra). Ultimately, the forma-tion of PASO3

� was ruled out by using an authentic sample.The primary dissociation pathway of PASO3

� is loss of theacryloyl group (CH2¼CHCO) to form the phenoxyl radical, lossof CO or CO2 does not occur at all (Supporting Information).In contrast, formation of the α-lactone is consistent with all

data. Formation of the lactone, either by a concerted loss ofNH2–H or through a zwitterionic intermediate (Eqn 7), involvesloss of only the exchangeable protons, and is therefore consis-tent with the results for the deuterated ions.

Loss of CO like that observed here is a well-known decompo-sition pathway for α-lactones,[24–36] and therefore, loss of COfrom the M–NH3 anion is consistent with the presence of theα-lactone. Although decarboxylation is generally not considereda decomposition pathway for that type of structure, it is not

unprecedented,[47] and computational studies described hereinpredict that it is expected to occur for αLSO3

�.Based on these results, we conclude that the M–NH3 ion found

upon CID of PheSO3� is most likely the α-lactone, αLSO3

�. Al-though α-lactones are nominally high-energy reactive species,they have previously been proposed as products in the CID ofα-substituted carboxylates in mass spectrometry experiments,[48–51]

which is essentially how it is shown being formed via the zwitter-ion in Eqn 7. The difference between the previous studies and thiswork is that the α-lactone formed upon fragmentation charge-remote fragmentation of PheSO3

� is ionic, and therefore detect-able by mass spectrometry. Previous studies of halogenatedcarboxylates[48–51] have inferred α-lactone structures based onthe observation of halide products.

Dissociation of the α-lactone, αLSO3�

As described in the aforementioned section, the PheSO3�–NH3

product, assigned to be the α-lactone (αLSO3�), dissociates by

loss of either CO or CO2. Loss of CO from α-lactones is well-knownand predicted to occur with a barrier of about 28–30 kcal/mol[52]

to form the corresponding aldehyde or ketone (Eqn 3). The CO-loss product obtained upon CID of αLSO3

� in this work is foundto undergo further dissociation by loss of 29 mass units, whichwe assign to HCO as shown in Eqn 8.

Dissociation to form the benzyl radical product, BzSO3�, is also

observed upon CID of control substrates, such as sulfonatedphenylpropionic acid or phenethylamine (Eqn 9), and so it isnot surprising that it is also observed for the aldehyde.

The loss of CO2 has not been reported previously for the gas-phase decomposition of α-lactones, although it has been foundto occur, along with decarbonylation, upon photolysis of thebis-n-butyl-substituted α-lactone in solution,[47] forming the ole-fin (Eqn 10).

By comparing CID spectra of the m/z 183 product with an au-thentic sample (Supporting Information), we have confirmedthat the product formed upon decarboxylation of αLSO3

� is sim-ilarly the corresponding styrene (Eqn 11).

(6)

(7)

(8)

(9)

(10)

CID OF CANONICAL AMINO ACIDS AND LACTONES

J. Phys. Org. Chem. 2015 Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc

157

Page 178: Mass spectrometric characterization of remotely charged

In summary, we have found that the dissociation of the gas-phase amino acid, PheSO3

�, results in the formation of an α-lactone, αLSO3

�, which dissociates by decarbonylation to formthe aldehyde, and by decarboxylation accompanied by hydrogenshift to form the styrene.

DISCUSSION

The results described in the previous section are the first studiesof the dissociation of a simple (non-N-substituted), canonicalamino acid in the gas phase. The dissociation pathway for theamino acid is similar to that observed previously for substitutedalanine[23] and other substituted carboxylic acids,[24–36] but isvery different from what has been reported for glycine deriva-tives.[20–22] The difference in dissociation pathways between gly-cine and alanine derivatives has been noted previously[22] buthas not been explained satisfactorily.

That the decomposition of PheSO3� involves only the amino

acid moiety and not the SO3� group is evidence that the sulfo-

nate charge does not participate in the dissociation processes.Fragmentation of the SO3

� group is not even observed for anyof the dissociation products, except for the sulfonated styreneshown in Eqn 11, which dissociates by loss of SO2. Therefore,the experimental results indicate that the sulfonate group isserving extensively as a charge carrier, and does not significantlyaffect the possible dissociation pathways. Although the dissocia-tion pathway, loss of NH3, is the same as what is found fordeprotonated phenylalanine,[44–46] the product formed with thation is deprotonated cinnamic acid. Control experiments in thiswork show that cinnamic acid is not formed upon dissociationof PheSO3

�. The mechanism proposed for deprotonated phenyl-alanine[44–46] involves intramolecular proton transfer to thecarboxylate and subsequent proton transfer to the amine, andtherefore, the charge center is extensively involved in the pro-cess. The lack of formation of cinnamic acid with PheSO3

suggests that the charge center is not participating in the disso-ciation process. Deprotonated tyrosine also undergoes CID byloss of ammonia,[53] possibly via the phenoxide ion.

Electronic structure calculations predict that the sulfonatecharge does not have a large effect on the energies of the reactionpathways for PheSO3

�. A comparison of the relative energies ofobserved and other possible products in the dissociation ofPheSO3

� and neutral phenylalanine, Phe, calculated at theMP2/6-311+G**//B3LYP/6-31+G* level of theory, is shown inTable 1. Table 1 shows that the relative energies for all the possibledissociation products for PheSO3

� and Phe are very similar, withdifferences for the primary process of less than 3 kcal/mol.

Amino acid dissociation

A potential energy surface for the dissociation of PheSO3�,

shown in Fig. 4, provides insight into the product selectivity. Be-cause CID is a kinetic process, the most important features of thepotential energy surface are the activation barriers. Therefore,

we have calculated the barriers for NH3 and CO2 loss fromPheSO3

� and Phe at the MP2/6-311 +G**//B3LYP/6-31 +G* levelof theory. Despite multiple attempts, we could not find any tran-sition states that directly connect the canonical structure of theamino acid with the expected deamination or decarboxylationproducts. However, insight into the mechanism was obtainedfrom multidimensional potential energy surface calculations(Supporting Information). We find that loss of NH3 or CO2 occurseffectively by dissociation of the zwitterionic structure, in thatproton transfer from the carboxylic acid to the amino groupoccurs very early in the process at energies well below thoserequired for dissociation.

Table 1. Computed reaction energies for dissociation path-ways of PheSO3

� and Phe1

Relativeenergies

Products PheSO3� Phe

Amino acid (reactant) 0.0 0.0ArCH2C(CO)CH+NH3 (α-lactone, Eqn 7) 47.0 44.6ArCH¼CHCO2H+NH3 (trans-cinnamicacid, Eqn 4)

17.7 16.4

ArCH2CH2NH2 +CO2 1.1 -4.6ArCH2CHO+CO+NH3 34.9 31.9ArCH¼CH2 +CO2 +NH3 15.9 11.4ArOC(O)CH¼CH2 +NH2 39.4 39.01Electronic energy differences between products and theamino acid reactants, computed at the MP2/6-311 +G**//B3LYP/6-31 +G* level of theory.

Figure 4. Calculated energetics for the dissociation of PheSO3�

(X¼SO3�) and Phe (X¼H). Values correspond to electronic energies and

geometries calculated at the MP2/6-311 + G**//B3LYP/6-31 + G* level oftheory (a) not a stable geometry; correspond to an amino acid geometrythe same as that for zwitterionic PheSO3

�; (b) computed utilizing a3 kcal/mol barrier for the reverse reaction, obtained from a relaxed sur-face scan; see Supporting Information

(11)

D. KOIRALA, S. R. KODITHUWAKKUGE AND P. G. WENTHOLD

wileyonlinelibrary.com/journal/poc Copyright © 2015 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2015

158

Page 179: Mass spectrometric characterization of remotely charged

Loss of NH3 from Phe to form the α-lactone is estimated tohave a barrier of about 3 kcal/mol higher than the reaction en-ergy, about 46 kcal/mol. The mechanism for ammonia elimina-tion involves an intramolecular SN2 reaction (Eqn 12), wherethe carboxylate displaces the amine leaving group in the zwitter-ionic structure. Therefore, the dissociation of PheSO3

� to formthe α-lactone is analogous to the elimination reactions of α-halocarboxylates that have been reported previously.[50,54–60] Inan ab initio study, Davidson and co-workers[61] found that thereis no barrier in excess of the dissociation energy for the forma-tion of the α-lactone from chloroacetate, whereas a small barrierwas calculated for the zwitterionic structure of Phe in this work.

The barrier for elimination of NH3 to form the cinnamic acid,not shown in Fig. 4, is calculated to be 68 kcal/mol, which ac-counts for why that process does not occur. However, given thatcharge-remote fragmentation is analogous to pyrolysis, it is notsurprising that cinnamic acid is not formed, considering that py-rolysis of alkylated amines does not occur by concerned elimina-tion of NH3, but results in the formation of imines.[62] Similarly,we have calculated the barrier for the formation of the phenylacrylate, PASO3

�, as shown in Eqn 6. Whereas for the neutralphenyl alanine, we find a concerted transition state for the rear-rangement accompanied by ammonia loss, the transition statefor the sulfonated version has the ammonia completely dissoci-ated, and is part of a stepwise process. In either case, the barrierfor phenyl acrylate formation is very high, computed to be about80 kcal/mol from the zwitterion, and 90–100 kcal/mol from thecanonical structure.Although loss of CO2 from the amino acid is not observed in

this work, it has been reported previously, and therefore, wehave carried out calculations to determine why it does not occurfor PheSO3

�. Decarboxylation of the zwitterionic structure of Pheis calculated to occur without a barrier in excess of the dissocia-tion energy, such that the net barrier for the reaction is deter-mined by the energies of the dissociation products. Therefore,loss of CO2 has a barrier of more than 60 kcal/mol. As shown inFig. 4, loss of CO2 would necessarily lead to formation of thezwitterionic product (the ammonium ylide) and not thephenethyl amine. Although formation of the amine is exother-mic, Alexandrova and Jorgensen[63] have noted that the barrierfor proton shift in the ylide is formally symmetry forbidden andis therefore expected to have a very high barrier. Consequently,the formation of the amine from glycine is essentially a sequen-tial, two-step process consisting of decarboxylation followed byproton transfer, with a proton transfer barrier of approximately45 kcal/mol for the ylide in aqueous solution.[63] The barrier forproton shift in the ammonium ylide formed upon decarboxyl-ation of Phe in the gas phase is calculated to be 25.6 kcal/mol,such that the overall barrier for the formation of the amine upondecarboxylation of Phe is calculated to be 86.4 kcal/mol. Thehigher barrier for proton shift computed for glycine is likelydue to solvent stabilization of the ammonium ylide. Interestingly,Alexandrova and Jorgensen also calculated a slight (~8 kcal/mol)barrier in excess of the dissociation energy for the decarboxylation

step. However, the presence of the excess barrier is also likely aconsequence of explicitly including solvation in the QM/MM ap-proach, and would not be expected in a gas-phase calculation.[64]

The barriers for decarboxylation obtained for glycine byAlexandrova and Jorgensen[63] and for Phe in this work aresignificantly higher than what has been reported previouslyfor the decarboxylation of canonical N,N-dimethylglycine orN-phenylglycine.[20,21,65] However, the ~42 kcal/mol barrierssuggested in the previous studies do not seem reasonable be-cause they are lower than the energies required for decarbox-ylation, which we calculate to be 52.3 and 59.6 kcal/mol fordimethylglycine and N-phenylglycine, respectively. Consideringthat decarboxylation leads initially to the ammonium ylide,and that rearrangement of the ylide to the amine occurs in asecond step and is formally symmetry forbidden, decarboxyl-ation should not occur with an activation energy less than thatrequired for the formation of the ammonium ylide.

In summary, although loss of CO2 from the amino acid can oc-cur to form the ammonium ylide, formation of the α-lactone ispredicted to be energetically preferred by about 20 kcal/mol.Consequently, that is the only process that is observed forPheSO3

�.

Dissociation of the α-lactone

In this study, we have also been able to investigate the dissocia-tion of the α-lactone, αLSO3

�. α-Lactones are highly reactive mol-ecules that have been proposed as short-lived intermediates in avariety of chemical reactions, often involving α-substituted car-boxylates[50,54–60] or α,β-unsaturated acids,[66,67] in the photolysisof cyclic peroxides[47,68–71] and α-halocarboxylic acids,[72] in theoxidation of ketenes,[73] or in the gas-phase dissociation ofα-substituted carboxylic[24–29] or in the decomposition of aminoacids.[23,74] They have also been proposed as intermediates in at-mospheric oxidation reactions[75,76] in mass spectrometry,[77–80]

and as products of the reaction of carbenes with CO2.[81–85] De-

spite being highly reactive, α-lactones have been investigatedspectroscopically by using matrix isolation,[82–84] gas phase,[86]

and solution-phase time-resolved IR.[85] There has long beena discussion regarding the electronic structure of α-lactones,and whether they are best considered to be cyclic, canonicalstructures[56,58–60] or zwitterionic,[55] although those discussionsgenerally relate to the structure in solution. The calculated struc-tures of gas-phase α-lactones have exceptionally long (~1.55 Å)and weak[87] Cα–O bonds, and even in the gas phase, there is con-siderable ionic character.[87]

α-Lactones are highly reactive, and react rapidly by polymeri-zation, by nucleophilic addition, or by decomposition.[88] Al-though an early photolysis study reported dissociation by lossof CO and CO2,

[47] more recent studies of α-substituted carboxylicacid pyrolysis that proceeds through α-lactone intermedi-ates[24,25,30–36] have reported dissociation only by loss of CO. Inthis work, we observe loss of both CO and CO2 from the α-lactone,in similar amounts (if anything, CO2 loss is slightly favored). As de-scribed earlier, loss of CO occurs to form the aldehyde, as ex-pected, whereas loss of CO2 results in the formation of thestyrene (Eqn 10).

Computed potential energy surfaces for the decompositionchannels are shown in Fig. 5. The barriers for CO loss from theα-lactone derived from PheSO3

� and Phe are predicted to beabout 30 kcal/mol, similar to that predicted for alkyl-substitutedα-lactones.[52] Determining the barrier for styrene formation is

(12)

CID OF CANONICAL AMINO ACIDS AND LACTONES

J. Phys. Org. Chem. 2015 Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc

159

Page 180: Mass spectrometric characterization of remotely charged

more difficult. The experimental observation that CO and CO2

are formed similarly indicates that the barriers for the two pro-cesses are similar. Therefore, a stepwise mechanism consistingof decarboxylation to form the carbene followed by hydrogenshift likely has a barrier that is too high to compete with CO loss(Fig. 5). Considering that decarboxylation of β-lactones results di-rectly in the formation of the olefin and occurs with moderateenergy barriers,[89,90] we explored the possibility of a mechanisminvolving the dyotropic shift[91–94] to the β-lactone[95] followedby decarboxylation (Eqn 13a). However, the search for the transi-tion state for α-lactone to β-lactone rearrangement did not give astructure for a dyotropic process, as shown in 10a, but insteadgave a structure that consists of 1,2-hydrogen shift without assis-tance of the carboxylate anion, as shown in Eqn 13b. The struc-ture shown in Eqn 13b was confirmed to be a saddle point byvibrational analysis, and an intrinsic reaction coordinate calcula-tion[96,97] finds that it is a transition state that connects theα-lactone with the styrene and CO2.

The 1,2-hydrogen shift mechanism shown in Eqn 13b is facili-tated by the high degree of polarization expected for the α-lactone(Eqn 14),[87] and by the fact that the β-cationic carboxylate (β(+)–CO2) that would result from hydride transfer is unstable with

respect to decarboxylation in the gas phase. The coupling of hy-drogen migration and leaving-group loss is analogous to whathas been proposed for the Schmidt reaction,[94] and has been ob-served previously in reactions such as the elimination of proton-ated ethers[98,99] and acetates.[100]

The computed barriers for decarboxylation of the α-lactoneare 32.1 kcal/mol for αLSO3

�, and 42.1 kcal/mol for the non-ionicsystem, indicating that the sulfonate is having a large effect onthe stability of the transition state. Because the charge of the sul-fonate group does not delocalize into the π-system of the ringvia conjugation, the stabilization of the transition state is likelyan inductive effect that stabilizes the formation of the benzyliccation. The effect would be expected to be stronger for an ionicgroup that is conjugated with the aromatic ring, such as an O�

group of a phenoxide (Eqn 15), which accounts for why thephenoxide ion obtained from deprotonating tyrosine loses NH3

and CO2 as the main dissociation channel, and an M–NH3–COion is not observed at all.[53]

The calculated barriers for decarboxylation and decarbonylationof the ionic and non-ionic substrates are consistent with the exper-imentally observed results. The large difference in barrier heightsfor the non-ionic lactone accounts for why loss of CO is generallythe only observed channel in pyrolysis experiments. However, forαLSO3

�, the computed barriers for loss of CO and CO2 are very sim-ilar, which accounts for why the products are observed in nearlyequal amounts. As shown in Eqn 15, the resonant interaction be-tween the charge site and the benzylic site in deprotonated tyro-sine is expected to strongly favor the CO2 loss channel. Finally,preliminary results with para-trimethylammonium-substitutedPhe find that the corresponding α-lactone dissociates only by lossof CO, consistent with electrostatic destabilization of the benzyliccation.

CONCLUSIONS

By using an ion with the charged moiety isolated from an aminoacid, we have been able to utilize mass spectrometry to investi-gate the chemical properties of a gas-phase amino acid. In thiswork, we have characterized the decomposition pathways for aphenylalanine derivative. The only observed pathway is loss ofammonia, to form an α-lactone. This reaction is similar to whathas been observed previously for pyrolysis of N-substituted ala-nines[23] and other α-substituted carboxylic acids,[24–36] but is in-consistent with what has been reported for pyrolysis of glycine

Figure 5. Calculated energetics for the dissociation of αLSO3� (X¼SO3

�)and the non-sulfonated derivative (X¼H). Values correspond to elec-tronic energies and geometries calculated at the MP2/6-311 + G**//B3LYP/6-31 + G* level of theory, unless indicated (a) not a stable geome-try; corresponds to the geometry of the lactone with the CO2 removed

(13b)

(13a)

(14)

(15)

D. KOIRALA, S. R. KODITHUWAKKUGE AND P. G. WENTHOLD

wileyonlinelibrary.com/journal/poc Copyright © 2015 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2015

160

Page 181: Mass spectrometric characterization of remotely charged

derivatives.[20–22] The main difference between the studies thatfind deamination[23] (including this work) and those that find de-carboxylation to occur[20–22] is our work and that of Al-Awadiet al.[23] involve a direct investigation of the gaseous amino acid,whereas Chuchani and co-workers have tried to generate theamino acid in situ by pyrolysis of the corresponding ethyl esters.Considering that the computational results in this work and byothers[63] have shown that decarboxylation of gaseous aminoacids to form the amine has a prohibitively high barrier, theproducts that correspond to decarboxylation of the amine arelikely formed via a pathway not involving the amino acid.The α-lactone that is formed from the amino acid in this work

undergoes dissociation by loss of CO and CO2. Although loss ofCO is expected, the loss of CO2 has generally not been reportedin reactions involving α-lactones. However, given the calculatedrelative barriers for CO and CO2 loss, decarboxylation should oc-cur, to some extent, in α-lactones containing a β-hydrogen thatcan shift during lactone ring-opening, similar to the reactionshown in Eqn 13b. It may be that the formation of the benzyliccation may facilitate the hydride shift reaction, although decar-boxylation has been observed for a completely aliphatic deriva-tive (Eqn 9).[47] The results of this work and that reportedpreviously for deprotonated tyrosine show that stabilization ofthe β-cation can affect the branching ratio for CO versus CO2

loss. Therefore, just as anionic stabilization of the transition statefavors CO2, the presence of a cationic group would be expectedto disfavor decarboxylation, and favor loss of CO, which is what isobserved in preliminary studies.Finally, the results of this study show that mass spectrometry

can be used to investigate the properties of canonical aminoacids. Whereas this is a new approach for the investigation ofgas-phase amino acids, the use of an inert, remote charge to in-vestigate neutral chemistry is not new and is similar, in principle,to the distonic ion approach used by Kenttämaa and co-workers[101] to examine the reactivity of aromatic radicals. Inthe same way, it should be possible to use mass spectrometryto investigate further the reactivity of gas-phase amino acids,and investigate the effects of structure and solvation.

Acknowledgements

This work was supported by the National Science Foundation(CHE11-11777). Calculations were carried out using the resourcesof the Center for Computational Studies of Open-Shell and Elec-tronically Excited Species (iopenshell.usc.edu), supported by theNational Science Foundation through the CRIF:CRF program. Wethank Profs. Anastassia Alexandrova and Nouria Al-Awadi for thehelpful comments, Dr. Will McGee for the help with the triple-quadrupole experiments, and Hari R. Khatri for the assistancewith the synthesis and characterization of the cis-cinnamic acid.

REFERENCES[1] Y.-J. Kuan, S. B. Charnley, H.-C. Huang, W.-L. Tseng, Z. Kisiel,

Astrophys. J. 2003, 593, 848.[2] L. E. Snyder, F. J. Lovas, J. M. Hollis, D. N. Friedel, P. R. Jewell, A.

Remijan, V. V. Ilyushin, E. A. Alekseev, S. F. Dyubko, Astrophys. J.2005, 619, 914.

[3] A. C. V. Johansson, E. Lindahl, J. Chem. Phys. 2009, 130, 185101.[4] C. J. Chapo, J. B. Paul, R. A. Provencal, K. Roth, R. J. Saykally, J. Am.

Chem. Soc. 1998, 120, 12956.[5] R. R. Julian, M. F. Jarrold, J. Phys. Chem. A. 2004, 108, 10861.[6] J. H. Jensen, M. S. Gordon, J. Am. Chem. Soc. 1995, 117, 8159.

[7] E. Kassab, J. Langlet, E. Evleth, Y. Akacem, J. Mol. Struct.(THEOCHEM). 2000, 531, 267.

[8] M. F. Jarrold, Annu. Rev. Phys. Chem. 2000, 51, 179.[9] T. S. Zwier, J. Phys. Chem. A. 2006, 110, 4133.[10] J. M. Bakker, L. M. Aleese, G. Meijer, G. v. Helden, Phys. Rev. Lett.

2003, 91, 203003.[11] J.-P. Schermann, Spectroscopy and Modelling of the Biomolecular

Building BlocksElsevier, Netherland, 2008.[12] Y. Hu, E. R. Bernstein, J. Chem. Phys. 2008, 128, 164311/1.[13] Y. Hu, E. R. Bernstein, J. Phys. Chem. A. 2009, 113, 8454.[14] R. Antoine, P. Dugourd, Phys. Chem. Chem. Phys. 2011, 13,

16494.[15] Y. K. Gao, F. Traeger, K. Kotsis, V. Staemmler, Phys. Chem. Chem.

Phys. 2011, 13, 10709.[16] H.-W. Jochims, M. Schwell, J.-L. Chotin, M. Clemino, F. Dulieu, H.

Baumgartel, S. Leach, Chem. Phys. 2004, 298, 279.[17] J. Adams, Mass Spectrom. Rev. 1990, 9, 141.[18] C. Cheng, M. L. Gross, Mass Spectrom. Rev. 2000, 19, 398.[19] J. Adams, M. L. Gross, J. Am. Chem. Soc. 1989, 111, 435.[20] A. Ensuncho, M. J. Lafont, A. Rotinov, R. M. Dominguez, A. Herize, J.

Quijano, G. Chuchani, Int. J. Chem. Kinet. 2001, 33, 465.[21] R. M. Dominguez, M. Tosta, G. Chuchani, J. Phys. Org. Chem. 2003,

16, 869.[22] M. Tosta, J. C. Oliveros, J. R. Mora, T. Cordova, G. Chuchani, J. Phys.

Chem. A. 2010, 114, 2483.[23] S. A. Al-Awadi, M. R. Abdallah, M. A. Hasan, N. A. Al-Awadi, Tetrahe-

dron. 2004, 60, 3045.[24] N. A. Al-Awadi, A. Kumar, G. Chuchani, A. Herize, Int. J. Chem. Kinet.

2001, 33, 612.[25] G. Chuchani, R. M. Dominguez, A. Rotinov, I. Martin, J. Phys. Org.

Chem. 1999, 12, 612.[26] L. R. Domingo, M. T. Picher, J. Andres, V. S. Safont, G. Chuchani,

Chem. Phys. Lett. 1997, 274, 422.[27] L. R. Domingo, M. T. Picher, V. S. Safont, J. Andres, G. Chuchani,

J. Phys. Chem. A. 1999, 103, 3935.[28] A. Rotinov, G. Chuchani, J. Andre, L. R. Domingo, V. S. Safont, Chem.

Phys. 1999, 246, 1.[29] V. S. Safont, V. Moliner, J. Andres, L. R. Domingo, J. Phys. Chem. A.

1997, 101, 1859.[30] G. Chuchani, R. M. Dominguez, Int. J. Chem. Kinet. 1995, 27, 85.[31] G. Chuchani, R. M. Dominguez, Int. J. Chem. Kinet. 1999, 31, 725.[32] G. Chuchani, R. M. Dominguez, A. Rotinov, Int. J. Chem. Kinet. 1991,

23, 779.[33] G. Chuchani, I. Martin, A. Rotinov, Int. J. Chem. Kinet. 1995,

27, 849.[34] G. Chuchani, I. Martin, A. Rotinov, R. M. Dominguez, J. Phys. Org.

Chem. 1993, 6, 54.[35] G. Chuchani, A. Rotinov, Int. J. Chem. Kinet. 1989, 21, 367.[36] G. Chuchani, A. Rotinov, R. M. Dominguez, J. Phys. Org. Chem. 1996,

9, 787.[37] Macromodel, version 9.6, Schrödinger, LLC, New York, NY, 2009.[38] J. Kong, C. A. White, A. I. Krylov, D. Sherrill, R. D. Adamson, T. R.

Furlani, M. S. Lee, A. M. Lee, S. R. Gwaltney, T. R. Adams, C.Ochsenfeld, A. T. B. Gilbert, G. S. Kedziora, V. A. Rassolov, D. R.Maurice, N. Nair, Y. Shao, N. A. Besley, P. E. Maslen, J. P. Dombroski,H. Daschel, W. Zhang, P. P. Korambath, J. Baker, E. F. C. Byrd, T. V.Voorhis, M. Oumi, S. Hirata, C.-P. Hsu, N. Ishikawa, J. Florian, A.Warshel, B. G. Johnson, P. M. W. Gill, M. Head-Gordon, J. A. Pople,J. Comput. Chem. 2000, 21, 1532.

[39] Gaussian 03, Revision B.03, M. J. Frisch, G. W. Trucks, H. B. Schlegel,G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. J. A. Montgomery, T.Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J.Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda,J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M.Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J.Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R.Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A.Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich,A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K.Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S.Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz,I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y.Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson,W. Chen, M. W. Wong, C. Gonzalez, J. A. Pople, Gaussian, Inc.,Pittsburgh, PA, 2003.

CID OF CANONICAL AMINO ACIDS AND LACTONES

J. Phys. Org. Chem. 2015 Copyright © 2015 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/poc

161

Page 182: Mass spectrometric characterization of remotely charged

[40] J. E. Bartmess, “Negative Ion Energetics”. in: NIST Chemistry Webbook,NIST Standard Reference Database Number 69 (Eds: P. J. Lindstrom,W. G. Mallard), National Institutes of Standards and Technology:Gaithersburg, MD, 20899, http://webbook.nist.gov, (retrieved June2, 2015).

[41] T. T. Nha Tran, T. Wang, S. Hack, J. H. Bowie, Rapid Commun. MassSpectrom. 2013, 27, 1135.

[42] Z. Tian, A. Pawlow, J. C. Poutsma, S. R. Kass, J. Am. Chem. Soc. 2007,129, 5403.

[43] Z. Tian, X.-B. Wang, L.-S. Wang, S. R. Kass, J. Am. Chem. Soc. 2009,131, 1174.

[44] M. Eckersley, J. H. Bowie, R. N. Hayes, Int. J. Mass Spectrom. Ion Pro-cesses. 1989, 93, 199.

[45] S.-S. Choi, O.-B. Kim, Int. J. Mass spectrom. 2013, 338, 17.[46] A. M. Couldwell, M. C. Thomas, T. W. Mitchell, A. J. Hulbert, S. J.

Blanksby, Rapid Commun. Mass Spectrom. 2005, 19, 2295.[47] W. Adam, R. Rucktaeschel, J. Org. Chem. 1978, 43, 3886.[48] M. Baker, W. Gabryelski, Int. J. Mass spectrom. 2007, 262, 128.[49] T. S. Choi, J. Y. Ko, S. W. Heo, Y. H. Ko, K. Kim, H. I. Kim, J. Am. Soc.

Mass Spectrom. 2012, 23, 1786.[50] S. T. Graul, R. R. Squires, Int. J. Mass Spectrom. Ion Processes. 1990,

100, 785.[51] N. J. Rijs, R. A. J. O’Hair, Dalton Trans. 2012, 41, 3395.[52] L. R. Domingo, J. Andres, V. Moliner, V. S. Safont, J. Am. Chem. Soc.

1997, 119, 6415.[53] Z. Tian, S. R. Kass, J. Am. Chem. Soc. 2008, 130, 10842.[54] C. M. Bean, J. Kenyon, H. Phillips, J. Chem. Soc. 1936, 303.[55] W. A. Cowdrey, E. D. Hughes, C. K. Ingold, S. Masterman, A. D. Scott,

J. Chem. Soc. 1937, 1252.[56] E. Grunwald, S. Winstein, J. Am. Chem. Soc. 1948, 70, 841.[57] B. Strijtveen, R. M. Kellogg, Recl. Trav. Chim. Pays-Bas. 1987, 106,

539.[58] S. Winstein, J. Am. Chem. Soc. 1939, 61, 1635.[59] S. Winstein, R. B. Henderson, J. Am. Chem. Soc. 1943, 65, 2196.[60] S. Winstein, H. J. Lucas, J. Am. Chem. Soc. 1939, 61, 1576.[61] D. Antolovic, V. J. Shiner, E. R. Davidson, J. Am. Chem. Soc. 1988,

110, 1375.[62] Y. Hamada, H. Takeo, Appl. Spectrosc. Rev. 1992, 27, 289.[63] A. N. Alexandrova, W. L. Jorgensen, J. Phys. Chem. B. 2011, 115,

13624.[64] A. N. Alexandrova, personal communication ed., 2013.[65] J. R. Perez, M. Lorono, R. M. Dominguez, T. Cordova, G. Chuchani,

J. Phys. Org. Chem. 2008, 21, 402.[66] N. Pirinccioglu, J. J. Robinson, M. F. Mahon, J. G. Buchanan, I. H. Wil-

liams, Org. Biomol. Chem. 2007, 5, 4001.[67] S. Niwayama, H. Noguchi, M. Ohno, S. Kobayashi, Tetrahedron Lett.

1993, 34, 665.[68] W. Adam, J.-C. Liu, O. Rodriguez, J. Org. Chem. 1973, 38, 2269.[69] W. Adam, R. Rucktaeschel, J. Amer. Chem. Soc. 1971, 93, 557.[70] O. L. Chapman, P. W. Wojtkowski, W. Adam, O. Rodriguez, R.

Rucktaeschel, J. Amer. Chem. Soc. 1972, 94, 1365.[71] W. Adam, L. Blancafort, J. Org. Chem. 1996, 61, 8432.[72] S. Nishino, M. Nakata, J. Mol. Struct. 2008, 875, 520.

[73] M. Schmittel, H. von Seggern, Liebigs Ann. Chem. 1995, 1815.[74] J. Kenyon, H. Phillips, Trans. Faraday Soc. 1930, 26, 451.[75] S. A. Carr, D. R. Glowacki, C.-H. Liang, M. T. Baeza-Romero, M. A.

Blitz, M. J. Pilling, P. W. Seakins, J. Phys. Chem. A. 2011, 115, 1069.[76] H. G. Kjaergaard, H. C. Knap, K. B. Oernsoe, S. Joergensen, J. D.

Crounse, F. Paulot, P. O. Wennberg, J. Phys. Chem. A. 2012, 116,5763.

[77] D. Schroder, N. Goldberg, W. Zummack, H. Schwarz, J. C. Poutsma,r. R. Squires, Int. J. Mass Spectrom. Ion Processes. 1997, 165/166, 71.

[78] A. K. Y. Lam, R. A. J. O’Hair, Rapid Commun. Mass Spectrom. 2010,24, 1779.

[79] J. A. Wyer, L. Feketeova, N. S. Brondsted, R. A. J. O’Hair, Phys. Chem.Chem. Phys. 2009, 11, 8752.

[80] P. B. Armentrout, S. J. Ye, A. Gabriel, R. M. Moision, J. Phys. Chem. B.2010, 114, 3938.

[81] G. B. Kistiakowsky, K. Sauer, J. Am. Chem. Soc. 1958, 80, 1066.[82] D. E. Milligan, M. E. Jacox, J. Chem. Phys. 1962, 36, 2911.[83] S. Wierlacher, W. Sander, M. T. H. Liu, J. Org. Chem. 1992, 57, 1051.[84] W. W. Sander, J. Org. Chem. 1989, 54, 4265.[85] B. M. Showalter, J. P. Toscano, J. Phys. Org. Chem. 2004, 17, 743.[86] S.-Y. Chen, Y.-P. Lee, J. Chem. Phys. 2010, 132, 114303/1.[87] G. D. Ruggiero, I. H. Williams, J. Chem. Soc., Perkin Trans. 2. 2001,

733.[88] G. L’Abbé, Angew. Chem., Int. Ed. Eng. 1980, 19, 276.[89] I. Morao, B. Lecea, A. Arrieta, F. P. Cossio, J. Am. Chem. Soc. 1997,

119, 816.[90] R. Ocampo, W. R. Dolbier Jr., M. D. Bartberger, R. Paredes, J. Org.

Chem. 1997, 62, 109.[91] M. T. Reetz, Angew. Chem. Int. Ed. Engl. 1972, 11, 130.[92] M. T. Reetz, Angew. Chem. Int. Ed. Engl. 1972, 11, 129.[93] I. Fernandez, F. P. Cossio, M. A. Sierra, Chem. Rev. (Washington, DC,

U. S.). 2009, 109, 6687.[94] O. Gutierrez, D. J. Tantillo, J. Org. Chem. 2012, 77, 8845.[95] J. G. Buchanan, G. D. Ruggiero, I. H. Williams, Org. Biomol. Chem.

2008, 6, 66.[96] C. Gonzalez, H. B. Schlegel, J. Chem. Phys. 1989, 90, 2154.[97] C. Gonzalez, H. B. Schlegel, J. Phys. Chem. 1990, 94, 5523.[98] A. J. Ben, M. Karni, Y. Apeloig, A. Mandelbaum, Int. J. Mass

Spectrom. 2003, 228, 297.[99] N. Morlender-Vais, A. Mandelbaum, J. Mass Spectrom. 1997, 32,

1124.[100] I. Kuzmenkov, A. Etinger, A. Mandelbaum, J. Mass Spectrom. 1999,

34, 797.[101] R. L. Smith, K. K. Thoen, J. J. Nousiainen, E. D. Nelson, H. I.

Kenttamaa, J. Am. Chem. Soc. 1996, 118, 8669.

SUPPORTING INFORMATION

Additional supporting information may be found in the onlineversion of this article at the publisher’s web-site.

D. KOIRALA, S. R. KODITHUWAKKUGE AND P. G. WENTHOLD

wileyonlinelibrary.com/journal/poc Copyright © 2015 John Wiley & Sons, Ltd. J. Phys. Org. Chem. 2015

162

Page 183: Mass spectrometric characterization of remotely charged

Reactivity of 3- and 4-pyridinylnitrene-n-oxide radical anions

Damodar Koirala a, James S. Poole b, Paul G. Wenthold a,*aDepartment of Chemistry, Purdue University, 560 Oval Drive, West Lafayette, IN 47907, United StatesbDepartment of Chemistry, Ball State University, Muncie, IN, United States

A R T I C L E I N F O

Article history:Received 1 May 2014Received in revised form 4 July 2014Accepted 4 July 2014Available online 9 July 2014

Keywords:Flowing afterglowIon–molecule reactionsKineticsPyridine-n-oxidesNitrene radical anions

A B S T R A C T

Ion–molecule reactions in a flowing afterglow are used to examine the electronic structure of 3- and4-pyridinylnitrene-n-oxide radical anions. Reactions with nitric oxide are generally similar to thosereported previously for other aromatic nitrene radical anions. In particular, phenoxide formation bynitrogen–oxygen exchange is observed with both isomers. Oxygen atom abstraction by NO is alsoobservedwith both isomers. Very significant differences in the reactivity are observed in the reactions ofthe two isomers with carbon disulfide. The reactivity of the 3-n-oxide isomer with CS2 is similar to thatobserved previously for nitrene radical anions, and reactions of the n-oxide moiety are not observed,similar towhat is expected based on solution chemistry. The 4-n-oxide isomer, however, undergoesmanyreactions, including oxygen atom and oxygen ion transfer and sulfur–oxygen exchange, that involve then-oxide oxygen. The increased reactivity of the oxygen is attributed to increased charge density at theoxygen due to pi electron donation of the nitrene anion in the para position.

ã 2014 Elsevier B.V. All rights reserved.

1. Introduction

Nitrene radical anions are a fascinating class of hypovalent ions.Despite their unusual electronic structures, isoelectronic withcarbene radical anions [1–6], they are surprisingly easily generatedin mass spectrometry, cleanly and in high abundance, by simpleelectron ionization of azide-substituted precursors [7–20].Although azides are considered potentially explosive, given therecent interests in azides as reagents in “click” chemistry [21–24]they are commonly being synthesized and utilized in chemicalapplications, and, inprinciple, anyof these derivatives could be usedas precursors of nitrene radical anions. However, despite how easilythe ions are to form, few studies of nitrene radical anions have beenreported, and many of those reported have focused on the use ofthe anions as precursors for photoelectron/photodetachmentspectroscopy studies of the corresponding nitrenes [9–14].

Nevertheless, some studies have addressed issues of electronicstructures of nitrene radical anions. The simplest nitrene radicalanion, HN�, has been well-studied experimentally [25–31] andcomputationally [32]. Isoelectronic with OH, the HN� ion is foundto have a 2P electronic ground state. Ellison and co-workers [14]discussed the electronic structure of methyl nitrene radical anion,CH3N�, which has a 2E ground state, but, like CH3O, undergoes aJahn–Teller distortion [33,34].

[TD$INLINE]The electronic structures of phenylnitrene anions have been

discussed in the context of photoelectron [12,13] and photo-detachment [9–11] spectroscopy studies, and have been examinedby electronic structure calculations [13]. The planar phenylnitreneanion, PhN�, has an electronic structure that consist of threeelectrons in the two non-bonding molecular orbitals (NBMOs) ofPhN, which consist of an in-plane, s orbital, and a benzylic-likep orbital (Fig. 1). Unlike the case in C3v methylnitrene, the orbitalsin PhN are not degenerate, and therefore there are two possibleelectronic states that can be created, corresponding to s2p (2B2)andp2s (2B1). For substituted systems where the symmetry of thesystem is reduced to Cs, the electronic states correspond to 2A00 and2A0 , respectively. B3LYP calculations with large basis setspredict that the ground states of PhN and chloro-substitutedderivatives [13] are all p2s, with the s2p states lying approxi-mately 0.5 eV higher in energy.

Benzoylnitrene radical anion, BzN�, has been examinedcomputationally and experimentally [18–20]. Like PhN�, theground state of BzN� is predicted to be p2s (2A0), but the relativeenergy of the s2p state (28 kcal/mol) is predicted [18] to be muchhigher than that in PhN�, reflecting a greater preference for the

* Corresponding author. Tel.: +1 765 494 0475.E-mail address: [email protected] (J.S. Poole).

http://dx.doi.org/10.1016/j.ijms.2014.07.0171387-3806/ã 2014 Elsevier B.V. All rights reserved.

International Journal of Mass Spectrometry 378 (2015) 69–75

Contents lists available at ScienceDirect

International Journal of Mass Spectrometry

journal homepage: www.elsevier .com/ locate / i jms

163

Page 184: Mass spectrometric characterization of remotely charged

carboxylate-like anion over having the charge localized on anitrogen s orbital. Chemical reactivity of the BzN� is consistentwith the open-shell structure. In particular, reaction with NOresults via radical–radical coupling to give nitrogen/oxygenexchange (Eq. (1)), leading to the formation of benzoate anionand N2.

(1)

[TD$INLINE]In this study, we report an investigation of the reactivity of the

3- and 4-pyridinylnitrene-n-oxide radical anions, 3PNO� and4PNO�, respectively. The pyridine-N-oxide is an interestingstructural motif. The formal pyridinium can nominally be viewedas an

[TD$INLINE]electron-acceptor, but the adjacent oxide can act as ap donor, suchthat the electronic effect of the N-oxide is dependent on the extentof benefit that can be created, via electron donation or acceptance.However, the resonance effect of the n-oxide can only occur withthe nitrene nitrogen at the 4-position, and not at the 3-position.Moreover, for 4PNO�, there are two types of stabilization that canbe envisioned to result from interaction between the nitreneradical anion and the N-oxide, shown in Fig. 2. In thep2s state, thepyridinium can serve as an electron

pair acceptor, and the system can potentially be stabilized bycontribution from the quinone-like structure. Similarly, the s2pstate can be stabilized by radical delocalization, creating a highlystable nitroxyl radical. Electron delocalization in either state issuch that it eliminates the formal charge separation of then-oxides, which could provide additional stabilization.

We have carried out a computational study of the electronicstructures of3PNO� and 4PNO�, and an experimental investigationof their reactivity. Surprisingly, despite the possible interactionsbetween the nitrene radical anion andN-oxide in 4PNO�, wedonotfind significant differences in the computedelectronic structures orreactivities of the two isomers in the reactionwith NO. However, adramatic difference between the ions is found in the reaction withCS2, as the4-isomerundergoesreactivity that isnotobservedfor the3-isomer, attributed to differences in accessibility of the oxide endof the ion toward reaction. The results indicate that the resonance

interaction between the nitrene anion center and the n-oxideincreases the nucleophilicity of the oxygen in the n-oxide moiety.

2. Methods

2.1. Experimental procedures

The flowing afterglow used in the investigation has beenpreviously described [35]. Briefly, 3PNO� and 4PNO� radicalanions are generated in upstream end of flow tube by electronionization of corresponding azide precursor. The ions are cooled toambient temperature (298K) and carried downstream by heliumbuffer gas (0.400 Torr, flow (He) = 190 STP cm3/s), and are allowedto undergo ion–molecule reactionwith neutral reagent vapors. Theproduct ions are sampled through a 1mm orifice into a low-pressure triple–quadrupole mass filter and detected with anelectron multiplier. Reactions with mass selected ions can becarried out in the second quadrupole (Q2). Reactions in Q2 wereused to verify the reaction products observed in the flow tube. Inthese experiments, the Q2 dc pole offset was kept very low (�1–2V, laboratory frame) tominimize the possibility of translationallydriven reactions. The energy dependencies of the observedreaction products were examined to ensure that their intensitiesare maximized at nearly thermal collision energies and drop off athigher energies as expected for exothermic processes.

Reaction kinetics are determined by monitoring the depletionof reactant ion as a function of neutral flow rate for sampleintroduced at a fixed distance from the nose cone. Pseudo-firstorder reaction rate constants, krxn, are obtained from a logarithmicplot of ion depletion vs reagent flow rate. Reaction rates arereported as reaction efficiencies (eff), which are the ratios of themeasured rate constant to the collisional rate constant,kcoll, calculated by using the parameterized-trajectory methoddescribed by Su and Chesnavich [36]. Absolute uncertaintiesin measured rate constants are estimated to be �50%. Branchingratios in reactions with multiple observed products aredetermined by measuring the branching ratios at multipleneutral flow rates, and extrapolating to zero reagent flow.Uncertainties in branching ratios are estimated to be �10% onan absolute basis.

[(Fig._2)TD$FIG]

Fig. 2. Resonance structures of the P2s and s2P states of 4PNO�.

[(Fig._1)TD$FIG]

Fig. 1. Non-bonding molecular orbitals in phenylnitrene.

70 D. Koirala et al. / International Journal of Mass Spectrometry 378 (2015) 69–75

164

Page 185: Mass spectrometric characterization of remotely charged

2.2. Materials

The azidopyridine-n-oxide precursors were prepared usingpublished procedures [37–39]. The synthesis and characterizationof the samples used in this work has been reported previously [40].Other materials were obtained from commercial suppliers andused as received.

2.3. Computational methods

Structures and energies of the p2s and s2p states of 3PNO�

and 4PNO�, NO, CS2 and possible reaction products werecalculated at the B3LYP/6-31 +G* level of theory. Reaction energiescorrespond to differences in electronic energies between reactantsand products, and are not corrected for zero-point energies orthermal energy differences. Calculations were carried out by usingQChem [41], using the resources of the Center forComputational Studies of Open-Shell and Electronically ExcitedSpecies (iopenshell.usc.edu).

3. Results and discussion

In this section, we report the results of flowing afterglowstudies of the reactivity of 3PNO� and 4PNO� with nitric oxide(NO) and carbon disulfide (CS2). These reagents have been shownpreviously [18] to form characteristic products with nitrene radicalanions. For example, the N–O transfer that occurs in the reaction ofBzN� with NO (Eq. (1)) also occurs with PhN�, resulting in theformation of phenoxide anion [18]. With CS2, aromatic nitreneanions have been found to react by addition, and by C and CSabstraction to form S2� and S�, respectively. Similar reactionsshould be expected for 3PNO� and 4PNO�.

3.1. Ion formation

The 3PNO� and 4PNO� radical anions are formed in high yieldsby dissociative electron attachment to the corresponding azides,with m/z 108 (Eq. (2)).

(2a)

[TD$INLINE]

(2b)

[TD$INLINE]The most significant impurity observed in these experiments

(with both isomers) is an ionwithm/z 92, which is presumably thenon-oxidized pyridinylnitrene radical anion. The relative signal ofthem/z 92 ion is variable from day to day, but decreases with timeduring the course of an experimental run, indicating that it mostlikely results from ionization of a non-oxidized azidopyridineimpurity.

3.2. Ion electronic structures

Before addressing the reactivity results, we first consider thecomputed electronic structures of the ions. As described in the

introduction, aromatic nitrene radical anions can have either p2sor s2p electronic structures. Previous studies of substitutedphenylnitrene anions [13] have found that the p2s states arefavored. However, as shown in Fig. 2, resonance interactionswithinthe states could affect the relative energies.

At the B3LYP/6-31 +G* level of theory, both 3PNO � and 4PNO�

are predicted to have robust p2s ground states. The 2B1 state of4PNO� is calculated to be 13.2 kcal/mol lower in energy than the2B2 state, whereas the 2A0 state of 3PNO� is predicted to be16.0 kcal/mol lower in energy than the 2A00 state. The smalldifference in state energies of the sp2 can likely be attributed tothe preferential stability of the nitroxyl radical (Fig. 2), but thatstabilization is not large enough to overcome the benefits ofcharge delocalization. Nonetheless, the presence of the n-oxidegroup is not calculated to alter the ground state of the nitreneanion.

3.3. Reactivity studies

The results of our reactivity studies of ions 3PNO� and 4PNO�

are shown in Table 1. Overall, there are no measurable differencesin the rates of the reactions for the two isomers. The efficiencies ofthe reactions with NO (approximately 0.2) are much higher thanthose found for reaction with CS2, but similar to that reportedpreviously for the reaction of NO with BzN� (eff = 0.15) [18].Although there are no differences in the reaction rates for the twoisomers with these reagents, there are very large differences in theobserved products. In the sections below, we consider thedifferences in the products that are formed.

3.3.1. Reaction with NOWith nitric oxide, both 3PNO� and 4PNO� are observed to

undergoN–O exchange, similar towhat is observedwith PhN� andBzN� [18]. Many additional products are observed for the reactionof 3PNO� with NO, but, as shown in Table 1, they are assigned tosecondary fragmentation of the phenoxide anion. Possible productstructures are shown in Eq. (3)

(3).

[TD$INLINE]A notable difference in the reactivity of 3PNO� and 4PNO�with

NO is that reaction with 4PNO� leads to significant amount ofadduct ion, whereas only a trace is observed with 3PNO�. NOaddition was not reported for either PhN� or BzN� [18], and isgenerally associated with reactions of closed-shell anions [42,43],Unfortunately, the structures of the NO adducts are not known.However, charge distribution calculations reported below find thatthere is more charge on the oxygen in 4PNO� than in 3PNO�,which raises the possibility that the difference in the extent ofadduct formation is due to 4PNO� forming an adduct at theoxygen, as opposed to at the nitrogen.

D. Koirala et al. / International Journal of Mass Spectrometry 378 (2015) 69–75 71

165

Page 186: Mass spectrometric characterization of remotely charged

Both n-oxide anions are found to react with NO by oxygen atomtransfer to form the pyridinylnitrene radical anion, as shown for4PNO� in Eq. (4). Although the branching ratios for these reactionsin the flow tube cannot be determined due to the presence ofimpurity in ions, they were verified as products by using thereaction of NO with mass-selected ions in Q2.

(4)

[TD$INLINE]If the N�O BDEs in 3PNO� and 4PNO� are similar to that in

pyridine-n-oxide (approximately 72 kcal/mol) [44], then theoxygen atom transfer reactionwould be essentially thermoneutral.Alternatively, given that the reaction is carried out in Q2, albeitunder very low energy conditions, the oxygen atom transfer couldbe slightly endothermic and translationally driven.

3.3.2. Reaction with CS2Although the rates at which 3PNO� and 4PNO� react with CS2

are similar, there are significant differences between the reactionproducts. In particular, some products observed with 4PNO�

involve loss of the oxygen atom, and these products are notobserved with 3PNO�. One product involving the oxygen isobserved at m/z 124, which is 16Da higher than the reactant. Themost likely assignment for this product is that it arises fromaddition of sulfur and loss of oxygen, i.e., a sulfur–oxygenexchange. The resulting products would be the n-sulfide andcarbonyl sulfide, OCS.

The ion signal atm/z92 is also likelydue to the reaction involvingoxygen. As in the reaction with NO, we are unable to quantify theyield of the m/z 92 product due to background presence of thepyridinylnitrene radical anion. However, considering 4PyN� is alsoobservedtoreactwithCS2, it appears thatm/z92 is, in fact, formedinrelatively high yield (comparable to the yield of S2�). This is alsoindicated by the results on “clean” days, where the mass spectrumincludesminimal amounts of the impurity. The formation ofm/z 92in the reaction of 4PNO� with CS2 was also confirmed by usingmass-selected ions in Q2.

The identity of them/z 92 is not known unequivocally. Althoughthe ion with m/z 92 could be the pyridinylnitrene radical anion,there is a second possibility. The ion CS2O� is isobaric with thepyridinylnitrene radical anion, and could, in principle be formed byoxygen-anion transfer with 4PNO�, as shown in Eq. (5). Ourexperimental results suggest both structures are present.

(5)

[TD$INLINE]Evidence for the formation of CS2O� comes from the isotopic

distribution of the product. Fig. 3 shows amass spectrum of 4PNO�

taken on an exceptionally clean day (with minimal m/z 92contamination), with and without the addition of CS2. Thebackground signal of m/z 92 in the spectrum is less than 10kcps.However, upon addition of CS2, the signal increases to nearly100kcps. Most importantly, the signal at m/z 94 also increases, tonearly 8% of them/z 92 signal, which agrees well with the value of9% expected for CS2O�. Therefore, the M +2 isotope peak indicatesthat there is sulfur present in the m/z 92 ion.

However, reactivity evidence suggests the presence of 4PyN� aswell. Fig. 4 shows the mass spectra for reaction of m/z 92 ion,formed by reaction of 4PNO� with CS2, with NO. The formation ofm/z 94 is clear evidence for the presence of a nitrene radical anion,4PyN�.

Computationally,bothreactionsarecalculatedtobeenergeticallyfavorable. At the B3LYP/6-31+G* level of theory, the formation ofCS2O� and the triplet pyridinylnitrene (Eq. (5)) is computed to beexothermic by 72kcal/mol, or more than 3eV! For oxygen atomtransfer, there are multiple thermochemically accessible pathways.Direct transfer to formCS2O (Eq. (6a)) is computed to be exothermicby 7.7 kcal/mol. A second pathway, shown in Eq. (6b), involves theformation of CO+ triplet S2, and is computed to be exothermic by9.9 kcal/mol.

(6a)

[(Fig._3)TD$FIG]

Fig. 3. The m/z 92 region of the mass spectra of 4PNO� before (dashed) and after(solid) addition of CS2.

Table 1Reaction efficiencies and product branching ratios for reactions of 3PNO� and4PNO� with NO and CS2

Reagent Result Ion assignment 3PNO– 4PNO–

NO Effa 0.21 0.20m/z 138 [M+NO]� <1% 18%m/z 110 [M+NO�N2]� 44% 82%m/z 82 [M+NO�N2�CO]� 21% Nb

m/z 80 [M+NO�N2�NO]� 20% Nb

m/z 61 [M+NO�C3H3�N2]� 14% Nb

m/z 92 [M+NO�NO2]� Yc Yc

CS2 Effa 0.03 0.02m/z 152 [M+CS2]� 83% 77%m/z 64 S2� 15% 15%m/z 32 S� <1% <1%m/z 58 SCN� 2% 2%m/z 124 [M+CS2–CSO]� Nb 7%m/z 92 [M�O]� Nb Yc

a Reaction efficiency, which corresponds to kexp/kADO.b Not observed for this isomer.c Product is observed in Q2, but the flow tube branching ratio cannot be

determined due to the presence of impurities.

72 D. Koirala et al. / International Journal of Mass Spectrometry 378 (2015) 69–75

166

Page 187: Mass spectrometric characterization of remotely charged

[TD$INLINE]

(6b)

[TD$INLINE]Regardless of which reaction occurs, oxygen atom transfer to

form4PyN�or oxygen ion transfer to formCS2O�, theyboth involvereaction at the oxygen. Deoxygenation of tertiary-amine-n-oxidesin solution [45–47] has been proposed to occur by a mechanismsuch as that shown in Scheme 1, leading to the formation of thedithiiranone as in Eq. (6a). In solution, the dithiiranone can behydrolyzed to CO2 +HSSH [45] or can be utilized as a sulfur transferreagent [48].

In the gas phase, direct reaction of oxygen atom with CS2 givesCS+ SO as the major products [49]. However, CO has also beenobserved in the reaction [50], indicating that CO+S2 is a possibledecomposition pathway of CS2O. In fact, CO+S2 is energetically themost favored pathway [49] but is slow because there is a slight(6.6 kcal/mol) barrier [49] for the initial formationofCS2O.However,thisbarriercanbeovercomeinthereactionofCS2with4PNO�bytheenergy releasedupon formationof the initial ion/molecule complexin combination with the oxygen atom transfer exothermicity.Therefore, oxygen atom transfer to form CO and S2 products shouldbe energetically accessible for 4PNO�. Formation of CS + SO isapproximately3 eV less favorable [49] than formationofCO+S2andis therefore highly endothermic in the reaction of 4PNO� with CS2.

As noted above, sulfur–oxygen exchange is also observed in thereaction of CS2 with 4PNO�. The mechanism of sulfur–oxygenexchange likely involves a first step similar to that for oxygen atomor ion transfer, addition of the oxygen to the center carbon of CS2,as shown in Scheme 2. However, instead of forming the disulfidebond as in dithiiranone formation, a N��S bond is formed.Alternatively, homolytic cleavage of the N��O bond in the CS2adduct would result in formation of CS2O�. Therefore, all threereaction pathways (oxygen atom transfer, oxygen ion transfer,oxygen–sulfur exchange) occur addition of the oxygen to CS2.

A small amount of NCS� product is observed in the reaction ofCS2 with both 3PNO� and 4PNO�. NCS� has been observedpreviously in reactions of CS2 with closed-shell, nitrogen-basedanions [43,51,52], but has not been reported previously forreactions of open-shell anions. The mechanism is presumablysimilar to that shown in Scheme 2, although occurring at thenitrogen. The other products observed with CS2 (S�, S2� and CS2adduct) are similar to those observed previously with nitreneradical anions [18].

3.4. Comparison of isomers: oxygen nucleophilicity

There are significant differences in the reactions that occurbetween ions 3PNO� and 4PNO� and CS2. Specifically, 4PNO�

undergoes sulfur–oxygen exchange and oxygen-atom and/oroxygen-anion transfer reactions that are not observedwith 3PNO�.The common feature of these reactions is that they all involveinitial nucleophilic attack of the oxygen in the anion at the carbonof the carbon disulfide. The fact that 4PNO� undergoes reaction atthe oxygenwhereas 3PNO� does not reflects important differencesin their electronic structures.

The best example that illustrates the electronic structuredifferences between the ions is the oxygen-transfer reactionwith CS2. Although carbon disulfide is known to deoxygenatetertiary-amine-n-oxides [45], the reaction is generally notobserved with aromatic-n-oxides. Therefore, the lack of oxygentransfer with 3PNO� is consistent with the expectation that thesubstituent in themeta position does not interact with the n-oxidemoiety. Similarly, the nitrene anion para to the n-oxide can serve asa p-donor, as shown in Fig. 2, which makes the oxygen in the p2sstatemore nucleophilic. Hammett analysis of the deoxygenation ofaniline-n-oxides in solution has shown [45] that the reaction isfavored by strong electron donating substituents, which increasethe nucleophilicity of the oxygen. Apparently, the nitrene radicalanion is such a strongp-electron donor that it even enables oxygentransfer in the normally inert aromatic-n-oxides.

The difference in the reactivity cannot be attributed todifferences in overall thermochemistry for the reactions. All of

[(Fig._4)TD$FIG]

Fig. 4. The m/z 92 region of the mass spectra of 4PNO� reaction with CS2 (dashed)and after addition of NO (solid).

[(Scheme_1)TD$FIG]

Scheme 1.

[(Scheme_2)TD$FIG]

Scheme 2.

D. Koirala et al. / International Journal of Mass Spectrometry 378 (2015) 69–75 73

167

Page 188: Mass spectrometric characterization of remotely charged

the reactions observed for 4PNO� are also computed to beexothermic for 3PNO�, although they do not occur. For example,the sulfur–oxygen exchange reaction observed for 4PNO�

(Scheme 2) is computed to be exothermic by 45kcal/mol. Similarly,the sulfur–oxygen exchange reaction for 3PNO� is computed to beexothermic by 37kcal/mol, but does not occur at all. Therefore, thedifference in reactivity is more likely due to kinetic differences thatresult from differences in electronic structure.

The increased nucleophilicity of the oxygen in 4PNO� apparentin the reactivity studies is consistent with the computed chargedistributions in the p2s states. The B3LYP/6-31 +G* computedMullikin charges at the heavy atoms in 3PNO� and 4PNO� areshown in Fig. 5.

Although both ions have increased charge density at the oxygenthan is found in pyridine-n-oxide there are some differences incharge densities at the carbon atoms, the most significantdifference is for the oxygen, where the calculated charge is largerin in 4PNO� than in 3PNO�, which accounts for the increasedreactivity at that site. Increased reactivity at the oxygen accountsfor the observation of either oxygen atom (Eq. (6)) or oxygen ion(Eq. (5)) transfer with CS2.

As suggested above, the difference in the charge density at theoxygen may also account for the difference in the extent of adductformation with nitric oxide. This is most likely the case if theadduct is an electrostatic complex. However, as noted, thestructure of the adduct is not known. Aside from the extent ofadduct formation, there is little difference in the reactivity of withnitric oxide. The differences in the observed products can beattributed to differences in the ability of the resulting phenoxideion to fragment.

4. Conclusion

The reactivity of 3PNO� and 4PNO�, particularlywith CS2, showthat there are significant differences in their electronic structures,which can be understood as resulting from the resonanceinteraction between the nitrene anion and the n-oxide moietyin 4PNO�, and the lack thereof in 3PNO�. In the p2s state, themonovalent nitrogen anion is a strong p electron donor, whichincreases the charge density on the oxygen, as shown in Fig. 2. Theresult is consistent with the conclusions based on condensed-phase studies that oxygen atom transfer is favored by p-donors[45]. However, the reaction has not been observed previously foraromatic n-oxides. This work shows that oxygen atom transfer foraromatic n-oxides can occur with sufficiently strongp donors. Thedifferences in the electronic structures of 3PNO� and 4PNO� donot result in differences in reactivity with NO, although there aredifferences in the stabilities of the resulting products.

Acknowledgements

This paper is dedicated to Professor Veronica Bierbaum, inappreciation for all she has taught us about ion chemistry and

flowing afterglow kinetics. The work is supported by the NationalScience Foundation (CHE11-11777).

References

[1] M. Born, S. Ingemann, N.M.M. Nibbering, Reactivity of mono-halogen carbeneradical anions (CHX��; X = F, Cl and Br) and the corresponding carbanions(CH2X; X=Cl and Br) in the gas phase, J. Chem. Soc Perkin Trans. 2 (1996)2537–2547.

[2] J.J. Grabowski, S.J. Melly, Formation of carbene radical anions: gas-phasereaction of the atomic oxygen anion with organic neutrals, Int. J. MassSpectrom. Ion Processes 81 (1987) 147–164.

[3] R.N. McDonald, Generation, thermochemistry, and chemistry of carbene anionradicals and related species, Tetrahedron 45 (1989) 3993–4015.

[4] G.D. Sargent, C.M. Tatum Jr., S.M. Kastner, Carbene radical anions. New speciesof reactive intermediate, J. Amer. Chem. Soc. 94 (1972) 7174–7176.

[5] S.M. Villano, N. Eyet,W.C. Lineberger, V.M. Bierbaum, Gas-phase carbene radicalanions: new mechanistic insights, J. Am. Chem. Soc. 130 (2008) 7214–7215.

[6] S.M. Villano, N. Eyet, W.C. Lineberger, V.M. Bierbaum, Gas-phase reactions ofhalogenated radical carbene anionswith sulfur and oxygen containing species,Int. J. Mass Spectrom. 280 (2009) 12–18.

[7] R.N. McDonald, A.K. Chowdhury, Hypovalent radicals. 7. Gas-phase generationof phenylnitrene anion radical and its reactionwith phenyl azide, J. Am. Chem.Soc. 102 (1980) 5118–5119.

[8] R.N. McDonald, A.K. Chowdhury, D.W. Setser, Gas-phase generation ofphenylnitrene anion radical – proton affinity and DHf of PhN– and itsclustering with ROH molecules, J. Am. Chem. Soc. 103 (1981) 6599–6603.

[9] P.S. Drzaic, J.I. Brauman, A determination of the triplet-singlet splitting inphenylnitrene via photodetchment spectroscopy, J. Am. Chem. Soc. 106 (1984)3443–3446.

[10] P.S. Drzaic, J.I. Brauman, Electron photodetachment from phenylnitrene,anilide, and benzyl anions. Electron affinities of the anilino and benzyl radicalsand phenylnitrene, J. Phys. Chem. 88 (1984) 5285–5290.

[11] R.N. McDonald, S.J. Davidson, Electron photodetachment of the phenylnitreneanion radical: EA, DH�

f, and the singlet-triplet splitting for phenylnitrene,J. Am. Chem. Soc. 115 (1993) 10857–10862.

[12] M.J. Travers, D.C. Cowels, E.P. Clifford, G.B. Ellison, Photoelectron spectroscopyof phenylnitrene anion, J. Am. Chem. Soc. 114 (1992) 8699–8701.

[13] N.R. Wijeratne, M.D. Fonte, A. Ronenius, P.J. Wyss, D. Tahmassebi, P.G.Wenthold, Photoelectron spectroscopy of chloro-substituted phenylnitreneanions, J. Phys. Chem. A 113 (2009) 9467–9473.

[14] M.J. Travers, D.C. Cowles, E.P. Clifford, G.B. Ellison, P.C. Engelking, Photoelectronspectroscopy of the CH3N– ion, J. Chem. Phys. 111 (1999) 5349–5360.

[15] D.E. Herbranson, M.D. Hawley, Electrochemical reduction of p-nitrophenylazide: evidence consistent with the formation of p-nitrophenylnitrene anionradical as a short-lived intermediate, J. Org. Chem. 55 (1990) 4297–4303.

[16] S. Murata, R. Nakatsuji, H. Tomioka, Mechanistic studies of pyrene-sensitizeddecomposition of p-butylphenyl azide: generation of nitrene radical anionthrough a sensitizer-mediated electron transfer from amines to the azide,J. Chem. Soc. Perkin Trans. 2 (1995) 793–799.

[17] D.A. Van Galen, J.H. Barnes, M.D. Hawley, The electrochemical reduction offluorenone tosylhydrazone. Evidence consistent with the formation to thetosyl nitrene anion radical, J. Org. Chem. 51 (1986) 2544–2550.

[18] N.R. Wijeratne, P.G. Wenthold, Structure and reactivity of benzoyl nitreneradical anion in the gas phase, J. Org. Chem. 72 (2007) 9518–9522.

[19] N.R.Wijeratne, P.G.Wenthold, Benzoylnitrene radical anion: a new reagent forthe generation of M-2H anions, J. Am. Soc. Mass Spectrom. 18 (2007)2014–2016.

[20] N.R. Wijeratne, P.G. Wenthold, Thermochemical studies of benzoylnitreneradical anion: the N–H bond dissociation energy in benzamide in the gasphase, J. Phys. Chem. A 111 (2007) 10712–10716.

[21] Z.P. Demko, K.B. Sharpless, A click chemistry approach to tetrazoles byHuisgen1,3-dipolar cycloaddition: synthesis of 5-acyltetrazoles from azides and acylcyanides, Angew. Chem. Int. Ed. 41 (2002) 2113–2116.

[22] Z.P. Demko, K.B. Sharpless, A click chemistry approach to tetrazoles byHuisgen1,3-dipolar cycloaddition: synthesis of 5-sulfonyl tetrazoles from azides andsulfonyl cyanides, Angew. Chem. Int. Ed. 41 (2002) 2110–2113.

[23] H.C. Kolb, M.G. Finn, K.B. Sharpless, Click chemistry: diverse chemical functionfrom a few good reactions, Angew. Chem. Int. Ed. 40 (2001) 2004–2021.

[24] C.W. Tornoe, C. Christensen, M. Meldal, Peptidotriazoles on solid phase:[1,2,3]-triazoles by regiospecific copper(I)-catalyzed 1,3-dipolar cycloadditionsof terminal alkynes to azides, J. Org. Chem. 67 (2002) 3057–3064.

[25] M. Al-Za'al, H.C. Miller, J.W. Farley, Observation of metastable states in theautodetachment continuum of the negative molecular ion amidogen ion(1�)(NH�), Chem. Phys. Lett. 131 (1986) 56–59.

[26] P.C. Engelking,W.C. Lineberger, Laser photoelectron spectrometry of imidogen(–) (NH�): electron affinity and intercombination energy difference inimidogen, J. Chem. Phys. 65 (1976) 4323–4324.

[27] J.W. Farley, High resolution laser spectroscopy ofmolecular ions, Proc. SPIE-Int.Soc. Opt. Eng. 1858 (1993) 92–100.

[28] K.K. Lykke, D.M. Murray, W.C. Neumark, High-resolution studies of autode-tachment in negative ions, Philos. Trans. R. Soc. Lond. A 324 (1988) 179–196.

[29] H.C. Miller, M. Al-Za'al, J.W. Farley, High-resolution measurement of theinfrared rotation–vibration spectrum of nitrogen hydride anion (NH1�),AIP Conference Procceedings 160 (1987) 354–355.

[(Fig._5)TD$FIG]

Fig. 5. Calculated distributions in 3PNO�, and 4PNO� and pyridine-n-oxide.Charges on hydrogen have been summed into heavy atoms.

74 D. Koirala et al. / International Journal of Mass Spectrometry 378 (2015) 69–75

168

Page 189: Mass spectrometric characterization of remotely charged

[30] H.C. Miller, M. Al-Za'al, J.W. Farley, Measurement of hyperfine structure in theinfrared rotation–vibration spectrum of amidogen ion(1�) (NH�), Phys. Rev.Lett. 58 (1987) 2031–2034.

[31] D.M. Neumark, K.R. Lykke, T. Andersen,W.C. Lineberger, Infrared spectrum andautodetachment dynamics of NH, J. Chem. Phys. 83 (1985) 4364–4373.

[32] S. Srivastava, N. Sathyamurthy, Ab initio potential energy curves for the groundand low lying excited states of NH� and the effect of 2S� states onL-doublingof the ground state X2P, J. Phys. Chem. A 117 (2013) 8623–8631 (andreferences therein).

[33] U. Hoper, P. Botschwina, H. Koppel, Theoretical study of the Jahn–Teller effectin X�2E CH3O, J. Chem. Phys. 112 (2000) 4132–4142.

[34] D.L. Osborn, D.J. Leahy, D.M. Neumark, Photodissociation spectroscopy anddynamics of CH3O and CD3O, J. Phys. Chem. A 101 (1997) 6583–6592.

[35] P.J. Marinelli, J.A. Paulino, L.S. Sunderlin, P.G. Wenthold, J.C. Poutsma, R.R.Squires, A tandem selected ion flow tube-triple quadrupole instrument,Int. J. Mass Spectrom. Ion Processes 130 (1994) 89–105.

[36] T. Su,W.J. Chesnavich, Parametrization of the ion-polar molecule collision rateconstant by trajectory calculations, J. Chem. Phys. 76 (1982) 5183–5185.

[37] H. Sawanishi, K. Tajima, T. Tsuchiya, Studies on diazepines. XXVIII. Synthesesof 5H-1,3-diazepines and 2H-1,4-diazepines from 3-azidopyridines, Chem.Pharm. Bull. 35 (1987) 4101–4109.

[38] T. Itai, S. Kamiya, Potential anticancer agents. II. 4-Azidoquinoline and4-azidopyridine derivatives, Chem. Pharm. Bull. 9 (1961) 87–91.

[39] T. Itai, S. Kamiya, Potential anticancer agents. XI. Synthesis of 4- and5-azidopyridazine 1-oxide, Chem. Pharm. Bull. 11 (1963) 1059–1064.

[40] K.N. Crabtree, K.J. Hostetler, T.E. Munsch, P. Neuhaus, P.M. Lahti, W. Sander, J.S.Poole, Comparative study of the photochemistry of the azidopyridine 1-oxides,J. Org. Chem. 73 (2008) 3441–3451.

[41] Y. Shao, L.F. Molnar, Y. Jung, J. Kussmann, C. Ochsenfeld, S.T. Brown, A.T.B.Gilbert, L.V. Slipchenko, S.V. Levchenko, D.P. O’Neill, R.A. DiStasio, R.C. Lochan,T. Wang, G.J.O. Beran, N.A. Besley, J.M. Herbert, C.Y. Lin, T. Van Voorhis, S.H.Chien, A. Sodt, R.P. Steele, V.A. Rassolov, P.E. Maslen, P.P. Korambath, R.D.Adamson, B. Austin, J. Baker, E.F.C. Byrd, H. Dachsel, R.J. Doerksen, A. Dreuw, B.D. Dunietz, A.D. Dutoi, T.R. Furlani, S.R. Gwaltney, A. Heyden, S. Hirata, C.P. Hsu,G. Kedziora, R.Z. Khalliulin, P. Klunzinger, A.M. Lee, M.S. Lee, W. Liang, I. Lotan,

N. Nair, B. Peters, E.I. Proynov, P.A. Pieniazek, Y.M. Rhee, J. Ritchie, E. Rosta, C.D.Sherrill, A.C. Simmonett, J.E. Subotnik, H.L. Woodcock, W. Zhang, A.T. Bell, A.K.Chakraborty, D.M. Chipman, F.J. Keil, A. Warshel, W.J. Hehre, H.F. Schaefer, J.Kong, A.I. Krylov, P.M.W. Gill, M. Head-Gordon, Advances in methods andalgorithms in a modern quantum chemistry program package, Phys. Chem.Chem. Phys. 8 (2006) 3172–3191.

[42] S.A. Chacko, P.G. Wenthold, The negative ion chemistry of nitric oxide in thegas phase, Mass Spectrom. Rev. 25 (2006) 112–126.

[43] N.J. Rau, E.A. Welles, P.G. Wenthold, Anionic substituent control of theelectronic structure of aromatic nitrenes, J. Am. Chem. Soc. 135 (2013)683–690.

[44] H.Y. Afeefy, J.F. Liebman, S.E. Stein, Neutral thermochemical data, in: P.J.Linstrom, W.G. Mallard (Eds.), NIST Chemistry WebBook, NIST StandardReference Database Number 69, National Institute of Standards andTechnology, Gaithersburg, MD 20899, 2014 (retrieved 11.02.14).

[45] T. Yoshimura, K. Asada, S. Oae, Deoxygenation of tertiary amine oxides withcarbon disulfide, Bull. Chem. Soc. Jpn. 55 (1982) 3000–3003.

[46] M. Hamana, B. Umezawa, S. Nakashima, Tertiary amine oxides. XIV. Reactionsof N-(p-dimethylaminophenyl)nitrones, having pyridine, quinoline or its N-oxide, as a-substituents, with carbon disulfide, Chem. Pharm. Bull. (Tokyo) 10(1962) 969–974.

[47] J. Nakayama, A. Ishii, Chemistry of dithiiranes, 1,2-dithietanes, and1,2-dithietes, Adv. Heterocycl. Chem. 77 (2000) 221–284.

[48] M.F. Zipplies, M.J. De Vos, T.C. Bruice, The formation of thiiranes fromolefins inthe course of the deoxygenation of tertiary amine N-oxides by carbondisulfide, J. Org. Chem. 50 (1985) 3228–3230.

[49] V. Saheb, Quantum chemical and theoretical kinetics study of the O(3P) +CS2reaction, J. Phys. Chem. A 115 (2011) 4263–4269 (and references therein).

[50] J.W. Hudgens, J.T. Gleaves, J.D. McDonald, Infrared chemiluminescence studiesof the reactions of oxygen (3P) atoms with carbon disulfide and carbonmonosulfide, J. Chem. Phys. 64 (1976) 2528–2532.

[51] S.E. Barlow, V.M. Bierbaum, Reactions of O–+N2O at 300K: the totally labeledexperiments, J. Chem. Phys. 92 (1990) 3442–3447.

[52] C.H. DePuy, Fragmentation of organic anions induced by exothermic additionreactions, Org. Mass Spectrom. 20 (1985) 556–559.

D. Koirala et al. / International Journal of Mass Spectrometry 378 (2015) 69–75 75

169