+ All Categories
Home > Documents > New aspects of tetramethylene initiation in polymer chemistry....Order Number 9024500 New aspects of...

New aspects of tetramethylene initiation in polymer chemistry....Order Number 9024500 New aspects of...

Date post: 08-Feb-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
149
New aspects of tetramethylene initiation in polymer chemistry. Item Type text; Dissertation-Reproduction (electronic) Authors Clever, Hester Ann. Publisher The University of Arizona. Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author. Download date 29/06/2021 13:06:18 Link to Item http://hdl.handle.net/10150/184988
Transcript
  • New aspects of tetramethyleneinitiation in polymer chemistry.

    Item Type text; Dissertation-Reproduction (electronic)

    Authors Clever, Hester Ann.

    Publisher The University of Arizona.

    Rights Copyright © is held by the author. Digital access to this materialis made possible by the University Libraries, University of Arizona.Further transmission, reproduction or presentation (such aspublic display or performance) of protected items is prohibitedexcept with permission of the author.

    Download date 29/06/2021 13:06:18

    Link to Item http://hdl.handle.net/10150/184988

    http://hdl.handle.net/10150/184988

  • INFORMATION TO USERS

    The most advanced technology has been used to photograph and

    reproduce this manuscript from the microfilm master. UMI films the

    text directly from the original or copy submitted. Thus, some thesis and

    dissertation copies are in typewriter face, while others may be from any

    type of computer printer.

    The quality of this reproduction is dependent upon the quality of the

    copy submitted. Broken or indistinct print, colored or poor quality

    illustrations and photographs, print bleedthrough, substandard margins,

    and improper alignment can adversely affect reproduction.

    In the unlikely event that the author did not send UMI a complete

    manuscript and· there are missing pages, these will be noted. Also, if

    unauthorized copyright material had to be removed, a note will indicate

    the deletion.

    Oversize materials (e.g., maps, drawings, charts) are reproduced by

    sectioning the original, beginning at the upper left-hand corner and

    continuing from left to right in equal sections with small overlaps. Each

    original is also photographed in one exposure and is included in

    reduced form at the back of the book.

    Photographs included in the original manuscript have been reproduced

    xerographically in this copy. Higher quality 6" x 9" black and white

    photographic prints are available for any photographs or illustrations

    appearing in this copy for an additional charge. Con~act UMI directly

    to order.

    U·M·I UnlverSily Microfilms International

    A 8811 & Howell Information Company 300 North Zeeb Road. Ann Arbor. MI 48106·1346 USA

    313761·4700 800521·0600

  • Order Number 9024500

    New aspects of tetramethylene initiation in polymer chemistry

    Clever, Hester Ann, Ph.D.

    The University of Arizona, 1990

    U·M·I 300 N. Zceb Rd. Ann Arbor, MI 48106

  • " .'f

    NEW ASPECfS OF TETRAMETHYLENE INITIATION

    IN POLYMER CHEMISTRY

    by

    Hester Ann Clever

    A Dissertation Submitted to the Faculty of the

    DEPARTMENT OF CHEMISTRY

    In Partial Fulfillmeut of the Requirements For the Degree of

    DOCfOR OF PHILOSOPHY

    In the Graduate College

    THE UNIVERSITY OF ARIZONA

    1990

  • THE UNIVERSITY OF ARIZONA GRADUATE COLLEGE

    As members of the Final Examination Committee, we certify that we have read

    the dissertation prepared by Hester A. Clever

    entitled New Aspects of Tetramethylene Initiation in

    Polymer Chemistry

    and recommend that it be accepted as fulfilling the dissertation requirement

    for the Degree of Doctor of Philosophy

    /I. J(, flrc~1. J -'. Date

    '3 J:/ztJ , 3-7-90

    Date

    3/t11D Date

    iMitJ ? I J 717 Date J

    7 !J7a'I--Ce:. 9iJ Date

    Final approval and acceptance of this dissertation is contingent upon the candidate's submission of the final copy of the dissertation to the Graduate College.

    I hereby certify that I have read this dissertation prepared under my direction and recommend that it be accepted as fulfilling the dissertation requirement.

    Dissertation Director Date

  • STATEMENT BY AUTHOR

    This dissertation has been submitted in partial fulfillment of requirements for an advanced degree at The University of Arizona and is deposited in the University Library to be made available to bonowers under rules of the Library.

    3

    Brief quotations from this dissertation are allowable without special permission, provided that accurate acknowledgement of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgment the proposed use of the material is in the interests of scholarship. In all other instances, however, permission must be obtained from the author.

  • 4

    DEDICATION

    To my parents, Donald and Marilyn Clever, who made this all possible.

  • ACKNO~DGEMENTS

    The author would like to thank Dr. H. K. Hall, Jr. for giving me the

    opportunity to wolk in his laboratory and for all of his advice and encouragement

    through out my graduate career. I would also like to thank Dr. Anne Padias for all of

    her advice, encouragement and patience.

    I would like to acknowledge Dr. Tzepei Tien for his helpful discussions

    concerning the molecular orbital calculations, and Dr. Michael Barfield and his NMR

    class for running many of the decoupled spectra required for analysis of the substituted

    cyclohexenes.

    5

  • -.

    6

    TABLE OF CONTENTS

    Page

    UST OF ll..LUSTRATIONS

    UST OF TABLES

    ABSTRACT

    9

    11

    13

    1. INTRODUCTION • • • • • • 15 The Bond Forming Initiation Theory •••• • • 16 An Organic Chemist's Periodic Table. • • • • • • 18 Applications of the Bond Forming Initiation Theory 20

    Trapping of a Zwitterionic Intermediate Using Cationic Polymerization • • • • • • • • • 20

    Trapping of a Zwitterionic Intermediate Using Anionic Polymerization • • • • • • • • • • • • 22

    Trapping of Diradical Intermediates • • • • 22 Borderline Tetramethylenes ••••••••••• 26 Reactions with p-Leaving Group Olefins 27

    Application of the Bond Forming Initiation Theory to [4+2] Systems ••••••••••••••••• 29

    Proposed Research ••••••••••••••••• 35

    2. REACTIONS OF l-ARYL-l,3-BUTADIENES WITH INSUBSTITUTED ELECTROPHll..IC OLEFINS ••••

    Introduction ••••••••••••••••••• Background •••••• Results • • • • • •

    Control Reactions • • • • • • • • • • • • • • • Reactions of l-AryI-l,3-Butadienes with Tri-

    substituted Electrophilic Olefins • • • • • Experimental Observations ••••••••• Stereochemistry of the Cycloadducts • • • • •

    Reactions with Olefins Containing p-Leaving Groups • Experimental Observations ••••• Homopolymer Cycloadduct • • • • • • • • • •

    Discussion • • • • • • • • • • • Conclusion • • • • • • • • • • • • • • • •

    3. REACTIONS OF l-METHOXY-l,3-BUTADIENE • Introduction •••••• • • • • Background ••••• Results • • • • • • • • • • •

    Control Reactions

    37 37 38 39 39

    41 41 44 45 45 45 47 49 53

    55 55 55 S8 S8

  • 4.

    5.

    6.

    7

    TABLE OF CONTENTS - Continued

    Page

    Reaction of I-Methoxy-l,3-Butadiene and Maleic Anhydride. • • • • • • • • • • • • • • 58

    Reaction of I-Methoxy-l,3-Butadiene and Methyl 3,3-Dicyanoacrylate • • • • • • • • • • • • • • • •• 60

    Reaction of I-Methoxy-l,3-Butadiene and Dimethyl Cyanofumarate • • • • • • • • • • • • • 62

    Reaction of I-Methoxy-l,3-Butadiene and Tri-methyl EthylenetricaIboxylate • • • • • • • • 62

    Reactions of I-Methoxy-l,3-Butadiene and Acrylonitrile • • • • • • • • • • • • • • • • • • •• 67

    Reactions of I-Methoxy-l,3-Butadiene with Olefins . ~ontaining p-Leaving Groups • • • • • • • . • 67

    DIscussIon • . • • • • • . • • • • • • • • • • • • 67 Conclusion • • • • • • . • • • . • • • • • • • • 69

    REACTIONS OF ALKYL BUTADIENES AND ACRYLONITRll..E Introduction • • • • • • • • • • • • • Background • . . • • • • • • • • • • • • • Results •.•••.•••••.••••.• • • • •

    Control Reactions •••••••••••••••• Reactions of Cyclopentadiene with Acrylonitrile Reactions of Cyclohexadiene with Acrylonitrile Reactions of Isoprene with Acrylonitrile • • • Reactions of 1,3-Butadiene with Acrylonitrile • Reactions of trans-Pipenylene with Acrylonitrile Reactions of 2,3-Dimethyl-l,3-Butadiene with

    Acrylonitrile • • • • • • • • . • • • Reactions of Other Dienes with Acrylonitrile • Reactions of Dienes with Methyl Acrylate •

    Discussion • • • • • • • • • • • • • • • • • • • • Conclusion • • • • • • • • • • • • • • • •

    A COMPARISON OF THEORETICAL AND EXPERIMENTAL DATA •• Introduction. • • • • • • • • • • • • Background Results Discussion • • • • • • • • • • • • • • • Conclusion •

    APPLICATION OF THE BOND-FORMING INITIATION THEORY TO GROUP TRANSFER POLYMERIZATION • • • •

    Introduction/Background • • • • • Mechanism Studies. •

    Results and Discussion • • • • • •

    71 71 71 73 73 73 73 73 75 77

    77 80 80 83 85

    86 86 86 87 94

    102

    103 103 106 111

  • 8

    TABLE OF CONTENTS - Continued

    7. CONCLUSIONS • • •

    Page

    114

    8. EXPERIMENTAL... 116 Instrumentation Solvents • • • • • • • • • • • • 116 Dienes •••• • • • • • • • • • • 116

    Reagents • • • • • • • • • • • • • • • • • • • • • • 117 Synthesis of 1-Phenyl-1,3-Butadiene. • 117 Synthesis of p-Anisyl-1,3-Butadiene. • 118

    Olefins •••••••••••••• • • • • • • 119 Synthesis of Methyl 3,3-Dicyanoacrylate • • • • • 119 Synthesis of Methyl Glyoxylate Methyl Hemiacetal. • 119 Synthesis of Dimethyl Cyanofumarate • • • • • • • 120 Synthesis of Trimethyl Ethylenetricarboxylate. 120 Synthesis of 2,2-Dicyanovinyl Chloride. • • • • • 120 Synthesis of 2,2-Dicyanovinyl Iodide • • • • • • • • • • • 121 Synthesis of 2,2-Dicyanovinyl Tosylate. • • • • • • 121 Synthesis of 2-Carbomethoxy-2-Cyanovinyl Chloride 121 Synthesis of 2-Carbomethoxy-2-Cyanovinyl Iodide 122

    Synthesis of 4-Cyano-l-Methylcyclohex-l-ene • • • • • 122 Control Reactions. • • • • • • • • • • • • • 123

    Stability of p-Anisyl-l,3-Butadiene • • • • • • • • • • 123 Triethyl Borane/Oxygen Initiation •••••• 123 Typical Procedure for Reactions of I-Phenyl-1,3-

    Butadiene and p-Anisyl 1.3-Butadiene with Tri-substituted Olefms 3-5 and p-Leaving Group Olefins 7-11, p-Anisyl-l,3-Butadiene with Maleic Anhydrides I-Methoxy-l,3-Butadiene with Olefins 3,4,7,10. • • • • • • • • 124

    Typical Procedure for the Reactions of I-Methoxy-l,3-Butadiene with Trimethyl Ethylenetricarboxylate, Reactions of I-Methoxy-l,3-Butadiene, 1,3-Butadiene, Isoprene, trans-Piperylene, 2,3-Dimethyl-l,3-Butadiene, cis-Piperylene, 2,5-Dimethyl-2,4-Hexadiene, Cyclo-pentadiene and 1,3-Cyclohexadiene with Acrylonitrile and Reactions of Isoprene and Piperylene with Methyl Acrylate ••• • • • • • • • • • • • • • • • • • •

    Group Transfer Polymerization. • • • • • • Group Transfer Polymerization Procedure Photopolymerizations • • • • • • • Data for Polymers and Cycloadducts • • • •

    APPENDIX A ••

    REFERENCES .•

    124 125 125 125 126

    134

    140

  • LIST OF n.LUSTRATIONS

    Figure/Schemes

    Scheme 1

    Figure 1

    Scheme 2

    Scheme 3

    Bonding FOnning Initiation Theory •

    An Organic Chemist's Periodic Table' •

    Trapping of a Zwitterionic Intennediate by Cationic Polymerization • • • • • • • • • • • • • •

    Trapping of a Zwitterionic Intennediate by Anionic Polymerization • • • • • • • • • • • • •

    Scheme 4 Trapping of a Diradical Intennediate I. •

    Scheme 5 Trapping of a Diradical Intennediate IT

    Scheme 6 Mechanism in Trapping a Diradical Intennediate with TEf'dPO • • • • • • • • • •

    Scheme 7 Effect of Solvent on Product Distribution • •

    Scheme 8 BFI Theory Using p-Leaving Group Olefins

    Scheme 9 Reactions of Cyclohexadiene with Cis- and Trans-2-butenes • . • • • • • • • . • • • • • •

    Scheme 10 Reactions of l-Cyclopropyl-l,3-Butadiene and TCNE. •

    Scheme 11 Reactions of Cis and Trans Dienes. • • • •

    Scheme 12 The BFI Theory Applied to [4+2] Systems. •

    Scheme 13 General Mechanism of Reactions of Dienes with p-Leaving Group Olefins • •

    Figure 2

    Scheme 14

    Scheme 15

    Figure 3

    Scheme 16

    Major Isomer • • • • • •

    Fonnation of Hexatriene • •

    Concerted Mechanism • •

    Stepwise Mechanism • •

    Butler's Proposed Mechanism for the Spontaneous Copolymerization of l-Ethoxy-l,3-Butadiene and Acrylonitrile • • • • • • • • • • • • •

    9

    Page

    16

    19

    21

    23

    23

    24

    25

    26

    28

    32

    33

    34

    36

    38

    46

    47

    51

    52

    57

  • Scheme 17

    . Figure 4

    LIST OF ll..LUSTRATIONS - Continued

    Possible Mechanism for the Fonnation of Black Interactable Product. • • • • • • • • •

    Isomeric Cyclohexenes from the Reactions of Piperylene and Acrylonitrile ••••••

    Scheme 18 Group Transfer Polymerization • • •

    Scheme 19 Reversible Dissociative Mechanism. • •

    Scheme 20 Irreversible Dissociate Mechanism

    Scheme 21 Associative Mechanism • . • • •

    Scheme 22 Webster's Proposed Mechanism • •

    10

    Page

    69

    77

    103

    107

    107

    109

    110

  • Table

    1.

    2.

    LIST OF TABLES

    Bartlett's Results

    Control Reactions • •

    3. Reactions of l-AryI-l,3-Butadienes with Trisubstituted

    11

    Page

    31

    • •• 40

    Olefins • • • • • • • • • • • • • • • • • • • • • • • • • • 42

    4. Reactions of l-AryI-I,3-Butadienes with p-Leaving Group Olefms • • . • • • • • • • • • • • • •• • • • • • • • 43

    5.

    6.

    7.

    8.

    9.

    10.

    Isomer Ratios of [4+2] Cycloadducts in Reactions of p-Anisyl-l,3-Butadiene • • • • • • • • • • •

    AryI-l,3-Butadienes with p-Leaving Group Olefins •

    Reactions of I-Methoxy-l,3-Butadiene and Maleic Anhydride

    Reactions of I-Methoxy-l,3-Butadiene • • • • • • •

    Reactions of I-Methoxy-l,3-Butadiene with Trimethyl • Ethylenetricarboxylate

    Molecular Weight of MeOBD/I'NE Copolymers • •

    11. Molecular Weight of MeOBD/I'NE Copolymers as a Function

    • 46

    • • • • • 48

    •• 59

    • 61

    .64

    • • • • • 65

    of Time . • • . • • • • • • • • • • • • • • • • • • • 66

    •• 74 12. Reactions of Isoprene and Acrylonitrile. • • • • • • •

    13. Feed Ratio Effects of Spontaneous Polymerizations of Isoprene and Acrylonitrile • • • • • • • • • • • • • • • • • • • 76

    14.

    15.

    16.

    17.

    Reactions of 1,3-Butadiene and Acrylonitrile. • •

    Reactions of trans-Piperylene and Acrylonitrile. •

    Reactions of 2,3-Dimethyl-l,3-Butadiene and Acrylonitrile

    Reactions of Other Dienes and Olefins in Bulk

    • • • • • 78

    • • • 79

    81

    82

  • Table

    18.

    19.

    20.

    21.

    22.

    23.

    24.

    25.

    26.

    LIST OF TABLES - Continued

    Calculations for Tetrasubstiwted Olefms · · Calculations for Trisubstituted Olefins

    Calculations for p-Leaving Group Olefins. . . Calculations for Maleic Anhydride and Derivatives • · · Calculations for Tetrasubstiwted Quinodimethanes

    Calculations for Mono and Disubstiwted Olefins •

    Comparison of Calculated LUMO Values and Experimental Reduction Potentials of Trisubstiwted Olefins

    Comparison of Calculated and Experimental Data for Tetrasubstituted Oletins . . . . . . . . . Comparison of Calculated and Experimental Values Calculations for Increasing Cyano Substituted Olefins

    · ·

    12

    Page

    · · · · · 88 90

    · · · · 91 · · · · 92

    · · · · 93 9S

    · · · · . . 96

    · · · · · . . 98

    100

    27. Results of the Mechanism Swdy of Group Transfer Polymerization. • • • 101

    28. Polymerization Reaction of Methyl Acrylate and Styrene Monomers • 112

  • 13

    ABSTRACI'

    Hall's Bond Forming Initiation Theory, derived for [2 + 2] systems, was applied

    to [4 + 2] systems. Polymerizable electron-rich and electron-poor olefins were reacted

    to obtain evidence for reactive intermediates using polymerization as a trap.

    A diradical intermediate was successfully trapped in reactions of I-methoxy-l,3-

    butadiene with methyl 2,2-dicyanoacrylate, dimethyl cyanofumarate, trimethyl

    ethylenetricarboxylate, maleic anhydride and acrylonitrile. Diradical intermediates were

    also trapped in reactions of isoprene, 1,3-butadiene, 2,3-dimethyl-l,3-butadiene, cis and

    trans-piperylene and 2,S-dimethyl-2,4-hexadiene with acrylonitrile. Copolymers were

    obtained in all cases. Copolymerization was accompanied by [4 + 2] cycloaddition.

    A zwitterionic intermediate was successfully trapped in the reactions of p-anisyl-

    1,3-butadiene and l-phenyl-I,3-butadiene with various leaving groups in the P-position.

    Homopolymer of the arylbutadiene was obtained in all cases. Polymerization was

    initiated from a cationic species which arose from the elimination of the leaving group

    from a zwitterionic hexamethylene species. Polymerization was accompanied by [4 + 2]

    cycloaddition.

    Reactions of p-anisyl-l,3-butadiene and l-phenyl-I,3-butadiene with methyl 2,2-

    dicyanoacrylate, dimethyl cyanofumarate, trimethyl ethylenetricarboxylate gave only the

    [4 + 2] cycloadduct. No polymer was formed in this reaction and no evidence for a

    reactive intermediate was obtained.

    The BFI theory was also applied to Group Transfer Polymerization reactions to

    determine whether the reaction mechanism involved a tetramethylene diradical

    intermediate. No styrene was incorporated into the methyl acrylate polymer, and so no

    evidence for a radical intermediate was obtained.

  • 14

    LUMO energies calculated by AMI were compared to reduction potential and UV

    data for tetrasubstituted and trisubstituted electrophilic olefins. The trends exhibited by

    the calculated LUMO energies agreed well with the experimental data, implying that

    calculations may be a reasonable method of predicting reactivity.

  • CHAPTER 1

    Introduction

    15

    Throughout the years, organic chemists have been doing their best to ignore the

    polymeric byproducts in their reactions. They use reaction conditions that favor the

    small molecule products such as high reaction temperature, low c.oncentration and highly

    substituted reactants. If the reaction conditions still allow polymer formation, they add

    copious amounts of inhibitor to prevent polymerization. Polymer chemists, on the other

    hand, have done their best to ignore the small molecule products they often obtain in

    their polymerization reactions. They use low temperatures and high concentrations to

    favor polymerization. They immediately add initiator to the reaction mixture without

    first seeing if the polymerization may be spontaneous. Organic chemists often ignore

    the polymeric products found in the bottom of their flasks, while polymer chemists

    often throwaway the small molecule products. Both the organic chemists and the

    polymer chemists are losing mechanistic information by ignoring each others products.

    In efforts to explain the spontaneous polymerizations that often accompany [2 + 2]

    cycloadditions and to unify the organic chemistry and the polymer chemistry, Hall

    proposed the Bond Forming Initiation Theory shown in Scheme 1.1 According to the

    Woodward-Hoffmann rules,2 [2+2] cycloadditions are thermally forbidden, concerted

    reactions. When an electron-rich olefin reacts with an electron-poor olefin. a

    tetramethylene intermediate, called such because it contains four carbons, is formed.

    This intennediate closes to form the cycloadduct. Both zwitterionic or diradical

    tetramethylenes have been proposed as intermediates.'"

  • D

    =' - c(D A

    1

    ceD

    A

    --

    Scheme 1 Bond Forming Initiation Theol)'

    16

    ~ D A

    ~ D

    According to Huisgen, this tetramethylene lies on a continuous scale between

    zwitterionic and diradical structures, and may be regarded as a resonance hybrid of the

    two extreme forms. The predominant nature of the tetramethylene intermediate is

    determined by the terminal substituents.5.6 If they are strong donors and acceptors such

    as alkoxy, amino, or cyano, a tetramethylene with zwitterionic character is favored. If

    the termini are substituted with weaker donor and acceptor groups such as alkyl, phenyl,

    ~D A

    c(D A

  • 17

    or ester groups, a predominantly diradical tetramethylene is favored.

    The termini of the tetramethylene can interact with each other by through-bond

    interaction in both the gauche and the trans forms. Maximum interaction is achieved in

    the trans configuration by the carbon p-orbitals adopting a 90°-90° conformation. The

    zwitterionic intermediate should exist primarily in the gauche form because of

    Coulombic attraction between the oppositely charged termini. The diradica1 intermediate

    should exist in the trans form, as the attractive forces between the ends are not as

    strong.

    ~D A

    gaucho trans

    .,- • -,- ., +.-

    The Bond Forming Initiation Theory

    The Bond Forming Initiation TheoI}' extends the Huisgen hypothesis and proposes

    that these same tetramethylenes are the true initiators for the observed spontaneous

    polymerizations that often accompany [2 + 2] cycloadditions, and that this concept is

    valid for both ionic and radical polymerizations.

    If the tetramethylene is diradical in nature, it can initiate radical copolymerization.

    If it is zwitterionic in nature, it can initiate cationic homopolymerization of the electron-

  • rich olefin and/or anionic homopolymerization of the electron-poor olefin. It can also

    simply close to give the [2 + 2] cycloadduct (see Scheme 1).

    18

    Polymerization is a very sensitive technique for detecting diradical or zwitterionic

    intermediates.1 It offers two main advantages over conventional trapping techniques: 1.

    A high conversion of reactant is obtained from a minute amount of reactive intermediate

    in the form of polymer, which is easily isolated and identified. 2. The nature of the

    intermediate can be determined from the structure of the polymer. Isolation of

    copolymer is indicative of a diradical intermediate, while homopolymer is indicative of a

    zwinerionic intermediate (see Scheme 1).

    Note that only homopolymer can be obtained from a zwitterionic intermediate.

    Copolymerization cannot occur here since that would require attack by a cationic center

    on an already electron-poor moiety unable to stabilize a cation, or an anionic center on

    an already electron-rich moiety unable to stabilizp. an anion. A diradical center can

    propagate with either olefin, since the radical can be stabilized by either a donor or an

    acceptor group. The propagation is governed by the polarity of the radicals and

    monomers and an alternating tendency should result.

    An Organic Chemist's Periodic Table

    In a survey of reactions of donor and acceptor olefins, Hall developed the Organic

    Chemist's Periodic Table, shown in Figure 1.1 The electron-poor olefms are listed

    across in order of increasing electrophilicity. The electron-rich olefins are listed down

    in order of increasing nucleophilicity. Reactivity increases as one goes down and to the

    right. The table shows areas of mechanistic change from diradicals to zwitterions and

    from zwitterions to ion radical pairs. Weak donors and acceptors favor a diradical

    intermediate. Strong donors and acceptors favor a zwitterionic intermediate and very

    strong donors and acceptors favor radical ion pair intermediates.

  • 19

    Incr.ling eccaplor ability _

    ::-"c,,1!s ,-

  • 20

    Keeping in mind Huisgeu's idea of the nature of the hybrid tetramethylene being

    dependent on the tenninal substituents and the trends shown in the Organic Chemist's

    Periodic Table, Hall chose N-vinylcarbazole and p-methoxystyrene as donor olefins

    because of their ready polymerizability. He chose tetra- and trisubstituted electron-poor

    olefms with various combinations of ester and cyano groups for various reasons. First,

    they readily will stabilize a reactive center (radical or anionic). Second, they are either

    solids or high boUing liquids at ambient temperature, unlike many reactive olefins which

    are gases. This allows the reactions to be done under conditions that do not require

    high pressure glassware and extremely low temperatures. Third, these olefins do not

    homopolymerize, but they readily copolymerize. This makes isolation of two different

    polymers in cases of zwitterions (no easy task) WUlecessary, as only homopolymer of

    the electron-rich olefm will be obtained.

    Applications of the Bond Forming Initiation Theory

    Trapping of a Zwitterionic Intermediate Using Cationic Polymerization

    A zwitterionic intermediate was successfully trapped using the polymerization

    criterion in the reactions of N-vinylcarbazole and dimethyl 2,2-dicyanoethylene-l,l-

    dicarboxylate.' Cyclobutane, poly(N-vinylcarbazole) and a I-butene derivative were

    obtained as products (see Scheme 2). The formation of homopolymer in this reaction is

    indicative of a zwitterionic intermediate. The olefms initially form a brightly colored

    electron-donor-acceptor (EDA) complex which collapses to fonn the zwitterionic

    tetramethylene. At equimolar concentrations and dilute solution the latter closes to form

    the cyclobutane. This is the result of a cage effect, where the intermediate is

    surrounded by a cage of solvent molecules. The trans tetramethylene is unable to

    initiate polymerization since no monomer is nearby, so it either undergoes a proton shift

    to give the I-butene or rotates about the C;-c, bond back to the gauche tetramethylene

    where it simply closes to the cyclobutane. At higher concentrations or in excess N-

  • NCz

    =' -E CN-)=(

    E CN

    , • COOCH,

    NCI • M-C.rblll'r'

    EDA

    Complox

    _ r-!NCZ

    E--j\CN

    E CN

    1 NCz

    --

    NCr-(:! _ NC E

    1 'k'yJ:

    Nez

    HC. r='NCZ

    N~E H E

    Scheme 2 Trapping of a Zwitterionic Intermediate By Cationic Polymerization

    vinylcarbazole, the trans tetramethylene initiates cationic homopolymerization of the N-

    vinylcarbazole.

    21

    Note that the cyclobutane formation is reversible. If this cyclobutane is put in an

    excess of N-vinylcarbazole, poly(N-vinylcarbazole) is obtained. No color is observed in

    this reaction, indicating that the BDA complex does not participate in the initiation and

    that the zwitterion is the true initiating species.

    Only the cyclobutane in which the two cyano groups are a to the N-carbazolyl

    group is formed in this reaction. Cyano groups stabilize anions more effectively than

    ester groups. Hence, maximum stabilization of the zwitterionic intermediate is obtained.

    This is yet more evidence supporting the zwitterionic intermediate. This zwitterion can

    also be trapped using methanol to give the adduct shown below.

  • 22

    Trapping of a Zwitterionic Intermediate Using Anionic Polymerization

    Another zwitterionic tetramethylene was trapped using an e1ectton-poor olefin

    capable of anionic polymerizadon (see Scheme 3).' Isobutyl vinyl ether and methyl a-

    cyanoacrylate reacted at room temperature in equimolar amounts to give a dihydropyran

    and poly(methyl a-cyanoacrylate). The two olefins react to fonn a zwitterionic

    intennediate which can initiate anionic homopolymerization of the methyl a-

    cyanoacrylate or close through the ester carbonyl to give the [2 + 4] cycloadduct. This

    cycloadduct is not stable. It isomerizes to methyl 2-cyano-S·isobutoxypent-4·enoate. In

    dilute solution, only a trace of homopolymer is fonned, the dihydropyran being the

    major product. The zwinerion was trapped using methanol. The results here are similar

    to those obtained in the reactions of N-vinylcarbazole and dimethyl 2,2-dicyanoethylene-

    I,l-dicarboxylate. The presence of poly(methyl a-cyanoacrylate), the cage effect, and

    the methanol adduct are all evidence for a zwitterionic intennediate.

    Trapping of Diradical Intermediates

    Diradical intermediates were also successfully trapped using Hall's polymerizadon

    criterion. p-Methoxystyrene and styrene both reacted with methyl a-cyanoacrylate to

    give 1:1 alternating copolymer. A double Diels-Aldel' (Wagner-lauregg) adduct was also

    formed as shown in Scheme 4.9

    The system of p-methoxystyrene and dimethyl cyanofumarate was studied

    extensively.'o These two olefms reacted to give 1:1 alternating copolymer at high

    concentrations regardless of the feed ratio. Copolymer is indicadve of a diradical

    intermediate. At lower concentrations dihydropyran formadon began to compete with

    the copolymerizadon. Addition of the tetramethyl piperidine N-oxyl radical resulted in

    the adduct shown in Scheme S. A proposed mechanism for this adduct fonnadon is

    shown in Scheme 6.

  • 23

    ~ C+'" ~o'" OIBu Dc NC + CN =< CN COOCH 3 H3 COOC COOCH 3 1

    !J!

    OIBu

    ~OCH' 0rf. H3 00C eN CN

    Scheme 3 Trapping of a Zwitterionic Intermediate by Anionic Polymerization

    ~

    ~: ~.' Ar

    D

    E

    =< CN CN NC

    \ !J1 NC

    E f;yr{. E ArE eN

    Scheme 4 Trapping of a Diradical Intermediate I

  • ow. y

    Scheme S

    E

    /~ lo oddltlon eye

    ~~CH3

    COOC¥OCH 3 H3 CN

    rodlco

    Diradical Intennediate IT Trapping of a

    Iy morlzotlon I copo E

  • 25

    The kinetics of this system were studied at high concentrations so that dibydropyran

    formation did not play any role. The rate of polymerization was found to be second

    order in monomer concentration (first order in donor and first order in acceptor since

    the polymerization is strictly alternating). The derivation also took into account the

    propagation by the BOA complex. This term, which was third order in monomer, was

    found to be zero. This is evidence that the BOA complex plays no role in the

    propagation step. The kinetic data as well as the trapping experiment support a

    diradical intermediate.

    s)'" ~. :2t. - Ct ~ . :. E R CN E CN

    OCH.

    :& - ~:' q :' Scheme 6 Mechanism of Oiradical Intermediate Trapping with TEMPO

  • 26

    Borderline Tetrnmethylenes

    The Organic Chemist's Periodic Table shows an area of mechanistic change. The

    donor olefin may predict a zwitterionic intermediate, while the acceptor olefin may

    suggest a diradical intermediate and vice versa. Hall and Padias call these borderline

    cases "schizophrenic tetramethylenes."ll The nature of the tetramethylene can be altered

    by varying the solvent polarity. p-Methoxystyrene and methyl 2,2-dicyanoacrylate react

    to give copolymer as the major product in nonpolar solvents like carbon tetrachloride

    and dichloroethane. Dipolar aprotic solvents (DMSO, DMAc) favor a double Diels-

    Alder adduct and polar protie solvents (methanol, hexafluoroisopropanol) favor the

    cyclobutane adduct (see Scheme 7).

    Scheme 7

    ji'" - +

    $0 h,,,,

    OUSO, OU&<

    r=

  • 27

    In polar solvents, the diradical tetramethylene exhibits more polar character due to

    solvation effects. In other words, it becomes more zwitterion-like. A more polar

    intennediate will exhibit increased Coulombic attraction between the tennini favoring the

    gauche tetramethylene and hence the cyclobutane adduct. In nonpolar solvents, the

    diradical intennediate is free to rotate about the ~-c, bond, favoring the trans

    tetramethylene which initiates copolymerization. Polar aprotic solvents stabilize the 8+

    charged end of the tetramethylene, leaving the 8- end free to attack the phenyl ring,

    thus favoring the double Diels-Alder adduct The tetramethylene intennediate in this

    example is not a zwitterion as it does not react with methanol, but is believed to be a

    polar diradical.

    Reactions with ~-Leaving Group Olefins

    If an electron-rich olefin is reacted with an electron-poor olefin containing a p-

    leaving group, a tetramethylene with zwitterionic character will be fonned. This

    tetramethylene can close to give the cyclobutane or the ~-leaving group can be

    eliminated to give the cationic species shown in Scheme 8. This species with a less

    nucleophilic counterion can initiate cationic homopolymerization of the electron-rich

    olefm more effectively than the zwitterion itself.

    p-Methoxystyrene was reacted with l,2,2-tricyanovinyl trifluoromethane sulfonate,

    2,2-dicyanovinyl chloride, 2,2-dicarbomethoxyvinyl chloride,I2 2,2-dicyanovinyl iodide,

    and 2,2-dicyanovinyl tosylate,u All successfully initiated cationic homopolymerization

    of p-methoxystyrene. Both 2,2-dicyanovinyl tosylate and 2,2-dicyanovinyl iodide still

    initiated polymerization in the presence of a radical inhibitor, bis(3-t-buty14-hydroxy-S-

    methylphenyl) sulfide, illustrating that the reaction did not proceed by way of a radical

    mechanism. Polymerization also proceeded in the presence of a proton trap, 2.4,6-tri-t-

    butylpyridine. This indicates that no proton transfer is involved in the reaction. A

  • D ~

    A

    r=< X A

    X. 01f, OT •• I. CI

    'AA X A I

    dA X A

    D /

    ~~+ A :;1

    I D A, r-{. )=' . x-

    Scheme 8 BFI Theory Using P-Leaving Group Olefins

    28

    bimodal molecular weight distribution was obtained from this polymerization - polymer

    resulting from free ion and ion pair propagation. Ion pair propagation is slower and the

    molecular weight of the polymer is lower.

    Olefms containing the triflate leaving group were extremely reactive. The reactions

    could not be controlled. Tricyanovinyl chloride and dicyanovinyl chloride were less

    reactive and hence less efficient at initiating polymerization. This is to be expected as

    triflate is a much better leaving group than chloride. Polymerization was accompanied

    by butadiene formation. As the initiator concentration increased. so did the amount of

    butadiene formed in the reaction. Polymer yield also increased. but molecular weight

    decreased due to termination by chloride ion. Dicarbomethoxyvinyl chloride reacted

    very slowly. but it also initiated homopolymerization of p-methoxystyrene.

    Reactivity correlated well with leaving group ability. The triflates were the most

    reactive. while the chlorides were the least reactive. Tosylate and iodide fell in

  • 29

    between. Polymer yield and molecular weight also correlated well with leaving group

    ability. More polymer of higher molecular weight was obtained with the olefins with

    the best leaving. groups (tritlate, tosylate). Less reactive olefins are more subject to

    premature termination by the gegenion. In the case of the chloride initiators, butadiene

    derivatives are also formed.

    These results illustrate how changing the substituent on the reactants affects the

    character of the tetramethylene. p-Methoxystyrene has been shown to copolymerize with

    electron-poor olefins, indicating initiation by a diradical intermediate. By introducing a

    leaving group, the nature of the tetramethylene is changed to zwitterionic. Huisgen's

    hybridized tetramethylene and Hall's Bond Forming Initiation Theory are further

    supported.

    Application of the Bond Forming Initiation Theory to [4 + 21 Systems

    Hall has successfully trapped both zwitterionic and diradical intermediates using his

    polymerization criterion. The Bond Forming Initiation Theory appears to explain the

    spontaneous polymerizations observed in reactions of electron-rich olefins and electron-

    poor olefms. Perhaps this theory could also be applied to reactions of electron-rich

    dienes with electron-poor olefins.

    According to the Woodward Hoffmann Rules, a [4 + 2] cycloaddition reaction is an

    allowed, concerted, electrocyclic reaction. This reaction can, however, be a stepwise

    reaction. Many reviews have been written on the mechanism of cycloaddition reactions,

    and many contain examples of [4 + 2] cycloadditions that are not concerted.14,15.16,17 The

    arguments of Huisgen and Firestone" seem to typify the state of affairs as to the

    reaction mechanism.

    Bartlett studied the reactions of chlorotluoroethylenes with l,3-butadiene.'~ The

    results are shown in Table 1. The [2 + 2] cycloadduct is the major product in the

    reaction with l,l-dichloro-2,2-ditluoroethylene, while the [4 + 2] cycloadduct is the

  • 30

    major product with vinylidine fluoride. The chlorine is better able to stabilize the

    radical than the fluorine. A stepwise intermediate is favored over the concerted

    mechanism, hence the higher yield of cyclobutane. In the cases of trifluoroethylene and

    vinylidene fluoride, the concerted mechanism may play a larger role.

    Little proposed a diradica1 intermediate in the reactions of a-acetoxyacrylonitrile

    and butadiene.:W The cyclohexene/cyclobutane ratio remained roughly the same

    regardless of the solvent or the temperature. Since it is unlikely that competing

    mechanisms (concerted [4 + 2] cycloaddition and stepwise [2 + 2] cycloaddition) depend

    on the same reaction conditions, in this case differing solvent polarity and

    temperature.the two products may have originated from a common intermediate, namely

    a diradical intermediate.

    Further studies by Banlen and Schueller using 2,4-hexadiene and ct-

    acetoxyacrylonitrile seemed to refute these conclusions.21 Only two isomeric forms of

    the [4 + 2] adduct were obtained. in which the stereochemistry of the methyl groups

    was retained. In reactions of 2,4-hexadiene with ethylene, the stereospecific [4 + 2]

    cycloadduct was the only product. Both [4 + 2] and [2 + 2] adducts were obtained in

    the reactions of 1,3-butadiene and ethylene. the latter in only 0.02% yield.22 These

    results lead to the conclusion that the diradical addition becomes twenty times slower

    and the concerted reaction becomes ten times faster on substitution of the diene. This

    supports the theory of two competing mechanisms: the thermally allowed concerted [4 +

    2] and the stepwise [2 + 2]. Substitution hinders initial attack to form the diradical.

    Stewart proposed a zwinerionic intermediate in the reactions of piperylene and

    tetracyanoethylene in which he obtained both vinylcyclobutane and cyclohexene

    adducts.23 Eisch and Husk also proposed a zwinerionic intermediate in their

    cycloaddition reaction of 1,1-diphenyl-l,3-butadiene and tetracyanoethylene.:U

  • Table 1 Banlen's Results

    Olefin

    F C I

    >=< F C I

    ==

  • 32

    Van Mele and Huybrechts obtained some rather convincing evidence for a diradical

    intennediate in the cycloadditions and the retro-Diels-Alder reactions of cyclohexadiene

    with the cis and trans 2-butenes shown in Scheme 9.25 Stereospecificity of the starting

    diene was lost in each case and both isomeric bicyclo[2.2.2]octenes were fonned.

    SimUar results were obtained in the thennolysis reactions of disubstiwted

    bicyclo[2.2.2]octenes.26 Endo-cis-S-cyano-6-methyl-bicyclo[2.2.2]octene and exo-trans-

    S-cyano-6-methyl-bicyclo[2.2.2]octene showed loss of stereochemistry after undergoing

    retro-Diels-Alder reactions. According to the principle of microscopic reversibility, the

    forward reaction should proceed in the same manner as the reverse reaction.

    Nishida, et al. studied the reactions of l-cyclopropyl substiwted 1,3-butadienes with

    tetracyanoethylene (see Scheme 10).27 When the E-l-cyclopropyl-l,3-butadiene was

    used, both vinylcyclobutane and cyclohexene adducts were formed. The product ratio

    o o

    Scheme 9

    +

    !t.". -- ~'" I tH. tH. ).~".~ ~ ~~J ~:J

    " CH, Reactions of Cyclohexadiene with Cis- and Trans- 2-butenes

    was dependent on solvent polarity; increasing solvent polanty ravored the cyclobutane.

    This is characteristic in reactions involving zwitterions. The vinylcyclobutane

    isomerized to the cyclohexene. A zwitterionic intennediate was proposed to explain

    these results. The vinylcyclobutane was fonned as the kinetic product and the

    cyclohexene as the thennodynamic product. According to Houk, the best known

    examples of stepwise Diels-Alder reactions involve either nonpolar partners or highly

  • 33

    unsymmetrical polar partners.28 Nonpolar cycloadditions have little preference for the

    concened mechanism over the stepwise mechanism, while polar Diels-Alder additions do

    show a high preference for the concened mechanism. With highly polar diene-

    dieneopile combinations, in which at least one addend is highly unsymmetrical, stepwise

    reactions involving zwitterionic intermediates take place. This seems to be a relatively

    correct assessment in the case of polar diene-dienophile combinations.

    However, Bartlett has come to the conclusion that alkyl substituted dienes favor the

    concerted mechanism (alkyl substituted and still relatively nonpolar dienes).29 In

    thereaction of butadiene with ethylene, the ratio of cyclohexene to vinyl cyclobutane

    was found to be 99.98 to 0.02. Reactions of any substituted diene would yield a

    virtually undetectable amount, if any cyclobutane. This is supported by the fact that no

    H

    ~ + NC CN

    )=( NC CN ~

    H+ CN - CN

    I CN CN

    dtf;H CN

    eN CN

    eN

    \

    ~ CN I eN CN eN

    Scheme 10 Reactions of l-Cyclopropyl-l,3-Butadiene and TCNE

  • '.

    vinylcyclobutane is obtained in the reactions of 2,4-hexadiene with ethylene.

    Despite the examples of these apparently nonconcerted reactions. the general

    consensus is that [4 + 2] cycloadditions are concerted. and that there are a few

    exceptions to the rule. These are governed by the following criteria:'o

    1. Substituents at the 1.6 tennini that stabilize the biradical or the zwitterion can

    substantially lower the energy level of the intennediate.

    34

    2. The concened pathway requires an s-cis 1.3-diene. The smaller the s-cis content in

    the confonnational equilibrium. the higher the chance of the two-step reaction.

    3. The demands on the order of the transition state are less stringent for the fonnation

    of the intennediate than for the concerted pathway.

    4. The substitution of the diene is such that the concerted transition state is hindered.

    The diene must be in the s-cis fonn. which is present only in minute

    concentrations. for a concerted reaction to occur. The s-trans fonn of the diene is the

    0 0 0

    ~ (A j - f_ -1

    ~A • A

    0

    ~D 0 erA + erA

    -A

    Scheme 11 Reactions of Cis and Trans Dienes

  • 3S

    predominant fonn, and can also be expected to react. However, due to the geometry

    constraints, it will react in a stepwise manner. Addition of the olefin to the tenninus of

    the diene can lead to either diradical or zwitterionic intennediates as shown in Scheme

    11. The identity of the intennediate will be dependent on the substituents present

    Rotation about the olefin C-C bond will result in loss of stereospecificity. Rotation

    about the allylic bond is unlikely since this would disrupt the allylic stabilization.

    A stereospecific reaction is expected when the HOMO of the diene and the LUMO

    of the dienophile or vice versa is small and the steric conditions allow a cisoid

    confonnation of the diene and thus a smooth reaction in the 1,4 position. Large

    HOMO-LUMO separation slows down the rate of the synchronous reaction to such an

    extent that perhaps a slower two-step reaction becomes possible.

    Cis I-substituted dienes will favor the transoid configuration of the diene by

    sterically hindering the cisoid confonnation. A nonconcerted mechanism should be

    favored. Substituents capable of stabilizing intennediates will also favor the two step

    mechanism.

    Proposed Research

    The Bond Forming Initiation Theory, derived for [2 + 2] systems, could also be

    applied to [4 + 2] systems. When a transoid electron-rich diene reacts with an electron-

    poor olefm, an intennediate may be fonned as shown in Scheme 12. The resonance

    fonn where the active center is adjacent to the donor group should be favored because

    of extra stabilization. As Huisgen suggested for the tetramethylene intennediate, the

    hexamethylene intennediate may also exist as a hybrid between the zwitterionic and

    diradical extremes. The nature of the intennediate should also be substituent-dependent

    in the same way. Strong donors and acceptors should favor the zwitterionic

    intennediate and weak donors and acceptors should favor a diradical intennediate.

  • 36

    This hexamethylene intermediate in the extended conformation can initiate

    polymerization. If the hexamethylene intermediate is diradical in nature, it could initiate

    copolymerization. If it is zwitterionic in nature, it could initiate either cationic homo-

    polymerization of the electron-rich diene and/or anionic homopolymerization of the

    electron-poor olefin. TIle intermediate in the cisoid conformation can simply close to

    form the cyclohexene. This theory does not rule out the possibility that the cycloadduct

    can also be formed via a concerted mechanism. It only suggests that if an intermediate

    is present. polymerization may be a useful means of identifying it

    A cisoid diene should react via a concerted mechanism to give only the

    cyclohexene. The transoid diene should react via an intermediate to give either the

    cyclobutane or the cyclohexene. More cyclobutane than cyclohexene is expected from a

    diradical intermediate because rotation about the bond that is part of the allylic system

    is unlikely. At higher temperatures. more cisoid diene is present and the amount of [4

    + 2] cycloadduct should increase.

    D

    j + •

    1 I; &A c(

    A

    D

    Scheme 12 The BFI Theory Applied to [4+2] Systems

    ~ D A

    ~ D

    k'y1 A

  • Introduction

    CHAPI'ER 2

    Reactions of l-AryI-l,3-Butadienes With Trisubsdtuted Electrophilic Olefins

    37

    In order to apply the Bond Fonning Initiation Theory discussed in the previous

    chapter to [4 + 2] systems, reactants must be chosen so that an intermediate will be

    favored. According to Houk, stepwise Diels-Alder reactions are favored when the

    reactants are highly unsymmetrical and polar.n It seems logical that unsymmetrical,

    polar reactants with donor and acceptor substituents on the diene and the dienophile

    respectively, will favor an intermediate. The choice of substituents will depend on the

    type of intermediate desired for study.

    According to the Organic Chemist's Periodic Table (Figure I), phenyl substituents

    on the diene should favor a diradical intermediate. The trisubstituted electrophilic

    olefms methyl 3,3-dicyanoacrylate, dimethyl cyanofumarate, and trimethyl ethylenetricar-

    boxylate were used in the olefin-olefin reactions and excellent results were obtained.

    These olefins could also be used in the [4 + 2] cases.

    The isolation of polymer on reactions of the dienes with the olefms would indicate

    the presence of a reactive intermediate. If a copolymer were obtained, the reaction

    would be proceeding by way of a diradical intermediate. A homopolymer would

    indicate that the reaction is proceeding via an ionic intermediate.

    Huisgen's hybrid tetramethylene theory can also be applied to hexamethylene

    intermediates. By introducing a leaving group at the (i-position of the olefin, the

    character of the hybrid is shifted towards zwitterionic. As the leaving group is expelled,

    the character of the intermediate should become more and more zwinerionic. The result

  • 38

    is a cationic intennediate that could initiate homopolymerization of the diene as shown

    in Scheme 13.

    The dienes selected for this study were I-phenyl-l,3-butadiene 1 and p-anisyl-l,3-

    butadiene 2. The trisubstituted elecuophilic olefins methyl 3,3-dicyanoacrylate 3,

    dimethyl cyanofumarate 4, and trimethyl ethylenetricarboxylate 5 were chosen to study

    2:'(~ J -~ k/~

    C-l/ x

    o

    A

    Scheme 13 General Mechanism of Reactions of Dienes with p-Leaving Group Olefms

    the diradical intennediate. The olefms with p-leaving groups selected to study the

    zwitterionic intennediate were 2,2-dicyanovinyl chloride 7, 2,2-dicyanovinyl iodide 8,

    2,2-dicyanovinyl tosylate 9, 2-cyano-2-carbomethoxyvinyl chloride 10, and 2-cyano-2-

    carbomethoxyvinyl iodide 11.

    Background

    I-Phenyl-l,3-butadiene has been polymerized by anionic and cationic means.32..""",33,36

    No mention of radical polymerization of I-phenyl-l,3-butadiene or p-anisyl-l,3-

    butadiene has been found. Appropriate control experiments would have to be perfonned

    in order to ascenain the polymerizabUity of these monomers. The trisubstituted olefins

  • used in this study do not homopolymerize. Thus only a homopolymer of the

    arylbutadiene would be obtained if the reactive intermediate were a zwitterion.

    Results

    Control Reactions

    39

    Control experiments were done to determine if the chosen systems would

    polymerize. The results are shown in Table 2. The reactants were deliberately initiated

    using both cationic and radical initiators. p-Toluenesulfonic acid and boron trifluoride

    etherate successfully initiated cationic homopolymerizations of I-phenyl-l,3-butadiene to

    give poly(phenylbutadiene). No radically initiated homopolymerization of the

    arylbutadienes was successful. Radical copolymerization reactions of the arylbutadienes

    and the trisubstituted olefins were also unsuccessful. The reaction of p-anisyl-l,3-

    butadiene with trimethyl ethylenetricarboxylate is a much slower reaction than reactions

    with methyl 3,3-dicyanoacrylate and dimethyl cyanofumarate. Initiation with AIBN at

    85°C gave only the [4 + 2] cycloadduct. Due to the rapid reaction rates of the

    spontaneous reactions between p-anisyl-l,3-butadiene and both methyl 2,2-

    dicyanoacrylate and dimethyl cyanofumarate, no deliberate AIBN initiated

    polymerizations were successful.

    Triethyl borane (in THF) and oxygen were used as the initiating system at -50°C.

    p-Anisyl-l,3-butadiene was dissolved in toluene and the triethyl borane, oxygen, and

    methyl dicyanoacrylate were added sequentially. A dark orange charge transfer complex

    formed immediately on addition of the olefin. After 16h at -50°C, a precipitate had

    formed. This precipitate proved to be the [4 + 2] cycloadduct. No polymer was

    obtained. Because of limited solubility of the diene in toluene at -SO°C, the reaction

    was repeated at -3SoC in dichloroethane. The results were the same - no copolymer

    was obtained. The same procedure was repeated using dimethyl cyanofumarate as the

  • 40

    Table 2 Control Reactions

    Monomers Initiator Solvent Time Temp. Polymer (h) (OC) Yield(%)

    1 p-TsOH DCE 0.03 25 81

    1 BF3 "Et 2O CH 3N02 0.02 25 79

    1 AIBN Bulk 48 70 0

    2 + 3 Et3B/02 CH 3 16 -50 0

    2 + 3 Et3B/02 DCE 16 -35 0

    2 + 4 Et3B/02 DCE 16 -35 0

    2 + 5 AIBN Bulk 16 85 0

    2 + 29 AIBN Bulk 72 80 0

    2 + 5 AIBN 29 41 80 0

    DCE = D~chloromethane

  • electrophilic olefin. The same results were obtained; the [4 + 2] cycloadduct was the

    only product formed. No copolymer was formed.

    41

    AIBN initiated polymerization of acrylonitrile in the presence of the p-anisyl-l,3-

    butadiene/trimethyl ethylenebicarboxylate system was unsuccessful. No polymer was

    obtained. AIBN initiated polymerization of acrylonitrile was inhibited by the presence

    of p-anisyl-l,3-butadiene. The diene appears to act as a radical polymerization inhibitor.

    According to Mulzer, I-phenyl-l,3-butadiene does not spontaneously

    homopolymerize.37 p-Anisyl-I,3-butadiene was heated at 175°C for IS hours. No

    decomposition, Oligomerization, or other reactions occurred. After 96 hours, there

    appeared to be some dimerization. The diene does not spontaneously homopolymerize

    with itself.

    Reactions of l-Aryl-l,3-Butadienes with Trisubstituted Electrophilic Olefms

    p-Anisyl-l,3-butadiene reacted in less than a minute with maleic anhydride 6 to

    give an essentially quantitative yield of the [4 + 2] cycloadduct 18 (see Table 3).

    Reactions of the electrophilic olefms 3-5 with the arylbutadienes 1-2 in dichloroethane

    and nitromethane gave exclusively [4 + 2] cycloadducts.

    Two stereoisomers were formed from endo and exo additions. Isomer ratios were

    determined by O. C. and are reported in Table 4. No [2 + 2] cycloadduct was

    observed in any case, nor was any polymer observed.

    Experimental Observations

    There was some experimental difficulty in doing the bulk reactions. All of the

    reactants are solid at room temperature. When the diene is melted and then

    supercooled, it stays in liquid form. The reactants can be mixed by pouring the diene

    into the dienophile. The resulting reaction mixture is very viscous and cannot be

  • 42

    Table 3. Reactions of 1-Aryl-1,3-Butadienes with Trisubstituted Olefins

    Reactants Solvent Cone. Time Temp. Adduct Adduct (Ml (hI (OCI 'Yield

    1 + 3 Bulk 1 65 12 87

    DCE 5 4 25 12 71

    DCE 1 4 25 12 73

    1 + 4 Bulk 1 25 13 84

    DCE 5 5 25 13 57

    DCE 1 16 25 13 61

    1 + 5 Bulk 96 75 14 71

    DCE 5 144 70 14 86

    2 + 3 Bulk 20 25 15 85

    DCE 5 2 0 15 76

    DCE 1 20 25 15 88

    CH 3NO, 5 16 25 15 91

    2 + 4 Bulk 20 25 16 67

    DCE 5 2 0 16 68

    DCE 1 20 25 16 77

    CH 3NO, 5 16 25 16 76

    2 + 5 Bulk 20 25 17 0

    DCE 5 48 150 17 21

    CH 3NO, 5 16 80 17 83

    2 + 6 Bulk 0.03 25 18 98

    DCE - Dichloroethane

  • 43

    Table 4. Reactions of 1-Aryl-1,3-Butadienes with P-Leaving Group Olefins

    Reactants Solvent Cone. Adduct Adduct Homopolymer (M) 'BYield 18 Yield

    1 + 7 Bulk 19 54 6

    DCE 5 19 43 2

    DCE 1 19 47 0

    CH 3N0 2 5 19 39 4

    1 + 8 DCE 5 20 46 16

    CH 3N0 2 5 20 34 25

    1 + 9 DCE 1.5 0 37

    CH 3N0 2 2 0 54

    1 + 10 DCE 5 21 77 1

    CH 3N0 2 5 21 80 1

    1 + 11 DCE 5 22 79 1

    CH 3N0 2 5 22 87 2

    2 + 7 DCE 5 23 80 5

    2 + 9 DCE 5 0 95

    DCE 5* 0 100

    2 + 10 DCE 5 24 88 8

    2 + 11 DCE 5 25 74 13

    DCE - Dichloroethane * Molar Ratio of p-anisyl-1,3-butadiene/2,2-dicyanovinyl tosylate

    10: 1.

  • 44

    stirred, hence the lower yields for the bulk reactions in some cases. In the case of p-

    anisyl-l,3-butadiene and ttimethyl ethylenetricruboxylate, the reactants solidified on

    mixing and no reaction occurred. Starting material, which was identified in both NMR

    spectra and O. C. analyses, remained in all of these reactions, but was Dot isolated.

    The reactions of dienes 1 - 2 with methyl 3,3-dicyanoacrylate and dimethyl

    cyanofumarate are very rapid and exothermic. The reactants were mixed at O°C in

    efforts to control the reaction rate. After the initial heat had dissipated, the reaction

    mixture was stirred at room temperature. The reactions of dienes 1 - 2 with trimethyl

    ethylenetricarboxylate are very slow. These reactions were mixed at room temperature

    and stirred at higher temperatures.

    Stereochemistry of the Cycloadducts

    Reactions of the two arylbutadienes with the trisubstituted olefins gave two isomers

    resulting from endo and exo addition. The predominant isomer was thoroughly

    investigated by 'H NMR. and has the structure depicted in Figure 2. This structure was

    determined by proton homodecoupling experiments. Coupling constants are discussed

    only for 4,4,5-tricarbomethoxy-3-phenyl-cyclohexene 14. The coupling constants of

    cycloadducts 12-17 were also determined and the same analysis was again applied to

    determine their stereochemistries.

    Irradiation of ~ and ~ shows that H, is coupled only to these two hydrogens.

    The doublet from H, is simplified to a singlet on irradiation of Hl , indicating that the

    methyne is adjacent to the quaternary carbon and to the vinyl proton, Ha. The

    stereochemistry at C, and Cs was determined by calculating coupling constants, ISla and

    JS,60' ISla is approximately 11Hz and Is,60 is approximately 7.5Hz. This is indicative of

    axial-axial and equatorial-axial coupling, respectively. Thus, Hs is in the axial position.

    Homoallylic coupling constants 131a and 13,60 were calculated to be 305Hz and 105Hz

  • 45

    respectively. Comparison with Barfield's calculated coupling constants indicates that H,

    is axial."

    Reactions with Olefins Containing p-Leaving Groups

    Homopolymer of the arylbutadiene was obtained in all reactions of arylbutadienes 1

    - 2 with Uisubstituted olefins having leaving groups in the P position (see Table 5).

    The [4 + 2] cycloadduct was also obtained in all cases as a mixture of two isomers,

    except in the reactions using 2,2-

  • 46

    A

    3 2 H H A

    :5 H

    Figure 2 Major Isomer

    Table S. Isomer Ratios of [4 + 2] Cycloadducts in Reactions of p-Anisyl-1,3-Butadiene

    Olefin Ratio Olefin Ratio

    3 2:1 7 3:1

    4 1:1 10 1:1

    5 1:1 11 3:1

  • 47

    backbone stereochemistry is cis-lA, or ttans-1,4. IR spectra show no terminal vinyl

    groups, so it is unlikely that 1,2 or 3,4 addition oCCUITed. The yield increased with

    increasing leaving group ability, as did molecular weight. The molecular weights of the

    polymers are low (see Table 6), which may be atUibuted to the fact that a 1:1 molar

    ratio of diene to olefin was used. Molecular weight increases with increasing leaving

    group ability. When 10 mole% of dicyanovinyl tosylate was used, a quantitative yield

    of poly(p-anisyl-l,3-butadiene) was obtained.

    Cycloadduct

    [4 + 2] Cycloadduct was obtained in all cases, except in the reactions with 2,2-

    dicyanovinyl tosylate. All reactions were done at high concentration or in bulk to favor

    polymerization. Two isomers of the cycloadduct were observed (see Table 5). No

    extensive NMR studies of these [4 + 2] cycloadducts were done. However, because of

    similar splitting patterns, these isomers are believed to have the same stereochemistry as

    the cycloadducts found in the reactions with the trisubstituted olefins.

    Side products of a highly conjugated nature were also observed, as was the

    methanol adduct of the olefin starting material, formed during the work-up. The latter

    was isolated and obtained in quantitative yield. The former was present in very small

    quantities and was not isolated. This side product is probably the hexatriene shown in

    Scheme 14. Analogous reactions took place in reactions of p-methoxystyrene and these

    ~-leaving group olefins to form butadienes."

    )' f..-r-" '17 el Ne

    Scheme 14. Formation of Hexatriene

    ~ J' .y eN

  • 48

    Table 6 AryI-l,3-Butadienes with jl-Leaving Group Olefins

    Reactants Solvent Cone. Adduct Adduct Homopolymer (M) \Yield , Yield

    1 + 7 Bul'k 19 54 6

    DCE 5 19 43 2

    DCE 1 19 47 0

    CH 3NOz 5 19 39 4

    1 + 8 DCE 5 20 46 16

    CH 3NOz 5 20 34 25

    1 + 9 DCE 1.5 0 37

    CH 3NOz 2 0 54

    1 + 10 DCE 5 21 77 1

    CH 3NOz 5 21 80 1

    1 + 11 DCE 5 22 79 1

    CH 3NOz 5 22 87 2

    2 + 7 DCE 5 23 80 5

    2 + 9 DCE 5 0 95

    DCE 5* 0 100

    2 + 10 DCE 5 24 8e S

    2 + 11 DCE 5 25 74 13

    DCE - Dichloroethane • Molar Ratio of p-anisyl-1,3-butadiene/2,2-dicyanovinyl tosy1ate

    10:1.

  • 49

    Discussion

    Deliberately attempted copolymerization reactions of aryl butadienes and olefins 3 -

    5 were unsuccessful. Because of the rapid spontaneous reaction rates of the aryl

    butadiene with methyl 2,2-dicyanoacrylate and dimethyl cyanofumarate, an initiator

    capable of initiating polymerization at low temperatures was needed. Low temperature

    polymerization of lithium enolates and vinyl monomers had previously been done with

    a triethyl borane/oxygen system.40•41 This proved unsuccessful in the aryl butadiene

    reactions. Only [4+2) cycloadduct was obtained and no copolymer was observed.

    Because the reactions of the aryl butadienes with trimethyl ethylenetricarboxylate

    are much slower, AIBN was used as an initiator. Again, only [4 + 2) cycloadduct was

    obtained. AIBN-initiated copolymerization of p-anisylbutadiene and trimethyl

    ethylenetricarboxylate was also tried in the presence of acrylonitrile, a third radically

    polymerizable monomer. The diradical intermediate, if present, and the AIBN were

    expected to initiate polymerization of the acrylonitrile. Again, no polymer of any kind

    was obtained. AIBN initiated polymerization of acrylonitrile in the presence of p-

    anisyl-l,3-butadiene again failed - no polymer was obtained. An aryl radical formed in

    this reaction could be stabilized by the aryl group to such a degree that propagation is

    inhibited. The arylbutadiene acts as a radical trap. Thus, no copolymer can form and

    evidence for a diradical intermediate in these reactions must be found through different

    methods.

    The choice of l-aryl-l,3-butadienes as the electron-rich dienes proved to be poor.

    The reactions of the arylbutadienes with the trisubstituted olefms were carried out, even

    though it was unlikely that spontaneous polymerization would occur. Indeed, this was

    found to be the case. No copolymer was obtained, nor was any [2 + 2) cycloadducL

    The [4 + 2) cycloadducts were the only products.

  • SO

    The stereochemistry of the cycloadducts can provide insight into the mechanism of

    these reactions. If the reaction is concerted, two isomeric cycloadducts would be

    formed as shown in Scheme IS. If the olefin addition is endo, a cyclohexene where

    the substituents at C, and C~ are trans is obtained. If the olefin addition is exo, a

    cycloadduct where the substituents at C, and C~ are cis is obtained. In the reactions of

    dimethyl cyanofumarate, four differenl isomers would be expected if the reaction was

    stepwise - two isomers where the stereochemistry of the ester groups was retained and

    remained trans, and two where the C4-C~ bond undelWent rotation and the ester groups

    were cis. In the reactions of methyl 3,3-dicyanoacrylate and Uimethyl

    ethylenetricarboxylate only two isomeric cyclohexenes would be obtained regardless of

    the reaction mechanism.

    Only two isomeric cyclohexenes were obtained in all cases. These isomers were

    those where the stereochemistry of the ester groups in the reactions of the aryl

    butadienes with dimethyl cyanofumarate remained trans. There was no loss of

    stereochemistry and no evidence that the reactions proceed via a diradical hexamethylene

    intermediate.

    According to the Organic Chemist's Periodic Table, the methoxyphenyl group and a

    dicyano moiety are a strong donor and strong acceptor, respectively. Intermediates in

    the reactions of vinyl compounds having these groups as substituents behave as

    "schizophrenic" tetramethylenes. Changing reaction conditions, such as solvent polarity,

    can change the nature of the tetrarlethylene. Polar solvents favor a zwitterionic

    intermediate, while nonpolar solvents favor a diradical intermediate. p-Anisyl-l,3-

    butadiene and methyl dicyanoacrylate may behave in an analogous manner. By

    changing the solvent to more polar nitromethane, a zwitterionic intermediate may be

    favored. This intennediate could initiate cationic homopolymerization of the diene,

  • SI

    D D

    ~ !£ :CA A H X,A H A

    j I

    4 4 A H 1\1

    "' A A o;'Ay: o~v:

    H H H A A H

    Scheme 15. Concerted Mechanism

  • S2

    [

    D

    H

    CN CN.

    E E D

    H H H E

    Figure 3. Stepwise Mechanism

    which has already been shown possible.42 In this case using the more polar

    nitromethane did not induce the zwitterionic intermediate to initiate homopolymerization

    of the diene. Only the [4 + 2] cycloadduct was obtained. No homopolymer of the

    diene was obseJVed.

    According to Huisgen, the tetramethylene intermediate exists not as a diradical nor

    as a zwitterion, but a resonance hybrid of the two. Substituents will cause the

    tetramethylene to exhibit characteristics of one or the other. The same can be said of

    the hexamethylene intermediate. By introducing a leaving group at the 13 position of

    the olefm, the character of the hybrid is shifted towards zwitterionic character. As the

    leaving group is expelled, a cationic intermediate is formed which could initiate

    homopolymerization of the electron-rich aryl butadiene very efficiently. This is indeed

    the case.

  • Reactions of I-phenyl-l.3-butadiene and p-anisyl-l.3-butadiene with nisubstituted

    oleflIlS containing j3-leaving groups were run in dichloroethane. Homopolymer of the

    diene was obtained in all cases. Homopolymer is evidence that this reaction proceeds

    via an ionic intermediate.

    The cycloadducts were obtained in all cases with the exception of reactions of

    dicyanovinyl tosylate. As in the previous reactions. two isomers were formed. They

    have the same structure as those previously discussed. based on spectral data. In the

    cases of the 2-carbomethoxy-2-cyano vinyl olefins. no rotation about the olefin C-C

    bond was observed.

    53

    Reactions of the chloro and iodo substituted olefins gave both cycloadduct and

    homopolymer. the former more adduct and less polymer than the laner. Reactions of

    the tosyl substituted olefin gave only homopolymer. The tosyl group is the least

    nucleophilic and is less likely to interfere with the polymerization. Iodide is less

    nucleophilic than chloride and would be less likely to interfere with polymerization than

    chloride. The olefins having a tosylate group would be expected to initiate

    polymerization more efficiently than iodo substituted oleflIlS. which are expected to be

    more effective initiators than chloro substituted olefins. Molecular weights of the

    polymers also increased with decreasing nucleophilicity indicating that the tosylate was

    interfering less with the polymerization than were the iodide or chloride.

    Conclusion

    The cycloadducts formed in these reactions probably result from the concerted

    reaction of the cis diene. Since only two isomeric cycloadducts were obtained instead

    of four (in the cases where the substituents at C4 were different). there is no evidence

    that the cycloadduct formation was the result of ring closure of a diradical intermediate.

    Since no copolymer was formed. there is no evidence for a diradical intermediate. The

  • 54

    homopolymers obtained in the j3-leaving group olefin reactions could only have been the

    result of a zwitterionic intermediate. Thus, in the reactions of l-aryl-l,3-butadienes with

    (3-leaving group olefins, a zwitterionic intermediate was successfully trapped using the

    polymerization criterion.

  • 55

    CHAPTER 3

    Reactions of I-Methoxy-l.3-Butadiene

    Introduction

    The l-aryl-l.3-butadienes proved to be unsatisfactory for studying possible diradical

    intermediates. They appeared to act as radical polymerization inhibitors. I-Methoxy-

    1.3-butadiene should not be capable of stabilizing a radical to such an extent. and it

    may prove to be a suitable diene for this study.

    Background

    There is little mention of polymerization reactions of l-alkoxy-l,3-butadienes in the

    literature. l-Alkoxy-l,3-butadienes spontaneously polymerized in the presence of

    chloromethyl ethers.43 Flaig reports the spontaneous copolymerization of I-methoxy-

    1,3-butadiene and maleic anhydride.44 However, he was not interested in the copolymer

    and took steps to inhibit the polymerization with methylene blue. Even so, cycloadduct

    formation was still accompanied by polymerization, but to a lesser extent (80%

    cycloadduct and 14% copolymer with inhibitor compared to 35% cycloadduct and 45%

    copolymer without inhibitor).

    Stepek studied the reaction of l-methoxy-l,3-butadiene with maleic anhydride in

    great detail.45 He found that these two compounds spontaneously polymerized to give a

    1:1 alternating copolymer, where the I-methoxyl,3-butadiene copolymerized by 1,4

    addition. This polymer was found to be identical with that formed when dibenzoyl

    peroxidelp-bromo-N,N-dimethylaniline was used as the initiator. Since a copolymer was

    formed, and the reactivity ratios for these two monomers are essentially zero,

    characteristic of highly alternating copolymers, he postulated a radical mechanism.

  • The rate of polymerization was found to be second order. The rate of

    cycloaddition, measured in the presence of hydroquinone which inhibited the copoly-

    merization, was also found to be approximately second order.

    56

    Energies of activation for each of these processes were measured. Activation

    energies for the copolymerization reactions of I-methoxy-l,3-butadiene and maleic

    anhydride and l-ethoxy-l,3-butadiene and maleic anhydride were found to be 14.8

    kcal/mole and 16.3 kcal/mole respectively. The cycloaddition reactions were found to

    have activation energies of 12.1 kcal/mole and 11.4 kcal/mole respectively. These

    different activation energies for the copolymerization and the cycloaddition coupled with

    the fact that inhibitors had no effect on the fonnation of the cycloadduct led Stepek to

    conclude that there were two different mechanisms operating in this reaction.

    I-Methoxy-l,3-butadiene is readily polymerized by cationic initiators such as

    mineral acids, zinc chloride, boron trifluoride, and aluminum chloride.46.47 These

    reactions are very violent and are best done at low temperatures and in dilute solution.

    Butler and Chen studied the reactions of l-ethoxy-l,3-butadiene and acrylonitrile.48

    These reactants spontaneously copolymerized to yield highly alternating copolymers

    having cis-3,4, trans-3,4, cis-l,4 and trans-l,4 structures. They proposed a mechanism

    where the electron donor acceptor complex fonned ion radical pairs which then initiated

    the polymerization as shown in Scheme 16. Cycloaddition accompanied the copolymer

    fonnation.

    Since there is precedence for spontaneous copolymerizations in systems involving 1-

    alkoxy-l,3-butadienes, I-methoxy-l,3-butadiene should prove a good choice for further

    investigations of diradical intennediates. On reaction with the same trisubstituted olefins

    (3-5) used in the arylbutadlene reactions, it is expected that both copolymer and

    cycloadduct will be obtained. The fonner should result from initiation by the diradical

  • 57

    z u"= -~'--V

    , I

    1

    1 1 z u

    "= :z u -..... 0

    + -..... o~

    Scheme 16 Butler's Proposed Mechanism

  • S8

    intermediate. The later could result from either closure of the diradical intermediate or

    from a concerted reaction.

    Results

    Control Reactions

    I-Methoxy-l.3-butadiene was heated at 66°C for seven days under Ar. No

    reactions of any kind. neither spontaneous polymerization nor cycloaddition were

    observed. I-Methoxy-l.3-butadiene was also heated at 66°C in the presence of AIBN

    for seven days. Again. no reaction occurred. The reaction of I-methoxy-l.3-butadiene

    and trimethyl ethylenetricarboxylate was initiated with AIBN. Both 1:1 copolymer and

    two isomeric [4 + 2] cycloadducts were obtained in 31% and 69% yields respectively.

    The I-methoxy-l.3-butadiene/acrylonitrile system was also initiated with AIBN. Low

    molecular weight copolymer was obtained in 12% yield. Two isomeric cyclohexenes

    were obtained in 69% yield.

    Reactions of I-Methoxy-l,3-Butadiene and Maleic Anhydride

    I-Methoxy-l.3-butadiene and maleic anhydride were reacted in 2M dioxane solution

    at 25°C under argon. The results are shown in Table 7. A 1:1 diene/olefin ratio was

    used. A bright yellow charge transfer complex was observed on mixing. After an hour

    this yellow color was completely gone. The reaction mixtures were precipitated into

    ether containing a small amount of inhibitor. Both copolymer 30 and [4 + 2]

    cycloadduct 31 were obtained.

    Viscosity measurements were done to determine if copolymer molecular weight

    increased with time. Inherent viscosities were found to increase with increasing reaction

    time. This is evidence for a radical reaction terminating by coupling.

  • 59

    Table 7 Reactions of I-Methoxy-l,3-Butadiene and Maleic Anhydride

    Time Cycloadduct Copolymer (h) Yield % ~1nh Yield % ( l/g)

    1 69 8 3.26

    2 80 10 1.99

    3 72 8 2.44

    4 73 7 2.97

    5 74 4 3.02

  • 60

    Reactions of I-Methoxy-l,3-Butadiene and Methyl 3,3-Dicyanoacrylate

    I-Methoxy-l.3-butadiene 28 and methyl 3.3-dicyanoacrylate 3 were reacted in bulk

    and in SM solutions of dichloroethane and methanol. The results are shown in Table 8.

    AI: 1 diene/olefin ratio was used. A charge transfer complex was formed as evidenced

    by the bright yellow color of the reaction mixwre on initial mixing. As in the

    arylbutadiene reactions discussed in the preceding chapter. the bulk reaction became too

    viscous at the onset of the reaction to allow adequate mixing of the diene and olefin.

    The reaction mixture was warmed at 80°C for a few minutes to allow the stining of the

    reaction mixture. Both copolymer 32 and [4+2] cycloadduct 33 were obtained in bulk

    and in dichloroethane. Only [4+2] cycloadduct was obtained in methanol.

    The copolymer composition is approximately 1:1 in each monomer according to the

    elemental analysis. IH NMR spectra of low molecular weight fractions show that the

    predominant structure is 1.4. The copolymers tum yellow and become insoluble on

    exposure to air.

    The cyclohexene adduct. S-carbomethoxy-4,4-dicyano-3-methoxy-l-cyclohexene 33

    was obtained as a mixture of two isomers. The isomer ratio was determined to be 1:2

    by gas chromatography. The structure of the major isomer was determined by detailed

    analysis of the NMR spectra. 15.6a was found to be 11.3 Hz and 15.6a was found to be

    S.9 Hz. These coupling constants indicate that H5 is in the pseudoaxial position and

    that the ester group is in the pseudoequatorial position. If ~ were equatorial. both

    coupling constants would be expected to be between 0 and S Hz. H3 was determined to

    be in the axial position by comparing Barfield's expected homoallYlic coupling constants

    with experimental data.4P

  • 61

    Table 8. Reactions of 1-Methoxy-1,3-Butadiene

    Olefin Solvent Conc. Time Temp. Cycloadduct Copolymer CMI ChI C'CI yield ('1 yield ('I

    3 bulk 16 26 90 10

    bulk 16 26 92 5

    DCE 5 16 26 90 8

    DCE 5 67 26 91

    CH30H 5 16 26 42 0

    4 bulk 16 26 59 6

    bulk 47 26 57 40

    DCE 5 47 26 50 20

    DCE 5 67 26 84

    29 bulk 72 26 57 40

    DCt 5M 72 26 50 20

  • 62

    Reactions of I-Methoxy-l,3-Butadiene with Dimethyl Cyanofumarate

    I-Methoxy-l,3-butadiene and dimethyl cyanofumarate were reacted in bulk and in

    SM dichloroethane solution. The results are shown in Table 8. A charge transfer

    complex was formed as evidenced by the yellow color of the reaction mixture. All of

    the dimethyl cyanofumarate did not initially dissolve in the diene when no solvent was

    used. The reaction mixture was warmed at 80°C for a few minutes to facilitate the

    dissolving. Both copolymer 34 and [4+2] cycloadduct 35 were obtained.

    The copolymer composition is approximately 1:1 in each monomer as evidenced by

    elemental analysis. The copolymer is of too high molecular weight to obtain a highly

    resolved 1H NMR spectrum, so the mode of addition is not known. (probably mostly

    cis and trans 1,4) The copolymer turns yellow and becomes insoluble on exposure to

    air.

    The cyclohexene adduct, 4,S-wcarbomethoxy-4-cyano-3-methoxy-l-cyclohexene 35,

    was obtained as a mixture of two isomers in a 1.4: 1 ratio as determined by gas chroma-

    tography. The structure of the major isomer was determined in the same way as that

    for the adduct from methyl 3,3-dicyanoacrylate - by detailed analysis of the NMR

    spectra. J5.6& was determined to be 12.1 Hz and J5,60 to be S.6 Hz. These values agree

    with the coupling constants for axial-axial and axial-equatorial couplings respectively.

    Analysis of homoallylic coupling constants show H3 to be axial, and thus the methoxy

    group to equatorial.

    Reactions of I-Methoxy-l,3-Butadiene and Trimethyl Ethylenetricarboxylate

    I-Methoxy-l,3-butadiene and trimethyl ethylenetricarboxylate were reacted in bulk

    and SM, 1M, and O.SM solutions of dichloroethane at various temperatures. The

    results are shown in Table 9. At 2SoC, both copolymer 36 and [4 + 2] cycloadduct 37

    were obtained in bulk and SM solution. In 1M and O.SM solutions of dichloroethane at

  • 63

    room temperature, only cycloadduct was fonned. At 70°C, and at longer reaction times,

    both copolymer and cycloadduct were obtained. At 1:1 feed ratios. copolymer yield

    increased with increasing temperature. The molecular weight of the copolymer was

    quite high as shown in Table 10.

    The molecular weight of the copolymer was studied as a function of time. The

    results are shown in Table 11. At 28°C, copolymer yield increased with increasing

    reaction time. Inherent viscosity, measured at 32.SoC in chlorofonn showed no increase

    in molecular weight with time. Reactions were done at a 2:1 feed ratio of I-methoxy-

    1,3-butadiene to trimethyl ethylenetricarboxylate. No difference in product composition

    was noted. However, the molecular weight of the copolymer decreased.

    Reactions were also done in acetonitrile and methanol. Only 1 % copolymer was

    isolated from the reaction on acetonitrile. The molecular weight of this polymer was

    significantly lower than the polymer from reactions done in dichloroethane. No

    significant difference was noted in methanol.

    The copolymer composition was 1:1 according to the elemental analysis. The

    copolymer appeared to be hygroscopic. Analyses not done immediately after heating at

    100°C gave erroneous results. The IH NMR was inconclusive as far as copolymer

    structure was concerned. At high concentrations the copolymer appears to have either

    trans- 1,4 and/or cis-l,4 structure. At lower concentrations, at higher temperatures and

    in methanol solution, 3,4 addition competes with the 1,4 addition. Unlike the

    copolymers obtained in reactions of methyl 2,2-dicyanoacrylate and dimethyl

    cyanofumarate, these copolymers did not tum yellow and crosslink on exposure to air.

    The [4 + 2] cycloadduct was obtained as a mixture of two isomers. These isomers

    could not be resolved by gas chromatography, but analysis of the IH NMR spectrum

    showed the isomers present in equimolar amounts.

  • 64

    Table 9 Reactions of 1-Hethoxy-1,3-Butadiene vith Trimethyl Ethylenetr!carboxylate

    HeOBD/TriE Solvent Conc. Time Temp. Cycloadduct Copolymer ratio (H) (h) ('C) yield It) yield It)

    1:1 bulk 16 25 57 50

    DCE 5 16 25 77 22

    DCE 1 16 25 81 0

    DCE 0.5 16 25 59 0

    DCE 5 63 70 50 40

    2:1 DCE 5 16 25 79 12

    DCE 5 16 57 74 10

    1:1 CH3CN 5 13 25 80 1

    CH30H 5 13 25 70 20

  • 65

    Table 10 Molecular Weight of HeOBD/TriE Copolymers

    Feed Ratio Solvent Temp. Copolymer Molecular 'C yield It) Weight Hn

    1:1 DCE 25 22 1,231,000

    2:1 DCE 25 12 320,000

    1:1 DCE 57 25 364,000

    2:1 DCE 57 10 286,000

    1:1 CH3CN 25 1 87,000

    1:1 CHJOH 25 20 1,130,000

  • Table 11 Molecular Weight of MeOBDtrNE Copolymers as a FWlction ofThne

    Time Copolymer :J 1nh (h) Yield % (al/g)

    5 7 2.45

    6 10 1.53

    8 16 2.06

    13 47 1.57

    66

  • Reactions of 1.Methoxy·l,3.Butadiene with Acrylonitrile

    1·Methoxy·l,3-butadiene and acrylonitrile were reacted in bulk and in SM

    dichloroethane solution at room temperature for 72 hours. The results are shown in

    Table 8. Both copolymer and [4+2] cycloadduct were obtained in all reactions.

    Polymer yield decreased with decreasing concentration as expected.

    67

    The copolymer 38 appears to be predominantly cis-l,4 and trans-l,4. The 3,4

    structure appears to be present also. The copolymers tum yellow and become insoluble

    when exposed to air.

    The cycloadduct 39 was obtained as a mixture of two isomers as detennined by

    gas chromatography. In the bulk reactions the isomer ratio was approximately 1:1. In

    SM solution, it was 1 :3. These isomers are those where the cyano group is adjacent to

    the methoxy group and are either cis or trans to it, analogous to the cyclohexenes

    discussed previously.

    Reactions of l·Methoxy·l,3·Butadiene With Oletlns Containing p-Leaving Groups

    I-Methoxy-I,3-butadiene was reacted with 2,2-dicyanovinyl chloride and 2·

    carbomethoxyvinyl chloride in SM dichloroethane solutions. In both cases, traces of [4

    + 2] cycloadduct were obtained. The major product of these reactions was a black,

    insoluble solid substance. Attempts to dissolve it in acetone, chlorofonn, DMSO, and

    DMF were unsuccessful. The reactions were repeated using only 10% of the olefin.

    The results were the same. An intractable black solid was the major product.

    Discussion

    I-Methoxy-l,3-butadiene proved to be a better diene for the study of diradical

    intennediates in the spontaneous polymerizations of electron-rich dienes with electron-

    poor olefms. I-Methoxy-l,3-butadiene was reacted with three trisubstituted electrophilic

  • olefms and acrylonitrile at varying concentrations and reaction conditions. Copolymer

    and [4+2] cycloadduct were obtained in the reactions with all four olefins. Since

    copolymer is formed. there is evidence for a diradical intermediate.

    68

    The copolymer composition was approximately 1: 1 for all reactions. This is to be

    expected if the hexamethylene mechanism is operating in this reaction. The polar

    diradical should favor the alternating structure.

    The copolymers have a predominantly 1,4 structure. The copolymers containing

    cyano groups are readily oxidized in air. Care should be taken to insure that polymers

    be stored in sealed ampules.

    The cycloadducts were again formed as a mixture of two isomers. No loss of

    stereochemistry was observed in the reactions using dimethyl cyanofumarate. Therefore.

    there is no evidence that the cycloadduct is formed from the diradical intermediate.

    The reac


Recommended