Skip to main content

Full text of "Organic seminar abstracts"

See other formats


"L  I  B  HAHY 

OF  THE 
UN  IVERSITY 
Of    ILLINOIS 

Q,  54-7 

libs 

966 -67 

pt.2 


The  person  charging  this  material  is  re- 
sponsible for  its  return  on  or  before  the 
Latest  Date  stamped  below. 

Theft,  mutilation,  and  underlining  of  books 
are  reasons  for  disciplinary  action  and  may 
result  in  dismissal  from  the  University. 

University  of  Illinois  Library 


jm-nrm 


L161  — O-1096 


Digitized  by  the  Internet  Archive 

in  2012  with  funding  from 

University  of  Illinois  Urbana-Champaign 


http://archive.org/details/organicsemi1966672univ 


ORGANIC  SEMINAR  ABSTRACTS 
1966-67 


Semester  II 


Department  of  Chemistry  and  Chemical  Engineering 
University  of  Illinois 


0  7 

SEMIHAR  TOPICS 

II  Semester  1966-67 

Mechanisms  in  the  Biological  Chemistry  of  Pyrophosphate  Esters 

W.  Tc  Shier  221 

The  Configurational  Stability  of  ALkenyl  Radicals 

Peter  M»  Harvey  230 

Isomerization  of  Organic  Thioeyanates  to  Isothiocyanates 

Joseph  Co  Stickler  238 

The  Physiologically  Active  Constituents  of  Marihuana 

Donald  Co  Schlegel  2V7 

Benzene  Photolysis 

Warren  J«  Peascoe  255 

The  Photodimerization  of  Thymine 

Sheldon  A.  Schaff er  263 

Thermal  Rearrangements  of  Cycloheptatrienes 

W.  Do  Shermer  272 

Cyclization  Reactions  of  N-haloamines ,  -Amides ,  and  -Imines 

Daniel  RB  Bloch  28l 

The  Thermal  Endo-Exo  Isomerization  of  Some  Diels-AIder  Adducts 

Tommy  L.  Chaffin  290 

Reactions  of  NH  Radicals 

Terry  Go  Burlingame  296 

Geometric  Isomerism  in  Diazoketones 

Daniel  B»  Pendergrass  305 

Eo    S.   R.    Studies  of  Organic  Ground-State  Triplet  Molecules 

Robert  Jo  Basalay  313 

Recent  Studies  Concerning  the  Mechanism  of  the  Favorskii  Rearrangement 

Peter  Ao  Gebauer  322 

Prostaglandin  Syntheses 

Edward  Bertram  331 

Possible  Yinyl  Cation  Intermediates 

David  Ao  Simpson  338 

Sigmatropic  Reactions 

Rs  H.  Watson  3^7 

Photochemistry  of  Cyclobutanones  and  Cyclobutanediones 

Edward  F»  Johnson  356 


-  2  «• 

The  Mechanism  of  Papain  Catalysis 

Paul  Elliot  Bender  365 

The  Photosensitized  Cis-Trans  Isomerization  of  Olefins 

Robert  Kalish  373 

The  Abnormal  Claisen  Rearrangement 

James  E0  Shaw  382 


-221- 

MECHANISMS  IN  THE  BIOLOGICAL  CHEMISTRY  OF  PYROPHOSPHATE  ESTERS 
Reported  by  W.  T.  Shier  February  23,  19&1 

Introduction; 

For  decades  physical  organic  chemists  have  studied  the  mechanisms  of  the 
reactions  they  conducted  in  their  flasks ,  often  determining  to  a  high  degree 
of  certainty  the  detailed  path  of  the  atoms  and  electrons  involved  in  these 
conversions.   In  contrast  the  mechanisms  of  the  reactions  of  many  biologically 
important  functional  groups  have  been  largely  ignored  even  to  the  present  day. 
One  such  group  is  the  pyrophosphate  group. 

The  pyrophosphate  esters  found  in  nature  may  be  represented  by  the  for- 
mula 

0  0 
II   II 

r„0-P-0-P-0-R' , 

1  I 
OH  OH 

where,  depending  on  the  stratagems  of  nature,  R  or  R'  may  be  as  simple  as  a 
proton  or  as  complex  as  a  nucleoside.   The  pivotal  role  of  phosphorus  compounds 
in  the  chemistry  of  life  was  first  recognised  by  Fritz  Lipmann,  and  represents, 
perhaps,  one  of  his  greatest  contributions  to  the  understanding  of  the  chem- 
istry of  biological  systems.   As  will  be  shown,  pyrophosphate  esters  and  the 
pyrophosphate  ion  itself  are  endowed  with  unique  biochemical  properties. 

These  compounds  have  heretofore  invariably  been  considered  in  various  other 
contexts,  usually  based  on  some  non-functional  portion  of  the  molecule.   When 
considered  as  pyrophosphate  esters,  they  represent  a  large  and  disjointed  group 
of  natural  products.  A  method  of  organizing  and  correlating  this  heterogeneous 
array  is  offered  by  a  rapidly  emerging  branch  of  Organic  Chemistry,  Bioorganic 
Mechanisms  -■  that  hybrid  of  Physical  Organic  Chemistry  and  Enzymology1.   On 
the  basis  of  mechanistic  aspects  of  the  biological  function  of  these  compounds 
they  have  been  correlated  into  three  groups  and  their  biological  reactions  or- 
ganized by  three  general  equations. 

The  Role  of  the  Pyrophosphate  Ion  as  a  Leaving  Group: 

Th.e   pyrophosphate  ion  functions  as  a  leaving  group  in  many  biosynthetic 
coupling  reactions.   Typical  is  the  biosynthesis  of  nucleoside  coenzymes2, 
which  can  be  represented  schematically  as: 

nucleoside-O-P-P-P  +  R-O-P  ^  nucleoside-0-P-P-OR  +  PP(* 

The  role  of  the  pyrophosphate  ion  as  a  leaving  group  in  a  similar  substitution 
reaction  has  been  studied  by  Lipmann.3  In  the  biosynthesis  of  adenosine-5 ' - 
phosphosulfate  (APS)  in  a  yeast  extract  preparation,  the  thermodynamics  of  the 


Abbreviations  and  structural  symbols  used  in  this  work: 

-OPP,  pyrophosphate  monoester;  P<*,  inorganic  phosphate;   OPP,  inorganic  pyro- 
phosphate; Enz-SH,  sulfhydryl-containing  enzyme;  WAD,  nicotinamide  adenine 
dinucleotide;  MDP,  ( TPN)  nicotinamide  adenine  dinucleotide  phosphate;  PRPP, 
5-phospho-ribosyl-l-pyrophosphate;  -OP,  phosphate  monoester;  CDP,  cytidine  di- 
phosphate; CMP,  cytidine  monophosphate;  UDP,  uridine  diphosphate;  CoA,  coen- 
zyme A;  TPP,  thiamine  pyrophosphate;  FAD,  flavin  adenine  dinucleotide;  ATP, 
adenosine  triphosphate;  ADP,  adenosine  diphosphate;  AMP,  adenosine  monophos- 
phate; DCC,  dicyclohexylcarbodiimide. 


-cLdei- 

equilibrium 

ATP  +  S04=  ATP-sulfurylase  Aps  +  pp, 

strongly  favor  the  reverse  reaction,   When  pyrophosphatase  was  added  to  a  pur- 
ified ATP-sulfurylase  system,  the  equilibrium  concentration  of  APS  increased 
from  O.OluM./ml.  to  0«23uM,/ml,3  Thus,  the  pyrophosphatases,  which  catalyse 
the  very  exothermic  hydrolysis  of  the  pyrophosphate  ion,  provide  an  energy 
coupled  system  with  a  sufficient  over-all  free  energy  drop  to  favor  APS  synthesis, 

Pyrophosphate  Esters  as  Coupling  Intermediates: 

Polyisoprenoids  can  be  considered  to  be  biosynthesized  by  means  of  the  poly- 
merization in  defined  modes  of  A^isopenteny]  pyrophosphate  monomer.4  The  inter- 
mediacy  of  these  pyrophosphate  esters  in  the  biosynthesis  of  squalene  has  been 
demonstrated  by  Lynen  and  his  collaborators. 5>6  When  2i4C»mevalonate-5-phosphate 
was  added  to  a  crude  enzyme  system  from  yeast,  labeled  farnesyl  pyrophosphate 
and  geranyl  pyrophosphate  were  isolated  by  paper  chromatography  and  identified 
by  comparison  with  synthetic  standards,5  When  the  system  was  inhibited  with 
iodoacetamide  A-isopentenyl  pyrophosphate  accumulated.6  A  sulfhydryl-containing 
enzyme  was  isolated  in  crude  form  which  converted  A-isopentenyl  pyrophosphate 
into  dimethylallyl  pyrophosphate.   The  proven  and  postulated  products  of  A3~ 
isopentenyl  pyrophosphate  polymerization  are  summarized  in  the  following  table:7 

Degree  of  Polymerization: 


n=l 


-OPP 


A-isopentenyl  pyrophosphate 


-> 


-OPP 


n=2 


~ DPP 


Geranyl  pyrophosphate 


-^   Monoterpenoids 


n=3 


Farnesyl  pyrophosphate 


^   Sesquiterpenoids 
Squalene 


n=k 


n=10 


n=n 


Geranyl-geranyl  pyrophosphate 

CH3 
H(-CH2-C=CH-CH2)-ioOPP 

CH3 
H(-CH2-C=CH-CH2> 


-a,  Diterrenoids 

C>> Complex  Lipids  15 
^iphytoene 


-^.   Ubiquinones 


-^   Gutta  percha 


The  driving  force  for  the  condensation  of  monomer  units  has  been  attributed 
to  the  unique  effectiveness  of  the  pyrophosphate  ion  as  a  leaving  group.8 
It  is  contended  here  that  this  ability  may  result  not  so  much  from  any  chemical 
property  of  the  pyrophosphate  ion,  as  from  its  rapid  removal  from  the  system 
by  highly  active,  widely  distributed  pyrophosphatases.   The  pyrophosphate  con- 
centration can  reasonably  have  a  profound  effect  on  the  pyrophosphate  ion-elim- 
inating mechanism  in  the  enzyme -substrate  complex,  thus  inhibiting  the  overall 
reaction,   Again,  the  pyrophosphatases  may  energy  couple  the  hydrolysis  of  the 

jj^jL-«jphutijJua.bc    ICii    uu    caxuuj.i-uctx'LujLi   bond    f  Oj.iuaLiui.u 


These  condensation  reactions  can  be  summarized  by  general  equation  I: 


R-OPP  +  H-S:  qrr-rt  R-S: 


+  HOPP 

Pyrophosphatase 

pp(- 


l] 


In  this  equation  it  is  tempting  to  consider  H-S:  a  general  nucleophile. 
Since  Yuan  and  Block  observed  a  63^  conversion  of  l-3H-A3-isopentenyl  pyrophos- 
phate  to  squalene  in  a  yeast  autolysate,9  the  rate  of  the  biological  reaction 
is  greater  than  the  rate  of  non-enzymatically  assisted  dissociation,  which  would 
result  in  hydrolysis  to  isopentenol.   Hence,  a  strict  SnI  mechanism  is  ruled 
out,  although  an  anchimerically  assisted  dissociation  of  the  pyrophosphate 
ester  as  the  enzyme-substrate  complex  is  note   The  absence  of  any  isomeric 
reaction  product4  arising  from  allylic  rearrangement  in  squalene  biosynthesis 
also  argues  against  any  SmI  mechanism,  although  Ingraham  has  suggested  that 
the  cationic  species  may  be  bound  to  the  surface  of  the  enzyme  in  a  manner  that 
permits  further  attack  only  at  the  terminal  carbon.10  It  has  been  also  sug- 
gested8 that  a  cationic  species  is  formed,  but  that  its  formation  is  concerted 
with  the  formation  of  a  carbon-carbon  bond;  that  this  is  anything  other  than 
an  Sn2  mechanism  is  not  made  clear. 


General  equation  I  can  be  seen  to  describe  the  polymerization  of  A^iso- 
pentenyl  pyrophosphate  by  considering  head-to-tail  condensation  as  resulting 
from  nucleophilic  attack  by  the  v.     electrons  of  the  double  bond  of  the  monomer 
either  directly  on  the  ester if ied  carbon  of  the  growing  chain12,  or  on  an  enzyme- 
growing-chain  complex: 


OPP 


OPP 


+ 


HOPP 


The  head-to-head  reductive  condensation  of  farnesyl  pyrophosphate  to  pro- 


duce squalene  according  to  Cornforth's  hypothetical  scheme13  can  be  seen  to 
follow  general  equation  I: 


CH3  jOPP  H 
RCH2C=CH-CH2  ^  S-Enz 


CH3 
">  RCH2C=CHCH2- S-Enz 


CH2-CH— CCH2R 


£opp 


CH- 


CH* 


B: 

H 
I 


r\ 


->   RCH2C=CH^CH- S-Enz 


Steven1 s 
Rearrangement 


CH- 
I 


-H© 


CHpCH— CCHpR 


PPO-CH2CH=CCH2R 

CH3   C  CH3 

I 

RCHpC=CHCH-S-Enz 


m3 

CH2CH=CCH2R 


l~\<-\ ), 


CH2CH=CCH2R 
I 

CH3      w  CH3 

RCH2C=CHCH-?S~Enz 

V 

(t     CH2CH=C-CH2R 
H^H  I 

H2N0CO>S  CH^ 

I 


CH< 


CH2CH=CCH2R 


I 


RCH^CH-CH^  +  S-Enz 

CH2CH=CCH2R 
» 
CH3 


\ 


(R=Geranyl;  B=proton  acceptor,  possibly  part  of  the  active  site) 

This  mechanism  satisfies  a  considerable  body  of  compelling  but  not  con- 
clusive evidence  obtained  largely  by  Popjak  and  collaborators,14  The  obser- 
vations and  conclusions  are: 

(1)  The  biosynthesis  of  squalene  from  farnesyl  pyrophosphate  by  washed  rat 
liver  microsomes  was  powerfully  inhibited  by  p-chloromercuri-benzoate,  N-ethyl- 
maleimide  and  Cu  ions  but  not  by  iodoacetamide.   Hence,  the  active  site  of  the 
enzyme  contains  a  functional  S-atom,  but  not  a  free~SH,  which  would  permit  in- 
hibition by  iodoacetamide.14 

(2)  When  squalene  was  biosynthesized  in  the  same  system  from  5-2H2-mevalonate, 
11  atoms  of  2H,  not  the  theoretically  possible  12,  were  retained.  Mass  spec- 
troscopic analysis  of  succinate  derivatives  obtained  from  the  centre  of  the 
chain  by  ozonolysis  of  the  deutero squalene  showed  mostly  trideutero  iaolecules. 
Hence,  the  labeling  at  the  centre  is  -CHD-CD2-.14 

(3)  In  the  biosynthesis  of  squalene  from  farnesyl  pyrophosphate  no  tritium 
from  H3HO  was  incorporated  into  squalene,  while  incorporation  of  up  to  0.8 
ug-atom  of  the  tritium  per  ug»mole  of  squalene  from  labeled  TPN3H  was  observed.14 
Hence,  the  condensation  is  a  reductive  process  involving  TPNH. 

In  Cornforth's  mechanism,  in  the  steady  state  all  the  experimental  obser- 
vations are  satisfied.   It  may  further  be  noted  that  in  each  -0PP  eliminating 
step  general  equation  I  is  followed.   That  is 


H 


R 


Enz 
I 

S:   +  ROPP 


•OPP 


II 


R' 


Enz 
S-R 


-H 


S-Enz 


Where  R=farnesyl  and  R'  =  H  or  farnesyl. 

Reactions  of  pyrophosphate  esters  other  than  the  synthesis  of  carbon-carbon 
bonds  can  be  shown  to  follow  general  equation  I.   In  the  biosynthesis  of  a  com- 
plex phospholipid  from  Halobacterium  cutirubrum  an  ether  linkage  is  formed  by 
phytanyl  pyrophosphate  and  an  -OH  of  glycerol-1-phosphate.15 

i.e.    R-OPP  +  H-0-C-  >  R-0-C-    +    HOPP 

Since  the  biosynthetic  studies  were  carried  out  in  a  crude  cell-free  extract 
with  the  above  reaction  monitored  only  by  analysis  of  32P  incorporation  into 
the  final  product,  no  detailed  mechanism  for  the  formation  of  the  ether  linkage 
has  been  put  forward.   It  was  observed,  however,  that  the  use  of  phytanyl  pyro- 
phosphate removed  the  requirement  for  ATP  in  the  biosynthesis.  Hence,  the  es- 
terified  pyrophosphate  group  supplies  the  energy  for  condensation.   It  can  read- 
ily be  seen,  also,  that  the  reaction  conforms  to  general  equation  I. 

In  the  biosynthesis  of  nucleotides,  PRPP  condenses  with  glutamine,  orotic 
acid,  purines,  or  pyrimidines  as  follows: 


P-0-H2C   0 


+ 


\  l  / 

N 


POH 


V 


opp 


+ 


OPP 


OH   OH 


OH  OH 


The  problem  of  strict  application  of  the  terms  of  physical  organic  chemistry 
arises-   Despite  the  fact  that  the  leaving  group  is  a  to  a  heteroatom16  a  strict 
SnI  reaction  is  highly  unlikely.   According  to  the  Michaelis-Menten  hypothesis,17 
in  the  biological  reaction  only  PRPP  is  bound  to  the  enzyme;  the  N-containing 
species  also  binds  to  the  enzyme  surface  where  reaction  occurs.  An  SH2  displace- 
ment of  OPP  by  the  lone  pair  of  the  N-containing  species,  while  it  and  PRPP 
are  both  bound  to  the  enzyme  is  the  simplest  mechanism,  although  not  the  only 
one  possible. 

It  is  observed  that  inversion  of  configuration  occurs  establishing  the 
stereochemistry  observed  in  nucleotides.18  This  inversion  of  configuration 
is  consistent  with  an  Sm2  mechanism,  but  not  conclusive  evidence  for  it.11 

Group  Transfer  Intermediates: 

The  nucleoside  pyrophosphate  diesters  represent  a  large  and  rapidly  increas- 
ing group  of  biosynthetic  intermediates.   While  the  basic  forms  -  the  nucleoside 
diphosphate  -sugars,  -alcohols  and  -diglycerides  -  were  known  a  decade  ago,  and 
extensively  reviewed  then,19  studies  demonstrating  their  universal  involvement 
in  the  biosynthesis  of  complex  lipids,  oligosaccharides,  homo-  and  hetero-poly- 
saccharides  in  all  levels  of  life,  from  the  cell  walls  of  Neurospora  crassa20 
to  the  complex  lipids  of  the  human  brain,21  have  occupied  the  time  of  hundreds 
of  research  workers  in  the  intervening  years.   To  introduce  order  into  this 
thriving  jungle  of  biosynthetic  pathways,  the  following  general  equations  are 
presented: 


II 


0  0 
H   II 

R-0-P-0-P~OCH2 

1  '     s  0-    5 
OH  OH 


HO   Y 


+  R'OH 


v 


0  0 
II  B 

R-O-ROR'  +  HO-P-OCH2 

1  I 
OH        OH 


III 


0   0 
II   'I 
R-O-P-0-P-0CH2     B 

"  0 


OH  OH 


y 


OH   Y 


+  R'OH 


0   0 
W        II 
-^R-O-R1   +  HO-P-0-P-OCH 


OH  OH 


w 


Oh  y 


Y=  -H  or  -OH;  B=  purine  or  pyrimidine 


-d.cLO- 

Neither  of  these  type  reactions  has  received  direct  mechanistic  study. 
Results  obtained  for  analogous  systems  will  be  considered,  and  the  results  ob- 
tained tentatively  applied  to  the  two  general  equations  pending  the  appearance 
of  systematic  studies  of  these  systems. 

The  alcoholysis  of  anhydride  phosphate  linkages  has  not  been  studied  in 
biological  systems,22  but  Cohn  showed  that  180  labeled  phosphate  underwent 
isotope  exchange  with  the  carboxylate  oxygen  atoms  of  3-phospho-glycerate  in 
the  following  oxidative  phosphorylation  of  ADP23 

0    0* 

'i  *  li  *e 

: *■   R-C-0  -P-0  +  H  +  +  DPMI 


0 

0 

R- 

II 
-CH 

*  11    ■ 

+  HO  -P-0 

i* 

0  H 

X©                      1 

+  DPN 

B 

= 

-CH( OH) CH2 

0P03H2 

0 

* 

0 

R- 

II 
-C-i 

*   ll       *Q 

D  -P-0       + 
U 
0  H 

(ADP)-0H 

I  * 
0  H 

*         * 
0  0 

s    N   R-C>  0  +  (ADP>0-P-0 

y0  OH 

The  reverse  of  the  second  reaction  was  interpreted  as  nucleophilic  attack 
by  the  carboxylate  oxygens  on  the  terminal  P  atom  of  ATP22  That  is, 

0 
0       0  0  HO  0  0   0 

/:  w  \\     \/,  il   II        N  e 

-CI6  +  H-0-P-0(-ADP)   »    -C-0-P-0(-ADP)         N    -C-O-P-OH  +  (ADPfO 

\\        1         < I  ^  I 

0       OH  OH  OH 


Extending  this  finding  to  the  alcoholysis  of  anhydride  phosphate  linkages, 
nucleophilic  attack  by  the  oxygen  of  the  alcohol  on  the  appropriate  phosphor- 
us atom  can  be  put  forward  as  a  working  hypothesis  for  further  research.   One 
experiment  which  suggests  itself  consists  of  feeding  R180H  to  an  active  cell  free 
extract  containing  the  appropriate  nucleoside  diphosphate  ester;  identification 
of  180  in  excess  in  the  phosphate  diester  and  not  in  the  nucleoside  monophosphate 
would  support  the  hypothesis. 

In  a  transfer  reaction  analogous  to  general  equation  HI  Douderoff,  Barker 
and  Hassid  explained  the  isotope  exchange  between  KH232P04  and  glucose -1-phosphate 
in  a  sucrose  phosphorylase  preparation  from  Pseudomonas  saccharophila  in  the 
absence  of  fructose,  the  phosphate  acceptor,  in  terms  of  a  glucosyl-enzyme 
complex.24 


i.e.   Glucose -I-OPO3H  +  Enzyme       ^    Glucose -Enzyme  +  H2P04~ 
(Glucose-1-032P03H)  (H232P04") 


The  analogous  substrate-enzyme  complex  has  been  suggested  for  systems  described 
by  general  equation^H.25  The  irreversibility  of  the  reaction  prevents  the  use 
of  32P-labeled  nucleoside  pyrophosphate  in  an  experiment  analogous  to  that  of 
Douderoff  et  al.24 

The  interesting  question  of  whether  P-0  or  C-0  cleavage  occurs  could  be 
answered  by  preparing  the  appropriate  nucleoside  triphosphate  labeled  with 
180  in  the  phosphate( s) ,  adding  it  to  a  cell-free  extract,  and  analysing  R-0-R' 
for  excess  180;  excess  l80  would  indicate  P-0  cleavage. 

The  nucleoside  pyrophosphates  function  as  transfer  intermediates  in  some  reactions, 


For  example,  the  following  kinase  catalysed  reaction2  fits  general  equation  II: 

2  nucleoside -OPP  - — *»  nucleoside-OP  +  nucleoside-OPPP 

This  reaction  is  general  to  both  nucleoside  and  deoxynucleoside  pyrophosphates.' 

Phosphatides  are  biosynthesized  from  CDP-diglycerides26  according  to  the 
scheme  of  general  equation  II: 

0 

CH2OCOR 

I 

CHOCOR, 


0 


CH20-P-0-P-QCH2  0> 


0   0 


+ 


R'-OH 


x- 


CH20C0R 
CHOC OR _ 
CHoO-P-OR'  +  CMP 


where  R  =  CH3(-CH2)-i4,  for  example,  and  R'OH  =  L-a  glycerol,  L-serine,  myo-inositol 
or  phosphatidyl  glycerol. 

Other  cytidine  diphosphate  alcohols  also  undergo  enzymatic  conversions26 
according  to  the  scheme  of  general  equation  II: 


0 
II 
CDP-OR  +  R'OH  rrr^  R-O-P-OR'  +  CMP 

where  ROH  is  ethanolamine  or  choline  in  an  alternate  synthesis  of  lecithins26 
or  a-glycerol  or  5-ribitol  in  the  biosynthesis  of  bacterial  cell  wall  polymers.27 
The  nucleoside  diphosphate-pentoses  and  -hexoses  represent  the  largest  group 
of  transfer  intermediates1  -  they  are  found  for  most  of  the  known  nucleosides 
and  deoxynucleo sides.19  The  sugar  moieties  include  hexoses,  pentoses, glycur- 
onic  acids,  hexosamines,  mucopeptides  and  oligosaccharides2  attached  to  the 
pyrophosphate  moiety  at  the  anomeric  carbon.   They  undergo  reaction  according 
to  the  scheme  of  general  equation  III.  For  example,  the  synthesis  of  glycogen 
from  UDP-glucose  in  rat  liver:29 


n 


CH2OH 


glucosyl 
transferase 


UDP)V 


+ 


n  UDP 


"Non-Functional"  Pyrophosphate  Esters: 

This  group  of  pyrophosphate  esters  consists  largely  of  the  coenzymic  f orm( s) 
of  the  B  vitamins.30  The  term  "non-functional"  is  somewhat  of  a  misnomer,  for 
the  pyrophosphate  moiety  serves  as  a  linkage  in  some  cases  (e.g.  CoA,  NAD,  FAD, 
etc.) ,  it  may  serve  as  a  binding  site  to  attach  the  coenzyme  to  the  enzyme 
(e.g.  TPP) ,  or  it  may  become  functional  after  conversions  have  been  made  on  other 
parts  of  the  molecule  (e.g.  mevalonate-5-PP) .12  On  any  account,  however,  the 
pyrophosphate  group  does  not  undergo  a  permanent  change  in  the  normal  biological 
function  of  the  molecule. 

In  the  conversion  of  the  B  vitamins  to  their  coenzymic  forms  the  vitamin 
is  usually  incorporated  intact:30 


ooH 


Vitamin  Bx  -»  Thiamine  Pyrophosphate 


0  0 

II    11 
CH2CH2-OrP-0-P-OH 

1  I 
OH  OH 


Pantothenic  Acid  -»  Coenzyme  A 
NH2 


n 


11 


o 


„  0  OH  CH3   i  0  0 
■  II  I   i     ill  )\ 


N 


HS(-CH2>  NfC(-CH2)-  N-C-CH-C-CH2Of-P-OPOH2  C 


I 

1 

Riboflavin  ->  Flavin  Adenine  Dinucleotide 

r---"o  ! 


CH 


^_J 


OH  OH 


HN 


_oj 


^k 


i\i 


Niacin  ->  Nicotinamide  Adenine  Dinucleo- 

NH2  tlde 


CH2  i 

I    I 

HCOH  , 

I    i 

HCOH  , 


N 


N' 


HCOH  '  0   Q 


CH20l-P-0-P-0CH2  n 
OH  OH 


ONH; 


"1 


0  0^^ 

II     11 

CH20-P-0-P-0CH2 

1  I 
OH  OH 


N 


■H  OR 


R  =  H  :   NAD 

R  =  -P03H2:   NADP 


BIBLIOGRAPHY 


1.  For  a  more  complete  definition  cf.  the  Preface  of  T.  C.  Bruice  and  S.  J. 
Benkovic,  "Bioorganic  Mechanisms/'  Vol.  1,  W.  A.  Benjamin,  Inc.,  New  York 
(1966). 

2.  A.  Kornberg,  Advances  in  Enzymol. ,  18,  191  (1957). 

3.  P.  W.  Robbins  and  F.  Lipmann,  J.  Am.  Chem.  Soc,  z£,  6409  (1956). 

k.      J.  R.  Richards  and  J.  B.  Hendrickson,  "The  Biosynthesis  of  Steroids,  Terpenes, 
and  Acetogenins,"  W.  A.  Benjamin,  Inc.,  New  York,  1964,  p.  198. 

5.  F.  Lynen,  B.  W.  Agranoff ,  H.  Eggerer,  U.  Henning  and  E.  M.  Moslein,  Angew. 
Chem.,  II,  657  (1959). 

6.  F.  Lynen,  H.  Eggerer,  U.  Henning,  and  I.  Kessel,  Angew.  Chem,,  70,  13$ 
(1958). 

7.  Modified  from  ref .  k. 

8.  Ref.  k,   p.  201. 

9.  C.  Yuan  and  K.  Block,  J.  Biol.  Chem.,  2j&,  2605  (1959). 

10.  L.  L.  Ingrahan,  "Biochemical  Mechanisms,"  John  Wiley  and  Sons,  Inc.,  New 
York,  1962,  p.  96. 

11.  D.  E.  Applequist,  Personal  communication,  Feb.  k,    1967- 


-229- 

12.  For  a  general  review  cf.,  G.  Popjak  and  J.  W.  Cornforth,  Adv.  Enzymol. ,  22, 
281  (i960). 

13.  cited  in  G.  Popjak,  "Proceedings  of  the  Fifth  International  Congress  of 
Biochemistry/'  G.  Popjak,  ed. ,  Pergamon  Press,  New  York,  1963>  P*  207. 

14.  G.  Popjak,  D.  Goodman,  J.  W.  Cornforth,  R.  H.  Cornforth,  and  R.  Ryhage, 
J.  Biol.  Chem. ,  2J>6,  1934  (I96l). 

15.  M.  Kates,  Personal  communication,  Jan.  5 ,   1967  • 

16.  E.  S.  Gould,  "Mechanism  and  Structure  in  Organic  Chemistry,"  Holt,  Rinehart, 
and  Winston,  New  York,  1959,  P«  284-5  and  footnote  6l. 

17.  For  a  general  discussion  of  the  Michaelis-Menten  hypothesis  cf  .,a.  White,  P. 
Handler,  E„  L.  Smith,  "Principles  of  Biochemistry,"  3^d.  ed. ,  McGraw-Hill 
Book  Co.,  New  York,  1964,  p.  221-8. 

18.  L.  Warren  in  "Metabolic  Pathways,"  Vol.  2,  D.  M.  Greenberg,  ed. ,  Academic 
Press,  New  York,  1961,  p.  4-59. 

19.  J.  Baddiley  and  J.  G.  Buchanan,  Quart.  Revs.,  (London),  12,  152  (1958). 

20.  Ref.  17,  p.  412. 

21.  Ref.  17,  p.  468. 

22.  D.  E.  Koshland  in  "The  Mechanism  of  Enzyme  Action,  "W.  D.  McElroy  and  B. 
Glass,  eds. ,  Johns  Hopkins  Press,  Baltimore,  1954.  p.  608. 

23.  M.  Cohn,  J.  Biol.  Chem.,  201,  735  (1953). 

24.  M.  Doudoroff,  H.  A.  Barker,  and  W.  Z.  Hassid,  J.  Biol.  Chem.,  168,  725  (1947) 

25.  W.  Z.  Hassid  in  "Chemical  Pathways  of  Metabolism,"  D.  M.  Greenberg,  ed. , 
Vol.  1,  Academic  Press  Inc.,  New  York,  1954,  p.  235. 

26.  E.  P.  Kennedy  in  ref.  13,  p.  113. 

27.  J.  Baddiley,  J.  G.  Buchanan,  and  B.  Carss,  Biochim.  Biophys.  Acta,  27,  220 
(1958). 

28.  a)   E.  Cabib,  Ann.  Rev.  Biochem. ,  32,  321  (1963) • 
b)   L.  F.  Leloir,  Biochem.  J.,  £1,  1  (1964). 

29.  L.  F.  Leloir  and  C.  E.  Cardini,  J.  Am.  Chem.  Soc,  79,  6340  (1957). 

30.  R.  J.  Williams,  R.  E.  Eakin,  E.  Beerstecher,  and  W.  Shive,  "The  Biochemistry 
of  the  B  Vitamins,"  Reinhold  Publishing  Corpn. ,  1950,  pp.  123-215. 


•230- 


THE  CONFIGURATIONAL  STABILITY  OF  ALKENYL  RADICALS 
Reported  by  Peter  M.  Harvey  February  27 ,  19&7 

INTRODUCTION 

There  is  evidence  in  the  literature  that  aliphatic  radicals  generated  at 
optically  active  carbon  atoms  are  unable  to  maintain  their  initial  configurations, 
and  that  they  yield  racemic  products.1  In  view  of  the  known  configurational  stabil- 
ities of  alkenyllithiums2  relative  to  alkyllithiums,3  it  is  of  interest  to  examine 
whether  alkenyl  radicals  are  configurationally  more  stable  than  alkyl  radicals. 
This  seminar  will  review  the  evidence  for  the  nonlinearity  of  vinyl  and  substituted 
vinyl  radicals  and  will  discuss  chemical  studies  in  which  isomeric  alkenyl  radicals 
are  generated  and  captured. 

If  it  can  be  established  that  the  vinyl  radical  is  nonlinear  and  does  not  main- 
tain its  configuration,  the  question  of  whether  loss  of  configuration  occurs  by 
rotation  or  by  inversion  is  not  a  trivial  one.  Although  no  experiments  have  been 
reported  which  distinguish  between  the  two  pathways,  one  might  expect  the  isomeriza- 
tion  to  proceed  by  inversion,  with  a  simple  migration  of  the  a-substituent  to  an 
alternate  site  of  high  electron  density,  rather  than  by  rotation,  in  which  the  jt- 
bond  of  the  vinyl  group  must  be  broken.  The  question  of  inversion  vs.  rotation  will 
not  be  dealt  with  further  in  this  review,  and  for  convenience  the  term  "inversion" 
will  be  used  to  refer  to  the  Interconversion  of  isomeric  alkenyl  radicals,  regardless 
of  the  mechanism. 

ELECTRON  SPIN  RESONANCE  STUDIES 

The  ESR  spectrum  of  the  vinyl  radical  should  show  eight  lines  if  the  three 
hydrogen  atoms  are  nonequivalent  or  six  lines  if  the  two  g_-hydrogen  atoms  are 
equivalent.  Fessenden  and  Schuler  irradiated  liquid  ethylene  with  2.0  Ilev  electrons 
in  the  cryostatted  cavity  of  an  ESR  spectrometer.4  Near  10U°K  they  observed  super- 
imposed on  the  twelve -line  spectrum  of  the  ethyl  radical  a  doublet  of  doublets  which 
they  attribute  to  the  vinyl  radical;  the  apparent  hyperfine  splittings  of  102. kk  and 
13=39  gauss  were  not  significantly  temperature  dependent.  The  formation  of  butane, 
1-butene,  and  1-hexene  during  the  electron  irradiation  of  ethylene  at  105-l63°K  has 
been  cited  as  evidence  that  ethyl  and  vinyl  radicals  are  indeed  formed,5  Further- 
more^, vinyl  radicals  have  been  trapped  with  labelled  methyl  radicals  generated  from 
iodomethane-C14  during  the  electron  irradiation  of  liquid  ethylene,6 

The  center  of  the  nine -line  pattern  attributed  to  the  trideuteriovinyl  radical 
from  the  irradiation  of  ethylene ~d4  was  displaced  0.40  gauss  from  that  of  the 
unlabelled  species.  Fessenden  and  Schuler  attribute  this  shift  upon  isotopic  sub- 
stitution to  a  second -order  effect  which  can  be  rationalized  if  the  102.  W-  gauss 
splitting  represents  the  sum  of  two  approximately  equal  coupling  constants.   If  the 
a-proton  of  the  vinyl  radical  is  inverting  with  a  frequency  comparable  to  the 
difference  of  two  nonequivalent  p-proton  splittings,  then  the  inversion  effectively 
interchanges  the  two  [3-protons  and  causes  a  change  in  the  spin  state  when  the  two 
j3»protons  are  antiparallel.  As  a  result,  the  inner  four  lines  corresponding  to 
these  spin  states  are  broadened,,  leaving  a  spectrum  whose  major  splitting  corresponds 
to  the  sum  of  the  two  g-proton  splittings.   If  this  interpretation  is  correct,  the 
13»39  gauss  splitting  can  be  assigned  to  the  a-proton. 

The  vinyl  radical  spectrum  was  also  observed  when  liquid  ethane  containing  0.5$ 
acetylene  was  irradiated.  No  signal  attributable  to  the  1,2-dideuteriovinyl  radical 
was  observed  in  a  similar  experiment  employing  acetylene-d^J  this  result  is  consist- 
ent with  an  inverting  radical  in  which  all  the  lines  are  broadened  beyond  detection. 
Hence  Fessenden  and  Schuler  conclude  that  the  vinyl  radical  is  nonlinear  and  that 
the  a-proton  is  rapidly  inverting.   They  feel  that  the  hybridization  on  the  a- 
carbon  atom  is  intermediate  between  sp2  and  sp  and  probably  lies  close  to  the  former. 
By  estimating  the  combined  line  widths  of  the  broadened  lines  to  be  on  the  order  of 
fifty  gauss,  they  place  limits  of  3xl0~8  and  3xl0~10  sec  on  the  lifetimes  of  the 
individual  configurations.  Assuming  a  classical  model  and  a  normal  prcexponential 
factor  of  about  1013  sec"1  (i.e.,  AS'^0) ,  this  range  of  lifetimes  corresponds  to  a 
barrier  to  inversion  of  about  2  kcal/mole;  If  the  lifetime  of  a  single  configuration 


-231- 
is  limited  by  quantum-mechanical  tunnelling,7  this  figure  represents  a  lower  limit 
on  the  estimated  barrier  height. 

The  electron-beam  irradiation  of  a  2.5-mole-$  solution  of  allene  in  liquid 
ethane  at  101°K  gave  rise  to  an  ESR  signal  which  was  interpreted  as  a  superposition 
of  the  spectra  of  the  ethyl,  2-propenyl  ( a-methylvinyl) ,  3-propynyl,  and  allyl 
radicals.4  The  sixteen  lines  attributed  to  the  2-propenyl  radical  can  be  grouped 
into  four  equally  intense,  overlapping  Ii2i2i±   quartets j  this  pattern  is  consistent 
with  a  slowly  ( on  the  ESR  time  scale)  or  noninverting  nonlinear  radical.  The  lower- 
ing of  the  inversion  rate  when  a  methyl  group  is  substituted  for  the  a-proton  of  the 
vinyl  radical  is  consistent  with  a  tunnelling  mechanism  for  inversion.  The  hyperfine 
splittings  for  the  2-propenyl  radical  are  19.48,  32.92,  and  57*89  gauss^  the  latter 
two  values  are  assigned  to  the  two  g-protons  since  their  sum  is  similar  to  the  sum 
of  the  two  g-proton  hyperfine  splittings  in  the  vinyl  radical, 

Cochran,  Adrian,  and  Bowers  generated  the  vinyl  radical  at  4.2°K  by  ultraviolet 
irradiation  (2537  A0)  of  a  solid  argon  matrix  containing  1$  hydrogen  iodide  and  9$ 
acetylene.8  They  observed  a  complex  unsymmetrical  spectrum  but  were  able  to  assign 
the  eight-line  gross  pattern  to  three  nonequivalent  protons.   Other  workers  observed 
a  similar  spectrum  in  the  solid  phase  at  77°K  after  atomic  hydrogen  generated  in  a 
silent  electrical  discharge  had  been  allowed  to  react  with  acetylene  at  20°C  in  the 
gas  phase  or  at  -196°C  in  the  solid  phase.9  The  complexity  of  the  spectrum  pre- 
sumably arises  from  the  failure  of  anisotropic  dipolar  interactions  to  average  to 
zero  through  rapid  thermal  tumbling.  The  hyperfine  splittings  due  to  the  three 
protons  are  15.7*  3^.2,  and  68.5  gauss °910   these  values  are  in  good  agreement  with 
the  values  given  above  for  the  a-proton  and  the  sum  of  the  g-proton  hyperfine 
splittings  for  the  vinyl  radical  in  the  liquid  phase.   In  the  solid  phase  at  4.2°K 
the  inversion  is  apparently  slowed  or  frozen  so  that  two  distinct  g-proton  couplings 
are  observed. 

When  a  mixture  of  hydrogen  iodide  and  acetylene -d2  was  photolyzed  under  similar 
conditions,  the  ESR  spectrum  consisted  of  two  groups  of  lines  separated  by  approx- 
imately 64  gauss.  The  spectrum  became  weaker  as  the  temperature  was  raised  to  32°K, 
but  no  new  lines  appeared.  Addition  of  a  hydrogen  atom  to  the  acetylene  molecule 
can  give  rise  to  either  or  both  isomeric  vinyl  radicals  I   and  II.  The  absence  of  any 

H^  R^    ^D" 

C=C  ^=C. 

I  II 

lines  with  splittings  of  3^  or  16  gauss  was  rationalized8  to  require  that  only  one 
species,  the  isomer  (I  or  II)  whose  g-proton  gives  rise  to  the  ok   gauss  splitting, 
is  present  in  detectable  amounts.  If  inversion  of  the  vinyl  radical  is  a  tunnelling 
phenomenon,7  the  rate  of  inversion  should  be  relatively  temperature  independent, 
since  there  is  no  classical  activation  energy  requirement.  An  inversion  process 
that  is  rapid  near  100°K  in  the  liquid  phase  but  slow  enough  to  maintain  configuration 
at  4.2°K  in  the  solid  phase  is  not  consistent  with  a  tunnelling  mechanism  unless  the 
transition  from  liquid  to  solid  at  4.2°K  is  accompanied  by  an  additional  stabilization 
of  the  predominant  radical  relative  to  the  transition  state  for  inversion  of  at  least 
150  cal/mole. 

A  single  predominant  isomer  might  be  the  kinetically-determined  product  of 
specific  cis  or  trans  addition  of  the  hydrogen  atom,  or  it  might  be  the  thermodynam- 
ically  more  stable  radical.  Cochran,  Adrian,  and  Bowers  feel  that  neither  of  these 
explanations  is  entirely  satisfactory.  They  believe  that  the  addition  of  a  hydrogen 
atom  to  a  molecule  of  acetylene  is  exothermic  and  should  give  rise  to  a  vibrationally 
excited  vinyl  radical,  which  should  invert  rapidly  before  being  cooled  to  4.2°Kj 
they  do  not  consider  the  possibility  that  the  rigid  matrix  may  hinder  the  atomic 
motion  necessary  for  inversion,  or  that  it  may  provide  a  lattice -relaxat ion  mechanism 
for  rapidly  deactivating  an  initially-formed  excited  cis  or  trans  radical.   If  the 
isomer  ratio  i/ll  is  thermodynamically  controlled,  the  predominance  of  a  single 
isomer  even  at  32°K  would  require  at  least  a  15$  difference  in  the  total  zero-point 
vibrational  energies  of  I  and  II  in  the  absence  of  a  significant  difference  in  the 
steric  interactions  of  the  matrix  with  the  cis  and  trans  radicals.8  The  fact  that 


170 

objections  can  be  raised  to  both  kinetic  and  thermodynamic  control  of  the  i/ll  ratio 
suggests  that  the  original  interpretation  of  the  ESR  spectrum  as  indicative  of  only- 
one  isomer  may  need  reexamination. 

Several  attempts  have  been  made  to  determine  the  geometry  at  the  a-carbon  atom 
of  the  vinyl  radical  by  comparing  theoretically  calculated  spin  densities  for 
different  configurations  with  the  observed  hyperfine  splitting  constants.  Using  a 
calculation  based  on  the  hyperconjugation  theory  of  Coulson  and  Crawford,11  Dixon 
calculated  that  the  carbon-carbon-a-hydrogen  bond  angle  0  (see  structure  III)  lies 
between  120°  and  I5O0.12  Adrian  and  Karplus  narrowed  this  range  to  I3O0  -  15O0  by 

using  a  valence  bond  calculation  and  by  assuming 
that  the  68  gauss  splitting  corresponds  to  the 
trans -f3 -proton.13  However,  the  agreement  of 
calculated  and  experimental  hyperfine  splittings 
for  any  value  of  0  may  be  fortuitous  in  this 
case)  Strauss  and  Fraenkel  have  shown  that  the 
same  type  of  valence  bond  method  gives  a  poor 
fit  between  calculated  and  observed  C13  hyper- 
fine splittings.14  An  extended  Hilickel  molecular 
orbital  treatment  by  Petersen  gives  the  best  fit  between  theory  and  experiment  for 
0  .  i^0.15 

RADICAL  ADDITIONS  TO  SUBSTITUTED  ACETYLENES 

Alkenyl  radicals  have  generally  been  assumed  to  be  intermediates  in  free  radical 
additions  to  alkynes.  Radical  additions  to  mono-  and  disubstituted  acetylenes  have 
been  briefly  discussed  in  several  reviews  concerned  chiefly  with  radical  additions 
to  olefins.16"18  The  pertinence  of  individual  stereochemical  studies  to  the  question 
of  the  configurational  stability  of  alkenyl  radicals  must  be  carefully  considered, 
since  conclusive  evidence  for  a  hemolytic  mechanism  is  often  lacking.  Even  in 
reactions  for  which  a  radical  mechanism  is  established,  free  alkenyl  radicals  may 
not  be  involved.   In  several  cases,  interpretation  of  the  results  is  complicated  by 
incomplete  product  studies,  ambiguous  stereochemical  assignments,  and  the  absence  of 
proof  that  product  ratios  are  kinetic ally  controlled. 

A  large  number  of  radical  additions  to  substituted  acetylenes,  most  of  which 
are  initiated  by  peroxides  or  UV- irradiation,  yield  products  corresponding  to 
preferential  trans  addition  to  the  triple  bond.   Included  in  this  group  are  the 
liquid -phase  addition  of  hydrogen  bromide  to  propyne,19*20  the  bromination  of  2- 
butynoic  acid,21  the  addition  of  trichlorosilane  to  mono-  and  dialkylacetylenes,22'23 
the  addition  of  alkyl-  and  arylthiols  to  ethoxyacetylene,24  propiolic  acid,25'26  and 
arylacetylenes,25*27*28  the  reaction  of  thiolacetic  acid  with  phenylacetylene27  and 
1-hexyne,29  the  addition  of  triphenyltin  hydride  to  phenylacetylene,30  and  the 
addition  of  ditin  and  diarsine  compounds  to  hexafluoro--2-butyne.31  On  the  other 
hand,  the  brominations  of  several  terminal  and  internal  alkynes,32'35  the  reaction 
of  hydrogen  bromide  with  1-bromoalkynes,36  and  the  addition  of  perfluoroalkyl 
iodides  to  acetylenic  compounds37*38  appear  to  proceed  predominantly  by  cis  addition 
to  the  triple  bond. 

Skell  and  Allen  have  rationalized  the  exclusive  trans  UV-catalyzed  hydro- 
bromination  of  propyne  in  the  liquid  phase  at  -780  to  -60u  by  1)  a  configurationally 
stable  cis  alkenyl  radical,  2)  a  bridged  bromine  radical,  or  5)  bromine  atom 
addition  to  an  initially -formed  alkyne -hydrogen  bromide  complex,  closely  followed  by 
hydrogen  atom  transfer.19'20  Oswald,  Griesbaum,  Hudson,  and  Bregman  have  suggested 
that  rapid  equilibration  of  cis  and  trans  alkenyl  radicals  and  stereoselective 
hydrogen  atom  transfer  to  the  less -hindered  cis  radical  can  account  for  the  apparent 
trans  addition  of  alkyl-  and  arylthiols  to  phenylacetylene.27  Bergel'son  has  con- 
cluded that  in  the  bromination  of  mono-  and  disubstituted  acetylenes,  the  reaction 
stereochemistry  is  determined  mostly  by  the  relative  thermodynamic  stabilities  of 
the  cis  and  trans  radicals,  which  in  turn  depend  on  the  steric  repulsions  between 
the  substituents  on  the  radicals.32'35  In  contrast  to  this  result,  Truce,  Klein, 
and  Kruse  have  found  that  the  steric  requirements  of  the  mesityl  groups  in  the 
transition  state  of  the  product -forming  step  in  the  addition  of  2-thiomesitylene  to 
mesitylacetylene  are  not  sufficiently  large  to  reverse  the  normal  trans  stereochemistry 
of  thiol  addition.28 


-233- 

RADICAL  REACTIONS  OF  ALKENYLMETALLIC  COMPOUNDS 

Beletskaya,  Karpov,  and  Reutov  have  claimed  a  radical  mechanism  for  the  reaction 
of  S-chlorovinylmercuric  chloride  with  iodine  in  benzene  and  in  carbon  tetrachloride 
to  form  l-iodo-2-chloroethylene„39  They  have  also  studied  the  stereochemistry  of 
the  j3-bromostyrenes  (IV)  produced  by  the  reaction  of  bromine  with  cis  and  trans -2- 
phenylethenylmercuric  bromides  (V)  ,40  In  methanol  in  the  presence  of  added  bromide 

j#CH=CH-HgBr  +  Br2  - — >   jfcH=CHBr  +  HgBr2 

V  IV 

ion,  the  reaction  proceeds  with  high  (88.5%-cisj  93 .5% -trans )  retention  of  stereo- 
chemistry at  the  double  bond.   In  carbon  tetrachloride  the  reaction  with  either 
cis  or  trans  V  gives  nearly  equal  amounts  of  cis  and  trans-(3-bromostyrenes ,  as 
evidenced  by  the  index  of  refraction  of  the  distilled  isomer  mixture j  the 
equilibrium  c is/trans  ratio  is  3>1»      ^.f   the  reaction  does  indeed  proceed  by  a 
radical  mechanism  and  if  the  product -forming  step  is  the  bromination  of  a  B-styrenyl 
radical,  the  similarity  of  the  product  ratios  from  cis  and  trans  organomercurials 
indicates  that  the  p-styrenyl  radical  either  is  linear  or,  if  nonlinear,  is  rapidly 
inverting.   In  the  latter  case,  the  loss  of  original  stereochemistry  in  the  presence 
of  as  good  a  radical  trap  as  molecular  bromine  requires  that  the  isomerization  be 
exceedingly  rapid. 

Gloc.kling  has  shown  that  the  thermal  decomposition  of  l-( 2-methylpropenyl) 
silver(I)   in  ethanol  at  -20°  probably  proceeds  by  a  radical  chain  mechanism. 41 
Whitesides  and  Casey  have  studied  the  thermal  decompositions  of  cis  and  trans-1- 
propenyl-and  2~but~2-enylcopper(  I)   and  silver(  I)  and  the  corresponding  tri-n- 
butylphosphine  complexes  in  ether  at  ambient  temperature j„42  The  sole  organic  pro- 
ducts are  2,^-hexadienes  and  3>^-dimethyl-2,4-hexadienes,  formed  in  high  yields  with 
greater  than  99%  retention  of  configuration  at  the  double  bond.   For  example,  eis-1- 
propenyl(tri-n-butylphosphine)  silver (I)  (VI),  prepared  from  1-propenyllithium  of  97% 
cis  stereochemistry,  gave  cis  ,cls-2,Wnexadiene  ( VII)  in  95%  yield  and  cis , trans- 
2,4~hexadiene  (VIII)  in  4."5%~yield,  corresponding  to  a  total  yield  of  cis  propenyl 
groups  of  97%  and  a  stereospecificity  of  100%,  based  on  the  isomeric  purity  of  the 
starting  organolithium. 


K    CH 


3 


Li^     /CH3   1'  {fU3?P£U   nBu3PA*       CH, 
\„_n^       ether, -78°   .-  J  °\„__/  * 


CT    CH3 
II     I 
V.C  C.  _  VII 

Cx 


A 


;c=c'    -^;     f   '  — -*     jc=c. 

IT  ^H     2°   dl0xane  H^     "H  CH3^    ^H 

VI  /C— C\    /CH< 

H       C=C^ 

VIII 

If  free  alkenyl  radicals  are  formed  in  these  reactions,  product  formation 
probably  occurs  by  coupling  of  an  alkenyl  radical  with  a  molecule  of  unde composed 
organometallic ,  present  in  ca  0.1  M  or  lower  concentration  under  the  reaction  con- 
ditions.  In  order  to  account  for  the  high  stereospecificity  of  diene  formation, 
coupling  must  proceed  at  least  102  times  faster  than  inversion  of  the  radical: 

k( coupling) x [alkenyl  radical]x[alkenylmetallic]  ^> 

102  x  k(  inver si on) x[ alkenyl  radical] 

If  the  configurational  lifetime  of  the  alkenyl  radical  is  approximately  that  of  the 
vinyl  radical,  on  the  order  of  10~8  or  10~10  r.ef  ;4  then  k(  coupling)  must  be  greater 
than  1011  l-mo.le"1sec~1.   Whitesides  and  Casey  argue  that  this  minimum  value  for 


.OX)i 


k( coupling)  is  unreasonably  large  and  conclude  that  the  reaction  does  not  proceed 
with  the  formation  of  free  alkenyl  radicals.  They  also  rule  out  geminate  combination 
of  radicals  within  a  solvent  cage  because  they  feel  that  the  100%  efficiency  of  cage 
combination  necessary  to  explain  the  product  yields  is  unlikely.  They  do  not  exclude 
the  possibility  that  alkenyl  radicals  are  formed  which  are  configurationally 
stabilized  by  jt-complexation  to  the  metal  atoms. 

GENERATION  OF  ALKENYL  RADICALS  BY  PERESTER  DECOMPOSITION 

Bartlett  has  proposed  a  nonconcerted  radical  mechanism,  in  which  the  initial 
step  is  cleavage  of  an  oxygen-oxygen  bond,  for  the  thermal  decomposition  of  tert- 
butyl  percinnamate.43  Kampmeier  and  Fantazier  decomposed  tert -butyl  cis  and  trans - 
a,P«dimethylpercinnamates  (IX  and  X)  in  cumene  at  110°. 44  The  observed  products  are 
consistent  with  the  following  reaction  scheme: 


X 


■OH 
?=C^     XI 

CH3   CH3 


XIII 


c=c^ 

CH3   CH^ 


XIV 


0     C-0-O-tBu 

CH3   CH3 


IX 


-£■»  tBuO»  +   ^=C^ 

CH3   CH3 


CH3COCH3 

+ 
( CH3)  3COH 


C02  +   ,£=Q 

CHq 


3    CH3 


or 


—   1>C-CH3 
CHo 


C=C^         -A>   tBuO  +   ^=C^ 
df3   g-0-O-tBu  CH3   C-0* 


K   /CH3 

co2  +   c=c 


X 


I 


^C=C^     XII 
CH3   C-OH 


A    J^3 

^C=C^    XV 
CH3   H 


Dicumyl  is  also  produced  in  significant  ( cis -43.4%;  trans -65.5$)  yields,   cis  and 
trans  -a  ,g -Dimethylcinnamic  acids  (XI  and  XII)  are  obtained  in  low  (1-2%)  yields  and 
have  the  same  stereochemistry  as  the  parent  perester.   3 A-Dimethylcoumarin  (XIII) 
is  formed  exclusively  from  the  cis  perester  in  12.8%  yield.   If  the  acids  and  the 
coumarin  arise  through  the  acyloxy  radicals  rather  than  by  a  minor  heterolytic 
pathway,  these  observations  indicate  that  the  cis  and  trans  acyloxy  radicals  do  not 
inter convert. 

In  separate  experiments,  Kampmeier  and  Fantazier  showed  that  trans  perester 
(X)  recovered  after  partial  decomposition  contains  no  cis  perester  ( IX) ,  that  cis 
and  trans -a ,8 -dimethylcinnamic  acids  added  to  decomposing  trans  perester  are 
recovered  nearly  quantitatively  and  unisomerized,  and  that  cis  and  trans -2 -phenyl - 
2-butenes  (XIV  and  XV)  are  not  isomerized  under  the  reaction  conditions.  The  sum  of 
the  yields  of  cis  and  trans -2-phenyl-2-butenes  accounts  for  most  of  the  alkenyl 
radicals  formed,  as  measured  by  the  evolution  of  carbon  dioxide,  so  that  hydrogen 
abstraction  is  the  major  reaction  of  xhe  alkenyl  radicals.  Tne  cis  and  trans 


peresters  give  mixtures  of  olefins  XIV  and  XV  with  the  same  c is/ trans  ratios,  1.1- 
1.2.  At  100°  the  equilibrium  cis/trans  ratio  is  approximately  four."45  The  common 
kinetically-controlled  isomer  ratio  again  supports  the  intermediacy  of  a  linear  or 
rapidly-inverting  nonlinear  radical. 

If  nonlinear  alkenyl  radicals  are  stereospecifically  generated  from  the  cis 
and  trans  peresters,  it  might  be  possible  to  trap  the  radicals  with  an  efficient 
scavenger  before  they  can  isomerize.  The  first-order  rate  constant  for  decomposition 
of  the  trans  perester  is  not  affected  by  the  addition  of  an  equimolar  quantity  of  2- 
thiomesitylene,  even  though  the  yield  of  trans  -a  ,ft -d  imethylcinnamic  acid  increases 
from  1.9$  to  36. 3$. 46  This  result  requires  separate  product-  and  rate -determining 
steps  and  suggests  that  Bartlett's  one-bond  cleavage  mechanism43  is  also  operating  in 
this  case.  As  the  concentration  of  added  2=thiomesitylene  is  increased,  the  cis/ 
trans  ratio  of  2=phenyl-2-butenes  from  both  cis  and  trans  peresters  increases  toward 
the  equilibrium  value.  The  failure  of  the  thiol  to  trap  a  mixture  of  radicals 
richer  in  the  trans  isomer  from  the  trans  perester  does  not  rule  out  the  stereo- 
specific  formation  of  nonlinear  radicals,  since  isomerization  of  the  olefinic  pro- 
ducts by  the  added  thiol  would  obscure  this  observation.  A  precedent  for  such 
olefin  isomerization  is  found  in  the  work  of  Oswald,  Griesbaum,  Hudson,  and  Bregman,27 
who  found  that  ethanethiol  and  thiophenol  catalyze  the  isomerization  of  cis-ft-thio- 
phenylstyrene  and  cis-ft-thioethylstyrene  to  the  more  stable  trans  isomers. 

Singer  and  Kong  have  investigated  the  thermal  decomposition  of  the  tert -butyl 
peresters  of  cis  and  trans -a-methylc innamic  acids  (XVI)  and  of  cis  and  trans -q- 
phenylcinnamic  acids  (XVII)  in  several  solvents  at  110°. 47>48  in  a  single  solvent, 

jZ5CH=C(CH3)C03tBu  -£»  j0CH=CHCH3 

XVI  XVIII 

jZ$CH=C(j#)C03tBu       -^>    J0CH=CH0 

XVII  XIX 

the  cis  and  trans  peresters  give  the  same  kinetically-controlled  (nonequilibrium) 
cis/trans  ratios  of  1-propenylbenzenes  (XVIII)  or  stilbenes  (XIX)  °9   the  cis/trans 
olefin  ratios  increase  as  the  solvent  is  changed  from  toluene  to  cyclohexene  to 
cumene.  The  authors  interpret  these  results  in  terms  of  stereoselective  capture  of 
rapidly-equilibrating  cis  and  trans  alkenyl  radicals  by  the  solvent.  Cumene  has 
the  largest  steric  requirement  in  the  transition  state  for  hydrogen  atom  transfer, 
so  that  this  solvent  shows  the  greatest  preference  for  hydrogen  transfer  trans  to 
the  a-phenyl  group  and  yields  an  olefin  mixture  containing  the  greatest  proportion 
of  the  cis  isomer.  Similar  reasoning  has  been  used  to  explain  the  variation  of  the 
cis/trans  decalin  ratios  observed  when  cis  or  trans-9-carbo-terjt-butylperoxydecalin 
is  decomposed  in  different  solvents.49 

GENERATION  OF  ALKENYL  RADICALS  BY  THE  HUNSDIECKER  REACTION 

The  brominative  decarboxylation  of  the  silver  salts  of  carboxylic  acids  is 
regarded  as  a  reliable  method  for  generating  alkyl  radicals.50  Berman  and  Price 
have  reported  that  the  bromination  of  silver  cis  and  trans -cinnamates  in  refluxing 
carbon  tetrachloride  gives  trans  -ft  -br omostyrene  (10$  and  17.5$,  respectively),  1,1,2- 
tribromo-2-phenylethane  ( 25%  and  35$),  and  the  c innamic  acid  of  unchanged  stereo- 
chemistry (12$  and  8$).51  Since  at  equilibrium  g-bromostyrene  contains  an  appreciable 
amount  of  the  cis  isomer,40  the  exclusive  formation  of  the  trans  isomer  from  both 
cis  and  trans  silver  salts  indicates  that  the  free  energy  of  the  transition  state 
for  bromination  of  the  cis  radical  is  significantly  (>3.2  kcal/mole  if  1$  of  cis -ft - 
bromostyrene  could  have  been  detected)  higher  than  that  of  the  trans  radical.   If 
the  thermodynamic  stabilities  of  the  cis  and  trans  radicals  are  comparable,  this 
explanation  corresponds  to  a  high  stereoselectivity  for  bromine  transfer  to  the 
intermediate  vinyl  radical.   Conclusions  drawn  from  this  study  must  be  regarded  as 
tentative,  since  large  percentages  of  starting  silver  salts  were  not  accounted  for. 

The  Hunsdiecker  reactions  of  silver  cis  and  trans -a-phenylcinnamates  give 
different  cis/trans  ratios  of  a-bromostilbenes.5i=;  Studies  of  product  stabilities 
under  the   reaction  conditions  are  not  reported. 


GENERATION  OF  ALKENYL  RADICALS  BY  THERMAL  DECOMPOSITION  OF  DIACYL  PEROXIDES 

Simamura,  Tokumaru,  and  Yui  decomposed  cis  and  trans-die innamyl  peroxides 
carbon  tetrachloride  and  in  bromotrichloromethane,53  The  stereochemistry  of  the  g- 
halostyrene  products  is  given  below; 

isomer  of  c is/trans  ratio  of  product 

peroxide        solvent       jfcH=CHC.l      jfciI=€HBr 

cis  CC14  18/82 

BrCCl3  27/73 

trans         CCI4  19/81 

BrCCl3  14/86 

Separate  experiments  showed  that  the  g_-halostyrenes  are  not  isomerized  under  the 
reaction  conditions „ 

A  common  linear  intermediate  radical,  is  ruled  out  as  the  sole  source  of  pri 
ducts  by  the  different  cis /trans  ratios  from  cis  and  trans  peroxides  in  bromot] 
chloromethane o  The  partial  retention  of  configuration  in  bromotrichloromethane  bui 
not  in  carbon  tetrachloride  is  consistent  with  a  competition  between  inversion  of 
stereospecifically -generated  alkenyl  radicals  and  trapping  of  the  radicals  by 
solvent  °9   the  smaller  activation  energy  requirement  for  breaking  a  bromine -cart- 
bond  in  the  chain  transfer  step  is  reflected  in  the  greater  ease  with  which  bromo- 
trichloromethane adds  to  olefins o54  In  carbon  tetrachloride  the  equilibration  of 
the  isomeric  alkenyl  radicals  is  complete j  if  the  equilibrium  constant  for  the 
radical  isomerization  is  near  unity ,  the  1:4  c is/trans  product  isomer  ratio  suggests 
that  the  phenyl  group  may  sterically  hinder  halogen  transfer  to  the  cis  radical, 

BIBLIOGRAPHY 

1.  J„  Hine,  "Physical  Organic  Chemistry,"  McGraw-Hill  Book  Co. ,  Inc.,  New  York, 
1962,  pi  473 . 

2.  Do  Yo  Curtin  and  J.  W.  Crump,  J.  Am.  Chem.  Soc.,  80,  1922  (1958). 

3.  R.  L„  Letsinger,  J.  Am.  Chem.  Soc.,  72,  4842  (1950}". 

4.  R.  W.  Fessenden  and  R.  H.  Schuler,  J.  Chem.  Fnys.,  39,  2147  ( I963) . 

5.  R.  A.  Holroyd  and  R.  W„  Fessenden,  J0  Phys.  Chem.,  gf,  274  5  (1963). 

6.  R.  A.  Holroyd  and  G„  W.  Klein,  Intern.  J.  Appl.  Radiation  Isotopes,  15,  633 
(1964). 

7.  Both  Fessenden  and  Schuler4  and  Cochran,  Adrian,  and  Bowers8  suggest  the 
possibility  of  tunnelling.  Experimental  methods  for  detecting  tunnelling  of 
protons  have  been  reviewed  by  M.  T„  Link,  MIT  Organic  Seminar,  I  Semester  1965- 
1966,  p.  I96. 

8.  E.  L.  Cochran,  F.  J.  Adrian,  and  V.  A.  Bowers,  J.  Chem.  Phys.,  40,  213  (1964). 
9o  N.  Ya„  Buben,  R.  V.  Kolesnikova,  N,  L,  Kuznetsova,  and  V.  I.  Trofimov,  Bull. 

Acad.  Sci.  USSR,  Div.  Chem.  Sci.,  I99O  (1964). 

10.  These  values  are  revised  values  attributed  to  Adrian  and  coworkers  and  quoted 
by  Fessenden.4 

11.  C.  A.  Coulson  and  V.  A.  Crawford,  J.  Chem.  Soc,  2052  (1953). 

12.  W.  T.  Dixon,  Mol.  Phys.,  9,  201  (I965). 

13.  F.  J.  Adrian  and  M.  Karplus,  J.  Chem.  Phys.,  4l,  56  (1964). 

14.  H.  L.  Strauss  and  G.  K.  Fraenkel,  J.  Chem.  Fnys.,  35.,  I738  (I96I). 
15c  H.  Petersen,  Ph.D.  Thesis,  University  of  Illinois,  I967,  pp.  82-92. 
I60  B.  A.  Bohm  and  P.  I.  At ell,  Chem.  Rev.,  62,  599  (1962). 

17.  C.  Walling  and  E.  S.  Huyser,  Org.  Reactions,  13_,  91  ( 1963) . 

18.  F„  W.  Stacey  and  J.  F.  Harris,  Org.  Reactions,  15,  150  (1963). 

19.  P.  S.  Skell  and  R.  G.  Allen,  J.  Am.  Chem.  Soc. ,80,  5997  (1958) . 

20.  P.  S.  Skell  and  R.  G.  Allen,  J.  Am.  Chem.  Soc,  W,   1559  (1964). 

21.  A.  Pinner,  Ber. ,  28,  1884  (1895). 


077 


22.   R.  A.  Benkeser  and  R.  A,  Hickner,  J.  Am.  Chem.  Soc.,  80,  5298  (1958). 
2.3„  R„  A«  Benkeser,  M.  L.  Burrous,  L.  E.  Nelson,  and  J.  V.  Swisher,  J.  Am.  Chem. 
Soc,  85,  4385  (1961). 

24.  H.  J.  Alkema  and  J.  F.  Arens,  Ree.  Trav.  Chim. ,  79,  1257  (i960). 

25.  Y.  Liu  and  H.  Wang,  Hua  Hsueh  Hsueh  Pao,  31,  ^51("l965)  >  Chem.  Abstr. ,  64, 
17391  (1966}. 

26.  L.  N.  Owen  and  M.  J.  S.  Sultanbawa,  J.  Chem.  Soc.,  3109  (1949). 

27.  A.  A.  Oswald,  K.  Griesbaum,  B.  E.  Hudson,  and  J.  M.  Bregman,  J.  Am.  Chem. 
Soc,  86,  2877  (1964). 

28.  W.  E.  Truce,  H.  G.  Klein,  and  R.  B.  Kruse,  J.  Am.  Chem.  Soc,  83,  4636  (1965)  . 

29.  J.  A.  Kampmeier  and  G.  Chen,  J.  Am.  Chem.  Soc.,  87,  2608  (1965)7 

30.  R.  F.  Fulton,  Ph.D.  Thesis,  Purdue  University,  196*0,  cited  by  H.  G.  Kuivila, 

in  F.  G.  A.  Stone  and  R.  West,  "Advances  in  Organometallic  Chemistry,"  Academic 
Press,  New  York,  1964,  Vol.  1,  p.  65. 
31o  W.  R.  Cullen,  D.  S.  Dawson,  and  G.  E.  Styan,  J.  Organometal.  Chem.,  3,  4o6 
(1965)o 

32.  L.  D„  Bergel'son,  Bull.  Acad.  Sci.  USSR,  Div.  Chem.  Sci.,  995  (i960). 

33.  I.  N.  Nazarov  and  L.  D.  Bergel'son,  J.  Gen.  Chem.  USSR,  27,  l6l4  (1957). 

34.  I.  N.  Nazarov  and  L.  D.  Bergel'son,  Bull.  Acad.  Sci.  USSR,  Div.  Chem.  Sci., 
826  (i960). 

35<>  1"  N.  Nazarov  and  L.  D.  Bergel'son,  Bull.  Acad.  Sci.  USSR,  Div.  Chem.  Sci., 

834  (i960). 
36.  L.  D„  Bergel'son,  Bull.  Acad.  Sci.  USSR,  Div.  Chem.  Sci.,  1145  (1960). 
37o  K.  Leedham  and  R.  N.  Haszeldine,  J.  Chem.  Soc,  1634  (1954). 
38.  N.  0.  Brace,  J.  Org.  Chem.,  27,  449  (1962). 
39-  lo  P.  Beletskaya,  V.  I.  Karpov,  and  0.  A.  Reutov,  "Proceedings  of  the  Conference 

on  the  Problems  of  the  Application  of  Correlation  Equations  in  Organic  Chemistry," 

Tartu ^  1962,  p.  355 9   cited  in  reference  40. 

40.  I.  P.  Beletskaya,  V.  I.  Karpov,  and  0.  A.  Reutov,  Bull.  Acad.  Sci.  USSR,  Div. 
Chem.  Sci.,  1615  (1964). 

41.  F.  Glockling,  J.  Chem.  Soc,  716  (1955)  i   3^40  (1956). 

42.  G.  M.  Whitesides  and  C.  P.  Casey,  J.  Am.  Chem.  Soc,  88,  454l  (1966). 

43.  P.  D.  Bartlett  and  R.  R.  Hiatt,  J.  Am.  Chem.  Soc,  80,  1398  (I958). 

44.  J.  A.  Kampmeier  and  R.  M.  Fantazier,  J.  Am.  Chem.  Soc,  88,  1959  (1966). 

45.  D.  J.  Cram  and  M.  R.  V.  Sahyun,  J.  Am.  Chem.  Soc,  8j?,  1257  (I963). 

46.  R.  M.  Fantazier  and  J.  A.  Kampmeier,  J.  Am.  Chem.  Soc.,  88,  5219  (I966). 

47.  L.  A.  Singer  and  N.  P.  Kong,  Tetrahedron  Letters,  2089  (1966). 

48.  L.  A.  Singer  and  N.  P.  Kong,  J.  Am.  Chem.  Soc.,  88,  5213  (1966). 

49.  P.  D.  Bartlett,  R.  E.  Pincock,  J.  H.  Rolston,  W.  G.  Schindel,  and  L.  A.  Singer, 
J.  Am.  Chem.  Soc,  87,  2590  (1965). 

50.  D.  E.  Applequist  and  A.  H.  Peterson,  J.  Am.  Chem.  Soc,  82,  2372  (i960). 
51o  J.  D.  Berman  and  C.  C.  Price,  J.  Org.  Chem.,  23,  102  (I95H). 

52.  E0  L.  Sukman,  Dissertation  Abstr.,  1£,  1211  (1958). 

53.  0.  Simamura,  K.  Tokumaru,  and  H.  Yui,  Tetrahedron  Letters,  5l4l  (1966). 

54.  C.  Walling,  "Free  Radicals  in  Solution,"  John  Wiley  and  Sons,  Inc.,  New  York, 
1957,  Po  250. 


-238- 

ISOMERIZATION  OF  ORGANIC  THIOCYMATES  TO  ISOTHIOCYANATES 

Reported  by  Joseph  C.  Stickler  March  2,  I.967 

INTRODUCTION 

The  observation  that  organic  thiocyanates  isomerize  to  isothiocyanates 
(equation  l)  was  made  about  a  century  ago.1'2  However,  not  until  this  decade 

R-I-C5NI  ^  R-N=C=^  (1) 

have  serious  investigations  been  initiated  to  obtain  a  mechanistic  description 
of  these  isomerizations.  Experimental  evidence  in  these  investigations  has  in- 
dicated that  they  may  proceed  by  at  least  four  pathways  depending  upon  the  struc- 
ture of  the  organic  substrate  and  reaction  conditions.  If  the  organic  moiety 
contains  an  allylic  double  bond,  the  reaction  may  proceed  by  a  shift  in  that 
bond  and  a  change  in  the  site  of  anion  attachment.3  6  Also,  conditions  can 
be  controlled  so  that  either  a  direct  displacement7  or  ionization  mechanism8 
is  responsible  for  the  isomerization  of  saturated  systems.   The  isomerization 
of  thiocyanates  upon  irradiation  is  believed  to  proceed  by  a  radical  pathway.9 
An  interesting  outgrowth  of  these  isomerization  studies  was  the  information  they 
yielded  on  the  ambident  reactivity  of  the  thiocyanate  anion.7 

ALLYLIC  SYSTEMS 

The  isomerization  of  allylic  thiocyanates  to  isothiocyanates  has  been  known 
since  I875.1"'2  The  fact  that  the  site  of  attachment  to  the  organic  substrate 
changed  in  allylic  isomerization  was  substantiated  when  it  was  demonstrated  that 
crotyl  thiocyanate  upon  heating  to  l40°  yielded  quantitatively  a-methylallyl 
isothiocyanate.4  The  proposed  mechanism  for  this  rearrangement  contains  a  six- 
membered-ring  transition  state,  which  may  be  represented  by  contributing  struc- 
tures I  through  IV  or  alternatively  by  V.  This  is  a  characterization  similar 
to  that  attributed  to  the  rearrangement  of  a,a-dimethylallyl chloride  by  Young, 
Winstein  and  Goering.10 


S— c  =N  S=  C-=JJ  ( SCN.) 

I  II  III 


c    c  cr   -c 

(SCN)6  L  .A 


Sr^1 


IV 


This  mechanism  is  supported  by  several  experimental  considerations.   The 
entropy  of  activation  for  the  allyl  and  2-methylallyl  thiocyanate  rearrangement 
are  -3*k  +  1  and  -8.7  +  1  cal./deg.-mole6  respectively  which  is  indicative  of 
a  high  degree  of  order  in  the  transition  state  relative  to  the  ground  state. 
These  values  are  the  same  sign  and  order  of  magnitude  as  several  reactions  be- 
lieved to  proceed  through  a  six-membered-ring  transition  state  such  as  the 
Claisen  rearrangement  of  allyl  vinyl  ether.11  These  rearrangements  obey  first-order 


-239- 
kineticso  The  rate  of  conversion  in  the  allylic  case  is  faster  than  in  the  sat- 
urated case.  For  instance,  allyl  thiocyanate6  in  toluene  at  86.4°  isomerizes 
at  a  rate  of  3.81  X  10~4  sec"1  as  compared  to  I.87  X  10~6  sec"1  for  benzhydryl 
thiocyanate  at  90° 0°  in  benzene,8  The  smallness  of  charge  separation  in  the 
transition  state  is  indicated  by  the  fact  that  neither  changes  in  solvents  towards 
greater  polarity  nor  added  salts  increased  the  rate  of  reaction  significantly. 
The  changes  in  rate  appear  especially  small  when  compared  to  the  data  compiled 
by  Streitwieser12  for  the  solvolysis  of  other  allylic  systems  believed  to  occur 
by  way  of  an  ionization  route.  For  instance,  for  the  solvolysis  of  Y/Y-dimethylallyl 


chloride,  the  ratio  k.rrar     ,,    . .. 

5  Op  ethanol  ethanol 

the  corresponding  thiocyanate , 5  k 


=  3300,  but  for  the  isomerization  of 

3«   Thus  the  con- 


acetonitrile  cyclohexane 
tributions  of  structures  III  and  IV  are  believed  to  be  small  compared  to  I  and 
II,  with  V  perhaps  being  the  best  representation. 

Although  the  isomerization  of  a  pure  isothiocyanate  to  thiocyanate  has  not 
been  clearly  established,  it  is  known  that  an  equilibrium  state  is  obtained  in 
the  isomerization  of  allylic  thiocyanates.5'*13'14  From  the  results  of  these 
equilibria  in  Table  I,  several  points  of  interest  have  been  made. 


TABLE  I5 
Equilibrium  thiocyanate/isothiocyanate  at 
100°  in  different  media. 


%   of  thiocyanate 

!  at 

equilibrium 

Compound 

Pure 

cyclo- 
hexane 

aceto- 

nitrile 

Cone, 
m/l 

Benzhydryl 

- 

<1 

2-3 

0.1-0.75 

allyl 

<i 

<1 

9-11 

10~3-10_1 

Y-methylallyl 

11 

5 

27 

10"  3 

y  /y-dimethylallyl 

ko 

18 

50 

10"3 

Analysis  by  U. V.  and  I.R.  spectrophotometry. 


As  the  polarity  of  the  medium  is  increased,  the  percent  of  thiocyanate  increas- 
es, which  is  consistent  with  the  fact  that  the  thiocyanate  has  a  higher  dipole 
moment  and  therefore  would  be  more  stable  in  a  polar  solvent.  If  the  rearrange- 
ment causes  the  organic  substrate  to  lose  stability  by  loss  of  hyperconjugative 
stabilization  of  a  y -methyl  group,  then  there  is  a  greater  amount  of  the  thio- 
cyanate in  equilibrium.   Contrasting  with  this  result  is  the  fact  that  the  more 
methyl  substitution  at  the  gamma  position,  the  faster  the  rate  of  conversion 
to  the  isothiocyanate  as  seen  from  Table  II.  Apparently  the  greater  the  ability 
of  the  organic  substrate  to  stabilize  a  positive  charge  the  more  contribution 
the  ionic  structures  III  and  IV  make  to  the  transition  state.  In  what  can  appar- 
ently be  considered  an  extreme  case  of  the  equilibrium  favoring  the  thiocyanate, 
Y-phenylallylthiocyanate  does  not  rearrange  at  all  with  an  allylic  shift  to 
a-phenylallyl  isothiocyanate  which  would  no  longer  have  the  allylic  double  bond 
conjugated  to  the  phenyl  group,  but  instead  isomerizes  to  "y-phenylallyl  iso- 
thiocyanate.15 The  mechanism  believed  to  be  responsible  is  the  same  as  that 
involved  in  the  isomerization  of  saturated  thiocyanates  and  will  be  discussed  in 
the  next  section. 


-2  Ro- 
table ii5 

First-order  rate  coefficients  at  60°  for  the 
isomeric  rearrangement  of  allylic  thiocyanates; 
RR'C^HCHaSCN  )    RR'C(NCS)C=CH2 

A 


Compound 

105  k/w)/          "i\ 
60u(sec   x) 

W6O0 

R             R> 

Ab                Cb 

H              H 

CH3C       H 
CH3         CH3 

1.8            3.3 
2?              31 
270d           96d 

0.5 
0.9 

3 

a 


Determined  by  U. V.  spectrophotometry. 

A  and  C  stand  for  acetonitrile  and  cyclohexane  solvents 
respectively. 

trans 

Extrapolated  from  lower  temperature  data. 


SATURATED  SYSTEMS 

Many  saturated  thiocyanates  are  known  to  isomerize  to  the  corresponding 
isothiocyanates.  The  mechanism  believed  to  apply  to  most  of  these  systems  is 
a  unimolecular  ionization  to  an  ion  pair,  which  can  return  to  starting  material 
or  go  to  the  isothiocyanate.   Before  considering  the  details  of  the  above  mech- 
anistic description,  the  experimental  evidence  which  excludes  other  pathways 
of  reasonable  a  priori  probability  will  be  presented. 

The  likelihood  of  a  concerted  pathway  in  which  there  would  be  some  bond 
making  of  the  nitrogen  to  the  a-carbon  before  complete  bond  breaking  of  the  sul- 
fur to  a-carbon  bond  seems  small.   This  route  would  involve  a  four -member ed- 
ring  transition  state  as  shown  in  equation  (2)  and  would  be  analogous  to  the 
Chapman  rearrangement  of  imino  esters.16  Such  a  transition  state  would  probably 


R-S-C=N 


->   R-N=C=S 


(2) 


have  a  higher  energy  of  activation  than  the  allylic  isomerization;  however,  the 
energies  of  activation  for  saturated  and  allylic  systems  are  quite  similar. 
Also,  one  might  expect  retention  of  optical  purity,  but  extensive  racemization 
is  found  when  an  optically  active  substrate  is  employed.17  The  argument  that 
the  geometry  of  the  C-S-C=N  group  is  prohibitive  to  the  concerted  mechanism  has 
been  made.8 

A  possible  second  route  is  represented  by  equation  (3).   This  route  is 
discredited  by  the  fact  that  the  kinetics  are  first  order. 


-2kl~ 
R 

2RSCN >     M  — )-  RNCS  +  RSCN  ( 3) 

RSC  SCN9 


Again,  one  would  expect  retention  of  optical  purity  in  the  product,  which  is 
not  found. 

A  third  pathway  might  he  an  a-elimination  to  a  carhene  and  thiocyanic  acid, 
and  then  recombination.  However,  attempts  to  identify  the  acid  by  spectroscopy 
and  to  trap  a  carbene  failed.6 

A  fourth  possible  mechanism  would  be  homolytic  cleavage  followed  by  simple 
recombination  (equation  k)   which  would  yield  the  necessary  first-order  kinetics, 
but  the  more  likely  chain  process  would  result  in  a  more  complicated  expression. 

R-SCN y    R.  +   *SCN  ^   R-NCS  (k) 

Besides  the  solvent  and  salt  effect  which  are  more  indicative  of  an  ionization 
mechanism,  no  side  products  were  isolated  which  would  be  expected  to  accompany 
a  radical  pathway.  However,  it  appears  that  conditions  can  be  controlled  so  as 
to  lead  to  an  isomerization  of  benzyl  and  benzhydryl  thiocyanates  to  the  iso- 
thiocyanates  by  a  radical  pathway.  Mazzucato9  recently  reported  that  when  ben- 
zyl thiocyanate  is  irradiated  with  a  low  pressure  mercury  arc,  using  the  2537  & 
line,  at  a  concentration  of  about  10  %  in  n-hexane,  at  room  temperature  and 
with  the  exclusion  of  oxygen  from  the  system,  a  photoequilibrium  with  the  iso- 
thiocyanate  was  established.  The  photostationary  equilibrium  mixture  was  about 
70%  thiocyanate  and  3C$>   isothiocyanate.  On  the  basis  of  detection  of  the  fluor- 
escence emission  of  the  benzyl  radical  in  the  ^60-530niM-  region  starting  from 
both  the  thio-  and  isothiocyanate  in  low  temperature  photolyses,  a  radical  mech- 
anism was  suggested. 

Having  disposed  of  the  above  pathways,  evidence  supporting  the  ionization 
mechanism  will  be  presented.   Perhaps  the  most  persuasive  evidence  is  the  com- 
parison of  this  system  to  the  many  similar  systems  studied  by  Winstein,18  Smith,19 
Goering,20  Darwish21  and  others.   The  proposed  mechanism  is  an  ionization  to  form 
an  ion  pair  intermediate  through  which  isomerization  occurs.   The  evidence 
overwhelmingly  points  toward  this  pathway  and  the  purpose  here  will  be  to  see 
how  well  a  detailed  picture  of  this  mechanism  is  fulfilled.  As  a  point  of  ref- 
erence, equations  (5)  and  (6)  representing  the  detailed  process  of  ionization- 
dissociation  described  by  Winstein  and  coworkers22  will  be  employed.   Several 

ionization                     dissociation 
I  )  II )  III  VlV  (5) 


external 
intimate  or 

or  solvent- 

internal  separated         dissociated 

ion  pair  ion  pair         ions 


li    +  -      ^  + 

RX  -^ R  X      ^ R 

k  ,  k 


x"    , R   +  X  (6) 


k 


-1         "-2  "-3 

types  of  organic  substrates  have  been  studied,  but  because  of  its  suitability  for 
extensive  studies  attention  was  mainly  focused  on  the  benzhydryl  substrate. 


olio 

~<—-TC—  - 


As  previously  stated,  the  reaction  obeys  first-order  kinetics.  The  energy 
and  entropy  of  activation  for  the  isomer ization  of  4,4' -dimethylbenzhydryl  thio- 
cyanate  in  acetonitriie  are  20.67  kcal./mole  and  -10.0  cal./mole-deg. ,  respectively.8 

In  a  study  of  solvent  effects  on  the  rate  of  isomer: ization ,  Fava  and  co- 
workers demonstrated  that  an  increase  in  polarity  accelerated  the  rate,  which 
is  compatible  with  the  ionization  mechanism.  For  example,  at  90°  for  the  benz- 
hydryl  case,  the  rates  in  methyl  ethyl  ketone,  acetonitriie  and  dimethylform- 
amide  relative  to  benzene  were  10,  150  and  280.  This  solvent  sensitivity  is 
similar  in  order  of  magnitude  to  that  of  other  reactions  believed  to  occur  by 
a  unimolecular  ionization.  For  instance,  the  ratio  of  rates  in  acetonitriie  and 
benzene  are  120  for  the  rearrangement  of  camphene  hydrochloride  and  2b  for  the 
rearrangement  of  9-decalyl  perbenzoate.8 

The  effect  of  structure  on  reactivity  revealed  that  with  increasing  stab- 
ility of  the  carbonium  ion  formed  by  ionization,  the  rate  of  isomerization  also 
increased.  A  qualitative  study  of  the  rates  of  the  series  n-butyl,  sec -butyl, 
and  t-butyl  thiocyanates  indicated  that  rate  increased  in  going  toward  the  more 
highly  branched  substrate.6  In  a  study  of  para- substituted  benzhydryl  thiocy- 
anates, Fava  and  coworkers  demonstrated  that  electron  releasing  substituents 
facilitated  the  rate,  while  electron  withdrawing  groups  hindered  the  rate. 
Using  Brown's  u  values23  a  Hammett  plot  of  log  k/ko  vs.cr  resulted  in  a  linear 
relationship  with  p  =  -3.40  in  acetonitriie  at  70°.  This  result  compares  well 
with  p  =  -^.05  for  the  solvolysis  of  benzhydryl  chloride  in  ethanoL.8  Hence 
structurax  modifications  of  the  organic  moiety  which  aid  the  stabilization  of 
a  positive  charge  increases  the  rate  of  isomerization. 

Salts  added  to  the  reaction  medium  also  increased  the  rate  of  isomerization, 
thus  supporting  the  ionization  mechanism.  Fava  and  coworkers8  investigated 
in  particular  the  effects  of  two  salts,  sodium  perchlorate  and  sodium  thiocyanate. 
The  investigation  was  carried  out  in  two  solvents,  methyl  ethyl  ketone  and  acet- 
onitriie }   the  effect  being  greater  in  the  less  polar  methyl  ethyl  ketone.  The 
results  fit  well  the  Winstein  equation24( 7)  for  "normal"  salt 


k  =  ko  (1  +  b  [salt]) 


(7) 


effects.  The  b  values  are  summarized  in  Table  III.   The  possibility  that 

TABLE  III 

Salt  effects  on  isomerization  of  benzhydryl  thiocyanate  at  90° 

in  different  media. 


b  values  from  Winstein  equation  (7) 

Salt 

Methyl  ethyl 
ketone 

Acetonitriie 

bMeEtC0 

bMeCN 

NaSCN 
NaC104 

16.7 
11.7 

4.27 
2.93 

3.92 
3.99 

a  concurrent  direct  displacement  by  the  nitrogen  end  of  the  ionic  thiocyanate 
might  account  for  its  greater  effect  as  compared  with  the  sodium  perchlorate 
was  not  completely  ruled  out.  However,  the  ratio  of  the  b  values  in  the  two 
solvents  was  independent  of  the  salt,  which  is  expected  if  the  effect  is  a 
specific  salt  effect.  Also,  the  degree  of  specificity  of  the  different  salts 
is  not  uncommon.8 

Although  the  solvent,  structural,  and  the  salt  effects  described  above  lead 
strongly  to  the  conclusion  that  an  ionization  pathway  is  operative,  they  do  not 
give  a  detailed  picture  of  the  mechanism.  Fava  and  coworkers  further  particularized 


the  mechanism  "by  performing  exchange  experiments  with  sulfur  labeled  sodium 
thiocyanate.8'25  The  main  portion  of  these  exchange  experiments  was  carried  out 
on  the  4,4'-dimethylbenzhydryl  thiocyanate  where  it  was  shown  that  direct  dis- 
placement by  either  the  sulfur  or  nitrogen  end  of  the  ionic  thiocyanate  on  the 
organic  substrate  is  negligible  (equations  8  and  9)  >  since  the  kinetics  indicated 


NCS*  +  RSCN 


*SCN   +  RSCN 


,NCS*R  +  SCN 


*SCER  +  SCN 


(8) 
(9) 


essentially  unimolecular  exchange  with  ionic  thiocyanate.   Therefore  all  radio- 
activity must  enter  the  organic  substrate  by  way  of  the  ionization  route.  The 
first  conclusion  drawn  from  these  experiments  was  that  since  the  amount  of  label, 
on  the  ionic  thiocyanate  always  remained  far  greater  than  that  on  the  organic 
thiocyanate  or  isothiocyanate,  the  organic  thiocyanate  could  not  have  become 
completely  dissociated  upon  ionization  or  the  labeled  ionic  thiocyanate  would 
have  become  equilibrated  with  the  organic  thiocyanate.   Thus  it  was  proposed 
that  the  isomerization  occurred  through  an  ion  pair. 

The  salt  effect  on  both  the  rate  of  isomerization  and  exchange  of  kfhl 
-dimethylbenzhydryl  thiocyanate  in  acetonitrile  at  25°  were  examined.  The 
results  are  shown  in  Table  IV.  As  can  be  seen  from  Table  IV  the  rate  of  exchange 


TABLE  IV 


Salt  effects  on  rates  of  exchange  and  isomerization  of 

h, h '-dimethylbenzhydryl  thiocyanate  in  acetonitrile 

b  values  from  Winstein  equation  (7) 

Isomerization 

Exchange 

NaC104 
NaSCN 

2 

k 

.38 

a  25.0  +  0.1° 
b  0o2  +  0.1° 


is  more  greatly  effected  by  added  salts  than  the  isomerization  rate.  The  author 
interpreted  this  to  mean  that  a  more  advanced  degree  of  ionization  (i.e.  greater 
charge  separation)  was  involved  in  exchange  than  in  isomerization.  Further 
indication  that  exchange  occurs  at  a  more  advanced  stage  of  ionization  is  the 
p  value  of  -4.5  for  exchange  at  70°  in  acetonitrile.26  This  is  more  negative 
than  the  p  =  -J.kQ  measured  under  the  same  conditions  for  isomerization  indicating 
a  more  highly  polar  transition  state.  Whether  exchange  might  occur  at  the  ex- 
ternal ion  pair  or  upon  dissociation  or  both  cannot  be  determined  from  this 
data.  However,  the  authors  have  speculated  that  for  the  highly  polar  acetonitrile 
and  the  benzhydryl  moiety,  the  dissociated  carbonium  ion  seems  the  more  probable 
intermediate  for  the  first -order  exchange  process.  Assuming  that  isomerization 
occurs  at  the  internal  ion  pair  stage,  and  that  if  in  some  instances  the  ionization 
proceeds  further  and  exchange  takes  place  with  labeled  ionic  thiocyanate,  then 
these  exchanged  species  will  proceed  back  to  the  internal  ion  pair  and  then 
partition  between  thiocyanate  and  isothiocyanate,  more  information  may  be  ob- 
tained. The  ratio  of  incorporated  label  in  thiocyanate  and  isothiocyanate  was 
made  and  it  was  determined  that  the  ion  pair  returned  to  covalency  with  the  sulfur 
end  five  times  more  often  than  the  nitrogen  end.   With  this  information  a  lower 


-2kk- 

limit  may  be  set  on  the  rate  of  ionization  by  the  expression  (10)  where  kj. 
is  the  rate  of  ionization  and  k.  the  isomerization  rate.   The  determination 

1  s 


ki  =  k_,  +  k 


N 


is 


6k, 


is 


(10) 


kN 


of  the  rate  of  ionization  in  this  manner  assumes  that  the  ion-pair  stage  at  which 
isomerization  occurs  is  not  preceded  by  another  ion  pair  which  returns  exclusive- 
ly to  the  thiocyanate.  The  stereochemical  evidence  indicated  that  racemization 
of  p-chlorobenzhydryl  thiocyanate  and  isomerization  takes  place  at  the  same 
intermediate,  thus  supporting  the  assumption  that  no  ion  pair  precedes  the  ion 
pair  stage  responsible  for  isomerization. 17>2G     This  method  is  similar  to  that 
employed  by  Goering  and  coworkers  in  their  study  of  the  jD-nitrobenzoate-carbonyl 
-018  system*20  the  main  exception  being  that  partition  between  the  isotopic 
oxygens  is  not  dependent  upon  reaction  conditions  as  is  the  ratio  k  /k... 

From  the  ratio  of  the  rate  of  exchange  to  the  rate  of  ionization  (equation 
11) ,  it  is  possible  to  calculate  an  upper  limit  to  the  percent  of  dissociation 
that  occurs.   Using  4,4' -dimethylbenzhydryl  thiocyanate  in  acetonitrile  at  0° 
with  0.01M  NaSCN, 


(11) 


the  upper  limit  is  calculated  to  be  5°3/°»The  authors  make  the  following  sum- 
mary to  the  result:   Out  of  100  internal  ion  pairs  formed,  about  5  undergo  furth- 
er ionization  and  95  return  to  the  covalent  state.  From  the  partition  data  mentioned 
above  it  was  concluded  that  of  the  95  returning  to  covalency,  about  79  return 
to  thiocyanate  and  16  to  the  isothiocyanate.  Also  concluded  was  that  under 
conditions  favoring  ionization  to  a  smaller  extent  (less  polar  solvent  or  less 
stable  carbonium  ion) ,  return  to  covalency  from  the  internal  ion  pair  predom- 
inates even  more  completely. 

An  outgrowth  of  these  exchange  experiments  employed  to  particularize  the 
isomerization  mechanism  was  an  extensive  mechanistic  study  on  the  isotopic 
exchange  between  substituted  benzhydryl  thiocyanates  and  ionic  thiocyanates. 
Fava  and  coworkers26 >srT   concluded  that  with  strongly  electron  attracting  sub- 
stituents  a  bimolecular  pathway  prevailed  and  with  strongly  electron  donating 
substituents  a  unimolecular  process  was  obtained.  However ,  with  substituents 
having  intermediate  electron-donating'  abilities ,  concurrent  bimolecular  and 
unimolecular  mechanisms  were  operative. 

The  isomerization  of  cyclopropylcarbinyl  thiocyanate  (VI)  is  of  special 
interest,  since  this  is  the  only  example  known  in  which  the  saturated  organic 
moiety  undergoes  a  rearrangement  of  the  carbon  skeleton  during  isomerization. 
Spurlock  and  Newallis28  have  reported  that  the  isomerization  of  cyclopropyl- 
carbinyl thiocyanate  at  155°C  in  acetonitrile  yielded  compounds  VII  through  XI, 
the  respective  percent  yield  indicated  below  each  structure.   The  kinetics  were 


SCN 


+ 


'SCN 


VI 


VII   (12) 


sew 


VIII  (6) 


+ 


ix  (75) 


KCS 


x  (5) 


'NCS 


XI   (2) 


^NCS 


-2li-5- 
first  order  and  the  rate  of  isomerization  at  15^.7°  was  2,55  X  10  5  sec  -1  in 
dimethylf ormamide .  The  authors  have  reported  that  control  experiments  indicated 
each  of  the  products  VII -XI  was  stable  to  the  reaction  conditions.  In  the  pres- 
ence of  added  potassium  thiocyanate  in  dimethylformamide  solutions,  the  isomer- 
ization rate  was  increased,  "but  the  products  and  their  distribution  were  un- 
affected.  Since  direct  displacement  by  the  ionic  thiocyanate  would  favor  the 
cyclopropylcarbinyl  isothiocyanate,  the  authors  concluded  that  the  acceleration 
of  the  rate  could  be  attributed  mainly  to  "normal"  salt  effects. 

The  isomerization  of  methyl  thiocyanate  is  somewhat  unexpected  since  exper- 
iments have  indicated  that  isomerization  of  n-butyl  thiocyanate  is  negligible. 
However,  at  131° ,  the  boiling  point  of  methyl  thiocyanate,  it  isomerized  and  was 
aided  by  dissolved  salts.29  Although  the  isomerization  proceeds  in  the  neat  methyl 
thiocyanate,  it  is  prevented  in  non-polar  solvents.6  The  kinetics,  which  were 
measured  in  the  pure  state,  were  found  to  fit  a  first-order  rate  law;  and  the  rate 
of  isomerization  at  120°  was  1.02  X  10  7  sec  1.  From  the  above  rate  data,  an 
energy  and  entropy  of  activation  were  calculated  to  be  kl   kcal./mole  and  31 
cal./mole-deg. ,  respectively.30  The  concerted  four-membered  ring,30  bimolec- 
ular6  and  ionization10  mechanisms  have  all  been  suggested  on  the  basis  of  this 
sparse  experimental  data.  A  more  systematic  and  comprehensive  study  of  this 
system  seems  to  be  necessary  for  classification  of  the  mechanism  involved. 

AMBIDENT  NATURE  OF  THE  THIOCYANATE  ANION 

One  of  the  interesting  aspects  of  the  isomerization  studies  is  the  infor- 
mation they  yield  about  the  ambident  reactivity  of  the  thiocyanate  anion.   So 
far  we  have  seen  that  there  are  three  general  mechanisms  by  which  these  isomer- 
izations  proceed;  by  allylic  rearrangement,  ionization  and  radical  pathways. 
In  order  to  provide  further  scope  to  the  study  of  this  ambident  reactivity,  Fava 
and  coworkers7  were  able  to  control  the  conditions  of  the  isomerization  so  as 
to  provide  the  isomerization  with  another  quite  general  pathway.   By  using  the 
benzyl  substrate  which  'undergoes  direct  substitution  easily  and  maintaining  a 
substantial  amount  of  ionic  thiocyanate  in  the  reaction  media,  the  isomerization 
was  found  to  proceed  mainly  by  an  S  .2  mechanism.  Radioactive  thiocyanate  was 
employed  and  the  rates  of  exchange,  which  proceeded  either  by  a  direct  displace- 
ment of  the  S  or  N  end  of  ionic  thiocyanate  as  well  as  the  rate  of  isomerization, 
direct  displacement  by  the  N  end,  were  measured.   Since  the  rate  of  isomerization 
was  on  the  order  of  10  6  and  the  isotopic  exchange  rate  on  the  order  of  10"4,  the 
rate  of  exchange  is  approximately  equal  to  the  reactivity  of  the  sulfur  end. 
The  rate  of  isomerization  is  equal  to  the  reactivity  of  the  nitrogen  end  of  the 
thiocyanate  anion.   Therefore  the  ratio  of  k  /k„   equals  the  reactivity  ratio 
kq/k  .  In  acetonitrile  and  methyl  ethyl  ketone  at  temperatures  ranging  from 
70-100°,  the  values  of  k  /k  were  on  the  order  of  102  to  103  for  the  benzyl  sub- 
strate. For  instance,  at  70°  the  ratios  of  k  /k  are  1000  and  725  for  the  methyl 
ethyl  ketone  and  acetonitrile  respectively;  while  values  of  65O  and  460  are  ob- 
tained at  100°  for  the  same  two  solvents.  On  the  other  hand,  in  reactions  stud- 
ied by  Cannell  and  Taft31  in  which  carbonium  ions  were  generated  in  aqueous  sol- 
utions independently  from  the  thiocyanate  ions,  k  /k  values  ranging  from  2  to  9 
were  determined  from  product  ratios.  In  Fava's  work   concerned  with  the  iso- 
merization by  the  ionization  mechanism,  a  value  for  k  /k  of  5  was  obtained  for 
the  isomerization  of  4,4' -dimethylbenzhydryl  thiocyanate  in  acetonitrile,  which 
is  on  the  same  order  of  magnitude  as  values  obtained  from  the  Cannell  and  Taft  work. 

As  indicated  from  the  above  results,  this  system  provides  a  good  method  for 
determining  how  various  factors  affect  the  relative  reactivities  of  the  two  ends 
of  the  thiocyanate  anion.  Further  investigation  into  the  solvent  effects  on  this 
ambident  anion  might  prove  useful  in  ascertaining  the  role  solvation  plays  in 
determining  relative  reactivities.  Also  a  systematic  study  of  the  thiocyanate 
radical,  in  cases  where  the  radical  pathway  is  operative,  could  reveal  more  about 
the  nature  of  the  relative  reactivity  of  the  two  ends  of  the  thiocyanate  radical. 
An  extensive  review  related  to  nucleophilic  ambifunctional  reactivity  by  Gompper32 
has  been  published. 


BIBLIOGRAPHY 

1.  G.  Gerlick,  Ann.,  17_8,  80  (1875)- 

2.  0.  Billeter,  Ber. ,  8,  462  (1875)- 

3.  0.  Billeter,  Helv.  Chira.  Acta,  8,  337  (1925). 

4.  0.  Mumm  and  H.  Richter,  Ber.,  7£B,  843  (1940). 

5.  A.  Iliceto,  A.  Fava,  and  U.  Mazzucato,  Tetrahedron  Letters,  No.  11 ,  27 

(I960), 

6.  P.  A.  S.  Smith  and  D.  W.  Emerson,  J.  Am.  Chem.  Soc,  82,  3076  (i960). 

7.  A.  Fava,  A.  Iliceto,  and  S.  Bresadola,  ibid. ,  87,  4791  (1965). 

8.  A.  Iliceto,  A.  Fava,  U.  Mazzucato,  and  0.  Rossetto,  ibid. ,  83,  2729  (1961). 

9.  U.  Mazzucato,  G.  Beggiato,  and  G.  Favaro,  Tetrahedron  Letters,  5455  (1966). 

10.  W.  G.  Young,  S.  Winstein,  and  H.  L.  Goering,  J.  Am.  Chem.  Soc,  73,  1958 

(1951). 

11.  A.  A.  Frost  and  R.  G.  Pearson,  "Kinetics  and  Mechanism,"  John  Wiley  and 

Sons,  Inc.,  New  York,  N.Y. ,  1953,  PP-  104-105. 

12.  A.  Streitvieser,  Jr.,  Chem.  Rev.,  j>6,  65O  (1956). 

13.  A. Iliceto  and  G.  Gaggia,  Gazz.  chim.  ital. ,  90,  262  (i960). 

14.  A.  Iliceto,  A.  Fava,  U.  Mazzucato,  and  P.  Radici,  ibid. ,  90,  919  (i960). 

15.  E.  Bergmann,  J.  Chem.  Soc,  1361  (1935). 

16.  A.  W.  Chapman,  ibid.,  127,  1992  (1925). 

17.  A.  Fava,  U.  Tonellato,  and  L.  Congiu,  Tetrahedron  Letters,  1657  (1965). 

18.  (a)   S.  Winstein  and  A.  H.  Fainberg,  J.  Am.  Chem.  Soc,  80,  459  (1958); 

(b)  S.  Winstein  and  J.  S.  Gall,  Tetrahedron  Letters,  No.  2,  31  (i960) ; 

(c)  S.  Winstein,  J.  S.  Gall,  M.  Hojo,  and  S.  Smith,  J.  Am.  Chem.  Soc, 
82,  1010  (i960);  (d)  S.  Winstein,  M.  Hojo,  and  S.  Smith,  Tetrahedron  Let- 
ters,Nc22,  12  (i960);  (e)  S.  Winstein,  A.  Ledwith,  and  M.  Hojo,  ibid.  , 
341  (I96I). 

19.  S.  G.  Smith,  ibid. ,  979  (1962). 

20.  (a)  H.  L.  Goering  and  J.  T.  Doi,  J.  Am.  Chem.  Soc,  82,  5850  (i960)  ;  (b) 

H.  L.  Goering  and  J.  F.  Levy,  Tetrahedron  Letters,  6^+4  (1961)5  (c)  H.  L. 
Goering  and  J.  F.  Levy,  J.  Am.  Chem.  Soc,  84,  3853  (1962);  (d)  H.  L.  Goering, 
R.  G.  Briody,  and  J.  F.  Levy,  ibid. ,  8£,  3059  (1965). 

21.  (a)  D.  Darwish  and  R.  McLaren,  Tetrahedron  Letters,  1231  (1962);  (b)  D.  Dar- 

wish and  E.  A.  Preston,  ibid. ,  113  (1964);  (c)  D.  Darwish  and  J.  Noreyko, 
Can.  J.  Chem. ,  4£,  1366  (1965) • 

22.  S.  Winstein,  E.  Clippinger,  A.  H.  Fainberg,  R.  Heck,  and  G.  C.  Robinson, 

J.  Am.  Chem.  Soc,  78,  328  (1956). 

23.  Y.  Okamoto  and  H.  C.  Brown,  J.  Org.  Chem.,  22,  485  (1957). 

24.  S.  Winstein,  E.  Clippinger,  A.  H.  Fainberg  and  G.  C.  Robinson,  J.  Am.  Chem. 

Soc,  36,   2597  (195*0- 

25.  A.  Fava,  A.  Iliceto,  A.  Ceccon,  and  P.  Koch,  ibid. ,  87,  1045  (1965). 

26.  A.  Ceccon,  I.  Papa,  and  A.  Fava,  ibid. ,  88,  4643  ( i960) . 

27.  A.  Fava,  A.  Iliceto,  and  A.  Ceccon,  Tetrahedron  Letters,  685  (1963). 

28.  L.  A.  Spurlock  and  P.  E.  Newallis,  ibid. ,  303  (1966). 

29.  J.  Gillis,  Rec  trav.  chim.,  jg,  330  (1920). 

30.  C.  N.  R.  Rao  and  S.  N.  Balasubrahmanyam,  Chemistry  and  Industry,  625  (i960) . 

31.  L.  G.  Cannell  and  R.  W.  Taft,  Abstracts  of  the  129th  A.  C.  S.  Meeting, 

46N  ( 1956) . 

32.  R.  Gompper,  Angew.  Chem.  Int.  Ed.,  J5,  560  (1964). 


nli  <~r 


THE  PHYSIOLOGICALLY  ACTIVE  CONSTITUENTS  OF  MARIHUANA 

Reported  by  Donald  C.  Schlegel  March  6,  1967 

INTRODUCTION 

Cannabis  sativa  Linn.,  more  commonly  known  as  hemp,  is  a  plant  which  grows 
wild  or  is  cultivated  throughout  much  of  the  world.  The  greenish  resin  known  as 
marihuana  (The  Americas),  hashish  (Middle  East),  or  charas  (Far  East)  can  be 
extracted  from  the  flowers  and  seeds  of  the  female  plant  and  the  leaves  of  both 
sexes.1  The  oil  is  composed  of  at  least  25  recognized  compounds,  the  major  con- 
stituents being  caryophyllene  ^5.7$,  (3-humulene  16$,  cc-selinene  8.6$,  f3-farnesene 
5.1$,  a-bergamotene  5$,  ^-phellandrene  2.7$. 2  In  smaller  content,  1.2$  or  less 
depending  on  the  geographical  origin,  are  two  tetrahydrocannabinols,  the  active 
constituents  which  make  marihuana  a  psychotomimetically  active  drug.3  The  search 
for  these  compounds,  their  structural  elucidation  and  most  recently  the  resynthesis 
of  their  optically  inactive  forms  are  the  topics  of  this  seminar. 

PHYSIOLOGICAL  EFFECTS 

Marihuana  is  a  psychotomimetically  active  drug.  It  is  known  to  have  a  profound 
effect  on  the  central  nervous  system.   It  can  be  taken  into  the  body  by  smoking  or 
direct  ingestion.  The  effects  have  been  "described  as  a  feeling  of  well  being 
alternating  with  depression,  distortions  of  time  and  space,  and  double  consciousness."1 
Illusion  and  fanciful  hallucinations  are  common  and  disorientation  and  delirium  may 
follow.  An  increased  sensitivity  of  the  eyes  to  light  is  also  observed.4 

STRUCTURE  ELUCIDATION 

In  the  1870's,  chemists  initially  began  to  investigate  marihuana.5  They  met 
with  little  success,  however,  and  it  wasn't  until  the  1890' s  and  the  work  of  Wood, 
Spivey  and  Easterfield5  that  much  was  known  about  the  supposedly  active  constituent 
or  constituents  of  marihuana.  At  that  time  they  isolated  a  material  C2;LH2602,  I,  from 
the  higher  boiling  fraction  of  the  marihuana  resin.  Treatment  of  this  compound  with 
cold  fuming  nitric  acid  yielded  a  yello\/  trinitro  derivative  C21II23N308,  II.  Further 
oxidation  of  this  product  with  hot  fuming  nitric  acid  gave  among  other  products  a 
mononitro  derivative  CnHuNC^,  III.  The  nitro  group  was  reduced  to  the  amine, 
diazotized,  treated  with  potassium  iodide  to  form  the  iodide  and  then  reduced  with 
sodium  amalgum  to  form  Ci:iH1202,  IV,  called  cannabinolactone.  'This  material  was 
reoxidized  with  basic  permanganate  to  give  a  CnH1004  compound,  V,  cannabinolactonic 
acid.  Reduction  of  this  gave  a  dibasic  acid  C1:LH1204,  VI.  Thirty  years  later  Cahn6 
further  oxidized  cannabinolactonic  acid  with  hot  dilute  nitric  acid  and  obtained 
trimellitic  acid,  VII,  thus  showing  the  aromatic  ring  substitution.  He  also 
demonstrated  that  the  acid  formed  a  lactone  with  a  tertiary  alcohol  by  collecting 
acetone  and  3-hydroxy-Wnethylbenzoic  acid,  VIII,  on  potassium  hydroxide  cleavage 
of  hydroxy-cannabinolactone. 

Cahn  applied  the  name  cannabincl,  earlier  used  to  denote  the  high  boiling 
fraction  of  marihuana,  to  the  compound  C2iH2602,  I.  The  material  formed  both  mono- 
methyl  ether  and  monoacetyl  derivatives.  Measurement  of  the  critical  oxidation 
potential  gave  a  value,  Ec  =  0.995  ±  0.10,  characteristic  of  substituted  phenols  but 
not  carbinols.  Recovery  of  •►hexanoic  acid  on  potassium  permanganate  oxidation 
indicated  the  presence  of  a  straight  chain  moiety.  Empirical  formula  considerations 
suggested  that  the  chain  was  a  n-pentyl  side  chain  on  the  phenolic  ring  with  the 
sixth  carbon  atom  coming  from  oxidation  of  the  ring.  Assembling  the  information 
gave  IX  as  the  basic  structure  for  cannabinol.   It  was  not  until  the  general 
structure  of  cannabidiol  was  determined  by  groups  led  by  Todd  and  Adams  that  the 
complete  structure  of  cannabinol  was  known. 


5^11 


Fuming 
HNO^ 


cold    CH3 


^     ^g 


II 


cf/hi     r^ 


KMnQ4 
KOH 


CH3- 


dilute  HNO3 

I    hot 


IX 


Cannabidiol,  another  material  found  by  Adams  and  Todd  in  the  high  boiling 
fraction  of  marihuana  was  seen  immediately  to  possess  a  molecular  formula  C2iH3002 
quite  similar  to  cannabinol.7  It  had  two  acidic  hydroxy  1  groups  which  were  readily 
acetylated,  Ultraviolet  spectral  comparison  showed  the  phenolic  hydroxyls  to  be  in 
a  resorcinol  arrangement.  Under  mild  hydrogenation  conditions  it  rapidly  consumed 
two  moles  of  hydrogen  without  great  change  in  the  UV  chromophore,  indicating  that 
the  phenolic  ring  remained  unchanged.8  Treatment  of  cannabidiol  with  pyridine 
hydrochloride  yielded  conclusive  proof  of  its  general  structure  when  p-cymene  and 
olivetol  were  formed.9 


olio. 


Py.'HClj 


5HH 


C5H11 
CH2  CH3 

The  point  of  attachment  of  the  two  rings  was  determined  by  oxidizing  tetrahydro- 
cannabidiol.  The  resulting  menthane  carboxylic  acid  was  isolated  as  its  anilide/ 

JHa  CH3 


KMnQ4 


■> 


acetone 
r.t. 


/\0H- 
CH3  CH3 


CkH 


5nll 


AH 

CH3  CH3 


Assuming  cannabinol  to  be  similar  to  cannabidiol,  the  last  data  suggested  the 
correct  structure  for  cannabinol.  Synthesis  of  this  structure  confirmed  the  identity 
of  cannabinol.10'11 

CH3 


5H11 

Recently,  using  modern  analytical  techniques  Mechoulam  and  Shvo12  have  completed 
the  structural  elucidation  of  cannabidiol.  Earlier  work  by  Adams  et  al.13  on  place- 
ment of  the  two  non-aromatic  double  bonds  had  demonstrated  that  a  terminal  double 
bond  was  present  by  recovering  formaldehyde  on  ozonization  of  cannabidiol.   Ultra- 
violet studies  of  model  compounds  further  showed  that  the  second  double  bond  was 
neither  conjugated  with  the  aromatic  ring  nor  with  the  terminal  isopropylene  double 
bond.13  Consequently,  the  remaining  double  bond  could  be  located  only  in  positions 
A5,  A6,  or  A1.   On  NMR  analysis,  Mechoulam  and  Shvo  observed  3  olefinic  protons  and 


5Hn 


5H11 


5H11 


AJ 


2  olefinic  methyl  groups  in  the  spectrum  of  cannabidiol.  These  data  eliminated  the 
A5  isomer  as  a  possibility.   The  NMR  spectrum  further  showed  the  C3  proton  at  t6.15. 
Normally  a  benzyl  proton  appears  around  T7tl3.14  This  datum  indicated  the  proximity 
of  another  deshielding  force,  the  double  bond.  The  C3  proton  appeared  as  a  doublet 
(J  =  11)  which  is  not  appreciably  coupled  with  the  C2  proton.   If  cannabidiol  has 
the  double  bond  in  the  A1  position,  then  02  would  be  at  an  angle  of  approximately 
850  and  consequently  not  appreciably  coupled  to  C3.   On  hydrogenation  the  product, 
tetrahydrocannabidiol,  shows  the  C3  proton  at  t7.^0.   If  the  A6  isomer  were  the 
correct  structure,  little  change  in  the  C3  proton  position  would  be  expected.  A 
final  experiment  was  a  selective  epoxidation  on  the  ring  double  bond  of  cannabidiol 


-250- 
bisdinitrobenzoate  with  perbenzoic  acid.  The  NMR  spectrum  of  the  product  showed  no 
change  in  the  position  of  the  methyl  group  or  protons  surrounding  the  side  chain 
double  bond  but  the  Ci,  C2  and  C3  groups  did  change.   In  particular  the  C3  proton 
was  now  observed  as  a  doublet  at  t6.70,  again  indicating  the  proximity  of  the 
epoxide  linkage.  From  the  data  it  was  concluded  that  cannabidiol  is  the  A1 
isomer.12 

Evidence  leading  to  the  stereochemistry  at  carbon  3  and  k ,   the  final  structural 
unknown,  initially  came  when  Adams  oxidized  tetrahydrocannabidiol  with  potassium 
permanganate  in  acetone  and  obtained  menthane  carboxylic  acid  as  one  of  the  pro- 
ducts.9 He  also  found  that  L-raenthyl  chloride  could  be  converted  via  Grignard 
formation  and  carbonation  to  the  same  menthane  carboxylic  acid.  Later  work  by 
Roberts,  Shoppee  and  Stephenson15  has  shown  that  both  3a  and  3p-bromocholestane  are 
converted  to  cholestane-3P-carboxylic  acid.  Assuming  a  similar  reaction  with 
menthyl  chloride,  an  equatorial  carboxyl  group  would  be  expected  thus  giving  the 
thermodynamically  most  stable  eee  isomer.   In  other  evidence,  the  coupling  constant 
J  =  11  for  the  C3  proton  coupling  with  C4  proton  is  indicative  of  a  diaxial  con- 
formation and  thus  a  trans  ring  junction.12  These  data  then  indicate  that  canna- 
bidiol has  a  3  >^ -"trans  ring  junction.  The  total  structure  is  seen  below. 


7CH3 


Ci8 

/\<k 

CH3  CH2 


10 


Early  investigations  by  Adams  et  al.  had  shown  that  naturally  occurring 
cannabidiol  could  be  converted  with  acids  to  two  optically  active,  psychotomimetically 
active  products.16  Sulfur  dehydrogenation  converted  these  materials  to  cannabinol 
thus  proving  their  structures  as  tetrahydrocannabinols.17  Furthermore,  both  of  the 
isomers  when  hydrogenated  yielded  the  same  two  epimers,  proven  similar  by  IR,  KMR 
and  chromoplate  comparison.18  Recent  investigations  have  shown  that  treatment  of 
cannabidiol  with  a  catalytic  amount  of  p -toluene sulfonic  acid  produces  the  A6-^>,k- 
trans -tetrahydrocannabinol,  XIV,  while  treatment  of  cannabidiol  with  dilute  hydro- 
chloric acid  in  ethanol  produces  XV,  the  A1-3,4-trans-tetrahydrocannabinol.18"21 

JH3 


CH3  CH 


5H11 


0^^— CgHn 

Lii3 
OH       XV 


;H 


5nll 


XIV 


Structure  proof  of  the  double  bond  position  in  the  isomers  came  from  an  NMR 
analysis.  In  non-rigid  cannabidiol  the  C3  proton  is  at  t6.15  while  the  C2  proton 
is  at  tk.kl.     The  fixed  ring  system  of  tetrahydrocannabinol,  however,  causes 
deshielding  by  the  aromatic  ring  of  the  C2  proton.  Therefore,  the  A1  double  bond 
was  assigned  to  the  isomer  with  a  C2  proton  signal  at  T3.58  and  the  A6  structure  to 
the  isomer  with  olefinic  C6  proton  at  t^.55,  a  normal,  unaffected  olefinic  proton 
value . 21 


-251- 

Both  the  A1  and  A6  isomers  are  psychotomimetically  active.  Taylor  et  al.22 
have  found  the  A1  isomer  to  isomerize  partially  to  the  A6  isomer  when  chromatographed 
at  280°  (column  10$  GE-SE-30  on  Diatoport  S) .  From  this  observation,  they  suggest 
that  on  smoking  marihuana,  the  psychotomimetic  effect  may  be  due  to  the  A6  isomer 
instead  of  the  A1.  Gaoni  and  Mechoulam20  disagree  after  finding  no  isomerization  of 
the  A1  isomer  to  the  A6  isomer  in  gas  chromatographic  experiments  up  to  300°  using 
a  different  column  (SE-30  on  Chromosorb  W) .  Recent  experiments  by  Lerner  and 
Zeffert3  have  shown  by  using  a  smoking  machine  and  analyzing  the  smoke  by  GC  that 
the  A6  isomer  increased  from  3$  to  only  9$  of  the  total  A1,  A6-tetrahydrocannabinol 
content.  This  would  suggest  that  although  some  isomerization  does  occur  on 
smoking,  the  major  psychotomimetically  active  component  is  still  the  A1  isomer. 

By  treatment  with  p -toluene sulfonic  acid  in  toluene,  the  A1  isomer  can  be 
converted  9Cf?o   to  the  A6  isomer.19  These  same  workers  have  also  noted  a  slow 
isomerization  of  the  A1  to  A6  isomer  on  long  standing.19  Gaoni  and  Mechoulam20 
suggest  the  facile  isomerization  occurs  because  there  is  much  less  steric  crowding 
in  the  A6  isomer  between  the  phenolic  hydroxyl  and  the  C2  protons. 

TOTAL  SYNTHESIS  OF  TETRAHYDROCANNABINOLS 

Mechoulam  and  Gaoni  recently  reported  the  first  total  synthesis  of  a 
psychotomimetically  active  constituent  of  marihuana,  dl-A1-3,^-trans -tetrahydro- 
cannabinol.23 Initially  citral  a  XVI  was  condensed  with  the  lithium  derivative  of 
olivetol  dimethyl  ether  XVII.  This  condensation  led  to  a  complicated  mixture  which 
was  dissolved  in  pyridine  and  treated  with  p-toluenesulfonyl  chloride  to  form  the 
sulfonate  ester  followed  by  ring  closure  and  sulfonate  elimination.   Chromatography 
of  the  mixture  over  10$  silver  nitrate  and  alumina  yielded  a  fraction  with  polarity 
similar  to  natural  cannabidiol  dimethyl  ether.  Re chromatography  of  this  material 
yielded  dl-cannabidiol  dimethyl  ether  XVIII  in  low  yield.  Demethylation  with  methyl- 
magnesium  iodide  gave  an  QCffo   yield  of  dl-cannabidiol  which  upon  treatment  with  0.05$ 
hydrochloric  acid  gave  dl-A1 -3,^4- -trans -tetrahydrocannabinol. 


CHO 


/  ^5-^11 


XVI 


OR   OCH3 


R  -  H 

R  =  SO2C6H4CH3 


0.05$  HC1 
.in  EtOH 


C5H11 


<CH3* 


155-65u 
15  min 


Ln 


5^11 


A  second  synthesis  of  both  dl-A1  and  A6-3,k -trans -tetrahydrocannabinol  was 
carried  out  by  Fahrenholtz,  Lurie  and  Kierstead.24  This  synthesis  involved  a 
Pechmann  condensation  of  olivetol  XIX  with  diethyl -cc-acetoglutarate  XX  to  form  the 
expected  coumarin  XXI.  Cyclization  of  XXI  using  sodium  hydride  in  an  aldol  type 
condensation  gave  XXII.  Ethylene  glycol  converted  XXII  into  the  corresponding 
ketal  XXIII  which  exists  in  two  polymorphic  forms.  Reaction  of  the  ketal  with 
methyl  magnesium  iodide  followed  by  acidic  hydrolysis  gave  XXIV.  Birch  reduction 
converted  XXIV  to  the  trans  ketone  XXV.  Carbinol  XXVI  was  obtained  by  conversion  of 
XXV  to  its  tetrahydropyranyl  ether  followed  by  treatment  with  methyl  magnesium 
iodide  and  subsequent  cleavage  of  the  protecting  group.  Dehydration  of  XXVI  using 


.o^o^ 


a  catalytic  amount  of  p -toluene sulfonic  acid  in  "benzene  gave  dl-A6-3,4-trans- 
tetrahydrocannabinol  as  a  single  isomer  shown  by  glpc.  Treatment  of  XXVI  with  Lucas 
reagent  gave  the  corresponding  chloride  XXVII.  Dehydrohalogenation  of  XXVII  with 
sodium  hydride  in  refluxing  tetrahydrofuran  gave  in  quantitative  yield  a  mixture  of 
the  A1,  lh1of   and  the  A6,  26$,  tetrahydrocannabinol  isomers  as  shown  by  glpc. 


OH 


QJJEt 

<L    PH3  OH 


C5H11"  v 

XIX 


XX 


1.  CH3MgI 
x2.  H"1" 


C=H 


5nll 


XXI 


DMSO 
NaH 
,15-20° 


XXIV 


c  (H0CH2)2 
XXIII 


5H11 


XXII 


XXIV 


Li/Ma. 
-78° 


1  Conversion  to  ether 

2  CHa^gl 

5!  Removal  of  hydroxyl 

C5H11  protecting  group 


XXV 


CrH 


5-n-ll   «. 


ZnCl2 


TIC1 


C=H 


5nll 


^05^11 


XXVI 


71$ 


-tol-S03H 


O^vx^CgHn 


Recently  Taylor,  Lenard  and  Shvo  carried  out  a  one  step  synthesis  utilizing 
citral  and  olivetol  in  the  presence  of  a  10$  benzene  solution  of  boron  trifluoride 
etherate.22  Chromatography  of  the  mixture  over  Florisil  followed  by  preparative 
VPC  separation  gave  a  20$  yield  of  dl-A6-3,4-trans -tetrahydrocannabinol.  ,   Some  of 
the  A6-cis  isomer  was  also  obtained  and  identified  by  comparison  with  a  sample  of 
the  cis  isomer  prepared  by  an  independent,  unequivocal  route. 


-PS^- 


53- 


CHO 


BF3  etherate 
benzene 


CsH 


5^11 


CH3^.0/    ^CsHu 
CH3 


BIOGENESIS 

No  labelling  studies  have  been  performed  as  yet,  but  by  examination  of  the 
other  constituents  of  marihuana,  the  following  biogenetic  pathway  has  been  postulated. 
Olivetol  probably  is  formed  via  the  usual  head  to  tail  linking  of  acetate  units  to 
form  phenolic  compounds. 


CH3CCH2CCH2 

^6=0 

CH2  CH2-CC>2H 


0=C 


I 
C=0 


CH3CCH2CCH2 

X/CO2H 

oit^'Sdh 


c=h 


5^11 


red, 


-CO; 


CH; 


It  is  uncertain  whether  the  pentyl  side  chain  is  formed  from  acetate  units  or 
whether  hexanoic  acid  is  incorporated  directly.25  Also  in  question  is  when  the 
decarboxylation  occurs. 

Once  formed,  olivetol  XXVIII  or  the  acid  precursor  XXIX  probably  condenses 
with  geranyl  pyrophosphate  to  form  either  cannabigerol  XXX  or  cannabigerolic  acid 

.OH 

'5H11 


R 


p2or3 


XXVIII  R  =  H 
XXIX  R  =  CO2H 


XXX  R  =  H 
XXXI  R  =  COsH 


PPi 


Both  cannabigerol  and  cannabigerolic  acid  have  been  found  in  extracts  of  marihuana,21 
These  two  materials  then  condense  further  to  the  corresponding  cannabidiol  XXXII 
and  tetrahydrocannabinol  XXXIII  or  cannabidiolic  acid  XXXIV  and  tetrahydrocannabinolic 
acid  XXXV.26'27  Whether  the  carboxylate  derivatives  are  converted  in  part  in  any 
of  the  preceeding  steps  to  the  decarboxylated  derivatives  is  still  uncertain.  The 
carboxylated  derivatives  have  a  high  antibiotic  activity  against  gram  positive 
bacteria  which  marihuana  lacks  and  they  may  only  compose  another  metabolic  pathway 
of  the  plant.26 


■cy'-r- 


C=H 


5-^11 


R  =  H  XXXII  R  =  H  XXXIII  R  =  H 

R  =  COaH  XXXIV  R  =  C02H  XXXV  R  =  CO2H 

SUMMARY 

Recent  studies  have  fully  elucidated  the  structures  of  the  natural  psycho- 
tomimetic ally  active  constituents  of  marihuana,  A1  and  A6-3^^ -"trans -tetrahydro- 
cannabinol and  a  likely  precursor,  cannabidiol.  Total  synthesis  of  the  dl  forms  has 
also  been  accomplished.  Speculations  on  the  biogenetic  pathway  have  been  advanced. 

BIBLIOGRAPHY 

1.  D.  Downing,  Quart.  Revs.,  16,  152  (1962). 

2.  M.  C.  Nigam,  K.  I.  Handa,  I.  C.  Nigam,  and  L.  Levi,  Can.  J.  Chem. ,  4j5,  3372-6 

(1965). 

3.  Chem.  and  Eng.  News,  44,  Dec.  26  (1966). 

4.  "Drug  Addiction,"  Collier's  Encyclopedia,  1966,  VIII,  395. 

5.  T.  Wood,  W.  Spivey,  and  T.  Easterfield,  J.  Chem.  Soc.,  20-36  (1899). 

6.  R.  S.  Cahn,  J.  Chem.  Soc,  1342-53  (1932). 

7.  R.  Adams,  M.  Hunt  and  J.  H.  Clark,  J.  Am.  Chem.  Soc,  62,  196-200  (1940). 

8.  R.  Adams,  C.  K.  Cain  and  H.  Wolff,  J.  Am.  Chem.  Soc,  S|,  732-3  (1940). 

9.  R.  Adams,  C.  K.  Cain  and  J.  H.  Clark,  J.  Am.  Chem.  Soc,  62,  735-7  (1940). 

10.  R.  Adams  and  B.  Baker,  J.  Am.  Chem.  Soc,  62,  2405-8  (194*67. 

11.  R.  Ghosh,  A.  R.  Todd  and  S.  Wilkinson,  J.  Chem.  Soc,  1393-6  (1940). 

12.  R.  Mechoulam  and  Y.  Shvo,  Tetrahedron,  12,  2073-8  (I963). 

13.  R.  Adams,  H.  Wolff,  C.  Cain  and  J.  Clark,  J.  Am.  Chem.  Soc,  62,  2215-9  (1940). 

14.  L.  M.  Jackman,  Applications  of  NMR  Spectroscopy  in  Organic  Chemistry,  Mac- 
millian  Co.,  N.Y.  ,  N.Y.  ,  I96TJ  5c~ 

15.  G.  Roberts,  C.  W.  Shoppee  and  R.  J.  Stephenson,  J.  Chem.  Soc,  27O5-I5  (1954). 

16.  R.  Adams,  D.  Pease,  C.  K.  Cain  and  J,  H.  Clark,  J.  Am.  Chem.  Soc,  62,  2402- 
4  (1940). 

17.  R.  Adams  and  B.  Baker,  J.  Am.  Chem.  Soc,  62,  2401-2  (1940). 

18.  Y.  Gaoni  and  R.  Mechoulam,  Tetrahedron,  22,  1481-8  (1966). 

19.  R.  Hiveiy,  W.  Mosher  and  F.  Hoffmann,  J.  Am.  Chem.  Soc,  88,  1832-3  (1966). 

20.  Y.  Gaoni  and  R.  Mechoulam,  J.  Am.  Chem.  Soc,  88,  5673-5  ^1966). 

21.  Y.  Gaoni  and  R.  Mechoulam,  J.  Am.  Chem.  Soc,  B£,  1646-7  (1964). 

22.  E.  Taylor,  K.  Lenard  and  Y.  Shvo,  J.  Am.  Chem.  Soc.,  88,  367-9  (I966). 

23.  Y.  Gaoni  and  R.  Mechoulam,  J.  Am.  Chem.  Soc.,  87,  3273-5  (1965). 

24.  K.  Fahrenholtz,  M.  Lurie  and  R.  Kierstead,  J.  Am.  Chem.  Soc,  88,  2079-80 
(1966).  "~" 

25.  T.  A.  Geissman,  Biogenesis  of  Natural  Compounds,  Ed.  by  P.  Bernfeld,  Macmillian 
Co.,  N.Y. ,  N.Y.,  1963,  580-9. 

26.  R.  Mechoulam  and  Y.  Gaoni,  Tetrahedron,  21,  1223-9  (I965). 

27.  F.  Korte,  M.  Haag  and  V.  Classen,  Angewande  Chemie  Int.  Ed.,  4,  872  (1965). 


-255- 

EENZENE  PHOTOLYSIS 

Reported  by  Warren  J.  Peascoe  March  9 >  19&7 

Many  examples  of  reactions  involving  electronically  excited  "benzene  have 
recently  been  reported.   This  seminar  will  review  these  reactions.  A  brief  sum- 
mary of  the  general  processes  which  occur  upon  absorption  of  a  quantum  of  ener- 
gy is  presented. 

Absorption  of  a  quantum  of  energy  by  a  molecule  in  the  ground  state  So 
produces  an  excited  singlet  which  may  return  to  the  ground  state  by  various  pho- 
tophysical  processes.1  Singlets  in  the  higher  excited  states  S2,  S3,  .  .  . 
undergo  internal  conversion  to  the  first  excited  singlet  state  Si  or  undergo 
inter system  crossing  to  produce  triplets.  The  Si  singlet  may  also  undergo  in- 
ter system  crossing  to  produce  a  triplet,  or  it  may  return  to  the  ground  state 
by  internal  conversion  or  fluorescence.   The  higher  excited  triplets  rapidly 
undergo  internal  conversion  to  the  Ti  state  and  may  return  to  the  ground  state 
by  intersystem  crossing  or  by  phosphorescence. 

The  irradiation  of  benzene  i,n^or  with  light  in  the  2300-2700  A  region  has 
recently  been  reviewed  by  Noyeir/ 58who  concluded  that  the  absorbed  energy  is 
efficiently  dissipated  by  photophysical  processes,,   Irradiation  at  a  shorter 
wavelength,  1.0^9  A.,  has  been  reported  by  several  groups  to  lead  to  the  formation 
01  polymer  deposited  on  the  walls  of  the  reactor,  two  transient  intermediates, 
and  minor  amounts  of  hydrogen,  methane,  acetylene,  ethane ,  toluene,  and  a  C2 
or  C3  substituted  benzene,3  8}55     The  quantum  yield  (number  of  molecules  which 
undergo  a  specific  reaction/number  of  molecules  electronically  excited)  of  the 
formation  of  hydrogen  was  reported  to  be  about  0.02  by  Lipsky  and  coworkers,3 
and  Foote  and  coworkers  reported  the  formation  of  acetylene  with  a  quantum  yield 
of  O.OI5.4  The  major  transient  intermediate  has  been  identified  as  fulvene  1 
independently  by  Ward  and  coworkers5  and  by  Kaplan  and  Wilzbach55. 

It  was  purified  by  preparative  gas  chromatography  and  had  the  same 
uv,  ir,  nmr,  and  mass  spectra  as  authentic  fulvene.  Fulvene  was 
shown  to  be  the  same  intermediate  detected  but  not  identified  by 
the  Foote4  and  Lipsky6  groups  by  a  comparison  of  the  vapor  phase 
ultraviolet  spectra  and  retention,  time  on  gas  chromatography. 
The  minor  intermediate  ir  approximately  l/6  the  concentration  of 
1       fulvene  was  detected  by  Kaplan  and  Wilzbach  by  gas  chromatography 
and  by  its  uv  spectrum.55  Since  it  absorbed  in  the  same  region 
as  fulvene,  its  uv  spectrum  was  determined  from  the  difference  between  the  uv 
spectrum  of  irradiated  benzene,  vs.  benzene  blank,  and  the  spectrum  of  fulvene 
vapor  in  the  same  concentration  as  in  irradiated  benzene.   On  the  basis  of  its 

uv  absorption  (k  '     2500  A)  and  extinction  coefficient  (log  e^  4.5)  , 

max 

l,3-hexadiene-5~yne  "was  proposed  as  a  possible  structure. 

The  Foote  and  Lipsky  groups  both  observed  that  fulvene  formed  rapidly  and 
reached  a  low  steady-state  concentration.   The  initial  rate  of  formation  of  ful- 
vene was  found  to  be  inversely  proportional  to  the  pressure  of  added  nitrogen 
by  Foote  and  coworkers  for  nitrogen  pressures  of  less  tnan  one  atm;4  Lipsky  and 
coworkers  found  that  the  rate  of  formation  decreased  with  increasing  nitrogen 
pressure  from  0.1  to  50  atm. 6  The  quantum  yield  for  the  disappearance  of  benzene 
was  also  inversely  related  to  the  pressure;  it  was  1  at  benzene  pressures  of  less 
than  0.1mm4 ;  fell  to  0.25  at  1mm,  and  decreased  further  if  nitrogen  was  added.6 
This  dependence  on   pressure  indicated  that  the  reaction  was  collision  quenched. 
The  polymer  seemed  to  be  formed  by  a  further  reaction  of  fulvene  since  the  ratio 
of  the  quantum  yield  of  benzene  disappearance  to  the  quantum  yield  of  the  formation 
of  fulvene  was  a  constant  as  the  nitrogen  pressure  was  varied. 

Lipsky  and  coworkers  calculated  from  the  absorption  oscillator  strengths 
that  the  second  and  third  excited  benzene  singlets  had  fluorescence  lifetimes 
of  less  than  10  12  sec.9  Since  no  fluorescence  was  observed  from  these  states, 
the  state  lifetime  must  be  even  shorter;  and  a  pressure  of  ca.  100  atm  is  required 


-256- 

in  order  to  have  the  mean  time  between  collisions  equal  the  calculated  fluores- 
cence lifetime.  Fluorescence  and  triplet  sensitized  emission  of  biacetyl  were 
observed  upon  excitation  of  benzene  to  the  Si  state  with  light  of  wavelength 
greater  than  2200  A.  However,  neither  fluorescence  from  any  singlet  state 
nor  sensitized  emission  of  biacetyl  was  observed  upon  excitation  with  light  of 
wavelength   in  the  2200-1600  A  region  corresponding  to  excitation  to  the  S2 
or  S3  state.   Thus  either  a  direct  photochemical  reaction  or  very  efficient  in- 
ternal conversion  to  the  ground  state  is  required  from  the  S2  and  S3  states. 
The  observed  collision  quenching  at  low  pressures  ruled  out  a  direct  photoreaction 
upon  irradiation  at  1849  A.   Lip sky  and  coworkers  were  led  to  propose  that  the 
initially  excited  S3  or  S2  benzene  very  rapidly  underwent  internal  conversion 
to  vibrationally  excited  ground  state  which  either  lost  its  vibrational  energy 
to  produce  benzene  or  isomerized  to  fulvene.6 

The  irreversible  formation  of  fulvene  and  polymer  upon  irradiation  of  liq- 
uid benzene  at  wavelengths  greater  than  2000  A  under  nitrogen  atmosphere  was 
reported  by  Bryce-Smith  and  coworkers,10-'11  with  an  approximate  quantum  yield 
of  0.01.  The  uv  spectrum  of  the  product  resembled  that  of  authentic  fulvene, 
and  the  one-to-one  adduct  formed  between  it  and  maleic  anhydride  was  found  to 
have  the  same  infrared  spectrum  and  mixed  melting  point  as  the  adduct  prepared 
from  authentic  fulvene. 

The  photolytic  isomerization  of  orthodi substituted  benzenes  has  been  re- 
viewed through  1.964  in  a  previous  seminar.12  The  isolation  of  substituted 
dewar  benzenes  and  isomerizations  through  interconvertible  dewar  benzenes  were 
discussed.  Additional  examples  of  isomerizations  of  substituted  benzenes  have 
been  reported.   The  formation  of  perfluorodewar  benzene  from  the  irradiation 
at  2537  A  of  perfluorobenzene  in  the  vapor  phase  has  been  independently  reported 
by  Haller13  and  by  Camaggi  and  coworkers.14  Lemal  and  Lokensgard15  have  con- 
verted hexamethyldewar  benzene  into  a  mixture  of  20$  hexamethylprismane  and  hex- 
amethylbenzene  upon  irradiation  at  2537  A  in  butane  solution.   The  hexamethyl- 
prismane could  be  phot ochemic ally  converted  to  the  dewar  benzene  and  hexamethylbenzene, 

Wilzbach  and  Kaplan  have  reported  that  1,3^5-tri-t-butylbenzene   2  photo- 
isomerized  at  2537  A  in  isohexane  to  l,2,4-tri-t~butylbenzene  J5  through  a  benz- 
vaiene   intermediate  4. 16  Irradiation  of  2  or  ^  led  to  a  photostationary  mixture 
with  7° 3°/°  2,  20.67c  j5,  less  than  0.7$  h,   7.1$  £,  and  64.8$  6.   Two  fulvenes  were 
also  produced  in  low  yield  from  2.  The  product  mixtures  were  analyzed  by  nmr 


(0.04) 


I  6 

utilizing  the  characteristic  t-butyl  peaks,  and  the  quantum  yields  shown  in 

the  diagram  were  determined  for  separate  isomers  2,  3,  4>  and  5*   The  prismane  6 
could  not  be  completely  separated  from  the  dewar  benzene  5,  a^d  the  quantum  yields 
for  its  reactions  are  approximate.   The  products  were  identified  on  the  basis 
of  their  spectral  properties  and  rearomatization  reactions  on  heating.   The 
benzvalene  4  showed  a  nmr  methyne  proton  ABX  pattern,  t  8,27,  8.35;  and  4.95 
for  Ay  B.  and  X  respectively,  with  J.=6.6cj<  J  -2. 45 ,  and  J==1.25  cps.   The 

A.D  AX.  £3A 

t-butyl  protons  were  observed  at  1   8.94  (18  H)  and  8.99  (9  fi)  .  The  benzvalene 
rearomatized  upon  heating  exclusively  to  j5  leading  to  the  assignment  of  structure  4. 

The  simplest  method  for  photoconversion  of  ^   into  4  involves  bending  the 


-2.57- 
planar  ring  of  J5  along  with  the  formation  of  transannular  bonds;  a  similar  ring 
bending  of  2  along  with  the  formation  of  transannular  bonds  leads  to  7  rather 


than  k.      Wilzbach  and  Kaplan  did  not  report  the  isolation  and  identification 
of  the  intermediate  initially  formed  from  the  irradiation  of  2,  and  thus  7  rather 
than  h  may  be  the  intermediate  in  the  photoisomerization  of  2  into  3«  It  is 
possible  that  J  may  not  have  been  detected  since  the  nmr  chemical  shifts  of  the 
t-butyl  protons  of  2  would  be  expected  to  be  very  similar  to  the  chemical  shifts 
of  the  t-butyl  protons  of  h.     A  second  possibility  is  that  4  was  the  intermediate 
and  that  it  was  formed  from  the  isomerization  of  7  in  the  ground  or  excited  state. 
As  a  member  of  the  class  of  valene  compounds ,  the  isomerization  of  benzvalene 
7  to  benzvalene  h  might  be  expected.17  Evidence  that  irradiation  of  2  leads 
to  the  initial  formation  of  7  in  either  an  excited  state  or  in  the  ground  state 
has  been  reported  by  Kaplan  and  coworkers o18  They  found  that  2  irradiated  in 
methanol  produced  8  with  a  quantum  yield  of  0.15  and  that  no  isomeric  tri-t- 

butylbenzenes  were  formed. 


o 


nv 
eOH 


) 


L  2 


OMe 


The  product  was  identified 
by  its  spectral  properties, 
and  the  proton  in  position 
6  was  assigned  the  endo- 
configuration  on  the  basis  of 
its  nmr  chemical  shifty  T  9«27> 
and  coupling  constants 


Ji,6=4.6  and  J5.,6=2.7  ops.   The  t-butyl  groups  were  shown  not  to  be  on  adjacent 
caroons  by  thermal  conversion  of  b   into  l^^-tri-t-butylfulvene. 

The  formation  of  benzvalene  upon  irradiation  of  benzene  in.  the  liquid  phase 
at  2537  A  has  been  reported  by  Kaplan  and  coworkers56.  Concentrations  of  product 
(ifo   based  on.  benzene)  high  enough  for  nmr  spectroscopy  and  preparative  gas  chrom- 
atography were  obtained  by  irradation  of  dilute  solutions  of  benzene  in  hexa- 
decane  at  650 -   Treatment  with  methanolic  HCT  of  either  irradiated  benzene  or 
the  product  purified  by  preparative  gas  chromatography  and  absorbed  in  isohexane 
produced  two  products  identified  by  methods  not  described  as  6-endo-methoxy 
(3.1.0)bicyclo-2-ene>  10  (R^CHa),  and  one  of  the  4-methoxy  isomers,  9  (R=0CH3). 
The  uv  spectrum  of  the  irradiated  benzene  product  showed  no  maximum  above  2100  A 
but  had  a  broad  shoulder  between  2200  A  and  2J00  A  ( S-^OO)  .   The  nmr  spectrum 
of  the  product  showed  three  resonances  of  equal  area*   an  unsymmetrical  triplet  (1.5-and 
1.7-cps  couplings)  at  T  k.Ok?   a  symmetrical  triplet  (l.5-eps  couplings)  at  T  6.47,  and 

a  quintet  (L   J=6.2  cpsj  at  T  8.16.   The  product  rearomatized  only  slowly  at 
room  temperature  and  was  assigned  the  benzvalene  structure. 

Photoaddition  reactions  have  been  reported  which  may  be  rationalized  as 
formal  additions  to  benzvalene  and  to  dewar  benzene.  Farenhorst  and  Bickel19 
reported  that  benzene  .irrarHqted  in  acetic  acid  with  a  low  pressure  mercury  lamp 
produced  polymer,  £  (*R--0C0CH3) ,  and  two  other  unidentified  products  in  lesser  amounts. 
The  acetoxy  group  of  Q  (R=-X0CH3)  is  reportedly  in  the  exo-position  from  nmr 
analysis,  but  only  chemical  shifts  without  any  splitting  patterns  were  reported. 
Without  nmr  data  for  the  endo-  and  exo-compounds  the  stereochemistry  must  remain 
tentative.  Compound  £  (R=-0C0CH3)  could  not  be  hydrolyzed  in  either  mild  acid 
or  base  but  in  both  cases  produced  polymer  and  a  yellow  discoloration  due  to 
fulvene  as  indicated  by  the  uv  spectrum.  Irradiation  of  aqueous  phosphoric 
acic  saturated  with  benzene  produced  two  products  tentatively  identified  as  alde- 

"*"°  °  /p-  OH)  with  the  hydroxy  group  in  the 


-258- 
exo -posit ion.  The  stereochemistry  of  the  hydroxy  group  was  based  only  on  report- 
ed nmr  chemical  shifts. 


5  k 


+     R 


R 


9 


10 


Kaplan  and  coworkers18  reported  that  benzene  irradiated  in  trifluoroethanol 
yielded  2.  and  ±2  (R=0CH2CF3)  in  the  ratio  of  2:1  with  a  total  quantum  yield  of 
0.05°  The  identification  of  the  products  was  based  on  spectral  data.   Nmr  double 
resonance  analysis  of  10  (R=0CH2CF3)  showed  H6  to  be  coupled  with  Hi  and  H5 
( J=7cps)  indicating  an  exo -posit ion  for  H6  to  the  authors.  The  coupling  constant 
is  closer  to  the  reported  vicinal  cyclopropyl  coupling  constants  5*2  to  8.0  cps 
for  trans  hydrogens  than  to  8.0  to  LI. 2  cps  for  cis  hydrogens,,57  and  the  opposite 
stereochemistry ,   H6-endo,  is  indicated  at  C6.  The  stereochemistry  of  9  (R=0CH2CF3) 
was  not  determined. 

The  photolytic  formal  addition  of  olefins  to  benzvalene  has  also  been  re- 
ported by  two  groups.  Wilzbach  and  Kaplan20  reported  that  products  correspond- 
ing to  11  could  be  isolated  by  gas  chromatography  from  the  product  mixtures  from 
the  irradiation  at  2537  A  of  benzene  and  ethylene,  cis-2-butene,  cyclopentene , 
or  2,3-dimethylbut-2-ene.   The  products  were  characterized  by  their  spectral 
properties  and  extensive  nmr  decoupling  experiments.   Bryce-Smith  and  coworkers21 
reported  that  equimolar  mixtures  of  cyclooctene  and  benzene  at  room  temperature 


R2 
\ 


/' 


R- 


R-3    R.4 


^i*^C  j/ 


+ 


R2 
R- 


/ 


rTT 

R. 


11 


2—f 

Ri 


12 


or  at  -60°  irradiated  in  the  2350-2580  A  region  produced  two  1:1  addition  products 
in  the  ratio  of  8:1.   Benzophenone  and  acetone  were  ineffective  as  sensitizers  but 
|3-propiolactune  increased  the  rate  of  addition  two-fold.   The  major  product  11 
was  characterized  by  spectral  studies }   products  from  catalytic  hydrogenation,  and 
formation  of  1:1  adducts  with  tetracyanoethylene  or  phenyl  azide.   But-1-ene,  oct-1-ene, 
cyclohexene,  cycloocta-l,5-diene ,  and  ethyl  vinyl  ether  also  gave  1:1  photoadducts 
analogous  to  11.   The  structure  of  the  minor  product  was  suggested  to  be  12  on  the 
basis  of  a  sharp  olefinic  singlet  at  T  k.kj.      Srinivasan  and  Hill22  reported  the 
-formation  of  12  in  50'°  yield  from  the  photoaddition  of  benzene  and  cyclobutene  at 
2537  A.   The  product  was  identified  on  the  basis  of  its  spectra  (a  sharp  nmr  reso- 
nance for  two  olefinic  protons  at  T  4.25)  and  the  products  produced  upon  catalytic 
hydrogenation.   Upon  heating  to  200° ,  the  product  decomposed  to  produce  benzene  and 
butadiene. 

Bryce-Smith  and  coworkers23  have  noted  that  both  11  and  12  provide  an  olefinic 
functional  group  and  should  be  able  to  add  a  second  molecule  of  benzene  in  a  step 
towards  photopolymerization.   They  irradiated  benzene  and  small  amounts  of 
cyclooctene  (0.35  mole-fa)  or  the  cyclooctene-benzene  1:1  adduct  (0.06  mole-fo) 
for  100  hr  with  a  medium  pressure  mercury  lamp  and  found  a  polymer  to  be  produced. 
The  polymer  was  fractionated  by  its  solubility  in  hexane,  acetone,  benzene,  and 
chloroform.   The  major  fraction  in  benzene  had  a  molecular  weight  of  about  1500. 
There  were  no  aromatic  protons  in  the  nmr  spectrum,  and  the  ir  spectrum  showed  six 
of  the  seven  bands  observed  in  monomeric  1:  ,   The  trace  amount  of  olefin  was 
sufficient  to  suppress  effectively  the  formauxon  of  fulvene.  Polymers  with  different 
properties  were  reported  to  be  formed  with  light  of  wavelength  of  ca.  2000  A. 

Koltzenburg  and  Kraft24-*25  reported  the  photoaddition  of  1,3-ciienes  to  benzene 
upon  irradiation  at  2537  A.   Gas  chromatography  showed  at  least  ten  products  from 
the  reaction  of  isoprene  with  benzene  with  a  hBp   ;yield  of  1/$  and  a  2j°r   yield  of  14. 


-259- 
The  products  1J5  and  ih   were  identified  on  the  basis  of  their  chemical  and  spectral 


o 


I 


h3L 


+ 


15  14  H 

properties.   Toluene,  o- xylene,  perfluorobenzene,  and  1,2,4,5-tetrafluorobenzene 
have  also  been  found  to  form  dimeric  adducts  with  isoprene.   The  photoadduct 
from  butadiene  and  benzene  dimerized  to  for::  15.   Attempts  to  sensitize  the  reaction 
with  benzophenone  led  only  to  the  formation  of  dimeric  dienes. 

A  2:1  photoadduct  formed  from  maleic  anhydride  or  N- substituted  maleimide 
and  benzene  has  been  reported.26  29  These  reactions  which  proceed  by  photo- 
activation  of  the  substrate  or  a  charge-transfer  complex  between  substrate  and 
benzene,  rather  than  by  direct  photoactivation  of  benzene,  will  not  be  discussed. 

Two  more  photoaddition  reactions  for  which  no  evidence  is  available  for 
determination  of  the  erMted  rpqctine;  species  must  be  considered.   The  photo- 
addition  of  methyl  acetylenecarboxylate  or  dimethyl  acetylenedicarboxyiate  to 
benzene  has  been  reported  by  two  groups  to  produce  substituted  cyclooctatetraenes 
17. 30  32  The  reaction  is  thought  to  proceed  by  1,2-addition  to  benzene  to  form 
lb  which  isomerizes  to  17.   The  cyclooctatetraene  acid  formed  by  saponification 


o 


-R 


■R1 


hv 


R 


R' 


16  17 

of  the  ester  was  identical  with  the  acid  prepared  by  an  alternate  route.   The 
diacid  was  assigned  structure  17  on  the  basis  of  its  spectra  and  products  produced 
by  catalytic  hydrogenation.  The  formation  of  18  by  photoaddition  of  2-methylbut-2-ene 


CN 


+ 


hv 


18  12 

to  benzonitrile  has  been  reported  by  Atkinson  and  coworkers.33  They  found  that 
benzophenone  effectively  quenched  the  reaction  and  that  aliphatic  acetylenes 
added  to  give  cyclooctatetraenes  19. 

To  account  for  the  photoisomerizations  and  photoaddition  reactions,  Bryce- 
Smith  and  Longuet-Higgins  have  proposed  a  mechanism  which  allows  the  rationalization 
of  the  observed  products  in  terms  of  the  lowest  benzene  singlet  and  triplet.34  The 
first  benzene  singlet  which  has  B   5^/mmetry35  is  antisymmetric  about  a  plane 

2U 

through  any  opposite  pair  of  caroun  atoms ;     the  lowest  electronic  configuration 

of  the  singlet  biradical  21  is  also  antisymmetric  about  its  plane  of  symmetry. 

An  orbital  correlation  diagram36  indicates  that  20  may  pass  adiabatically  into  21. 

)     Fulvene 


in  addition 


-)  R 


-}     Benzvalene 


4  fa 


R 


"^^ — >     rrrrg, 

>      Pr i  smane 


d\  1>1  "^ ">        1,4-adduct 


24 


-260- 
Similarly  one  component  of  the  lowest  benzene  triplet,  which  has  B   symmetry37 

and  is  antisymmetric  about  a  plane  bisecting  any  opposite  pair  of  carbon-carbon 
bonds,  may  adiabatically  pass  into  the  triplet  state  of  either  of  the  diradicals 
23  or  24o   The  lowest  electronic  configuration  of  both  23  and  2J+  in  the  trip- 
let state  is  antisymmetric  about  a  plane  bisecting  the  terminal  bonds  of  23 
and  the  double  bonds  of  24.   The  diradicals  are  proposed  to  react  as  indicated 
to  form  the  observed  products  from  electronically  excited  benzene.  There  is 
no  experimental  evidence  which  requires  the  diradicals  21,  2J5,  and  2k   to  be  on 
the  reaction  path  leading  from  electronically  excited  benzene  to  products,  and 
it  is  possible  that  benzene  in  the  lowest  singlet  or  triplet  state  reacts  dir- 
ectly to  form  the  observed  products.  The  experimental  determination  of  the  mult- 
iplicity of  the  electronically  excited  state  leading  to  the  formation  of  dewar 
benzene,  dewar  benzene  addition  products,  prismane,  and  1,4- addition  products 
would  be  useful  in  evaluating  this  mechanism  since  triplet  states  are  predicted. 
The  formation  of  cyclooctatetraenes  by  the  1,2-addition  of  acetylenes  to  benzene 
and  the  1, 2~add.it ion  to  benzonitrile  were  not  explicitly  considered  by  the  authors. 
It  has  not  been  established  that  these  reactions  must  proceed  by  attack  of  excited 
benzene  on  the  substrate.  An.  alternate  mechanism,  attack  of  excited  substrate 
on  benzene,  would  not  require  modification  of  the  proposed  reaction  scheme. 

An  alternate  mechanism  involving  a  single  highly  reactive  intermediate  has 
been  proposed  by  Farenhorst,38  van  Tamelen39  has  pointed  out  that  the  Woodward- 
Hoffmann  rules40  predict  preferential  conrotatory  ring  opening  of  cyclobutene 
systems,  and  that  conrotatory  ring  opening  of  dewar  benzene  leads  to  trans -benzene  25. 


conrotatory 


I? 


ring  opening 

The  trans-benzene  pi-orbitals  are  topologically  equivalent  to  a  conjugated  six- 
membered  Moebius  ring.   The  HMO  energy  calculated  by  Heilbronner41  for  a  ground 
state  six-membered  Moebius  system  was  shown  to  be  equal  to  the  HMO  energy  of 
benzene  in  its  first  excited  state.  Farenhorst  thus  proposed  that  benzene  in 
an  excited  state  isomerized  to  trans -benzene  in  the  ground  electronic  state  with 
a  Moebius  pi-electronic  structure  and  that  the  trans -benzene  reacted  to  form  the 
observed  products.  He  also  pointed  out  that  a  possible  transition  state  26 
leading  to  the  formation  of  dewar  benzene  and  a  transition  state  27  leading  to 
benzvalene  would  have  nearly  the  same  energy  as  trans-benzene.   These  transition 


/  i 

/  1   2 
26 

states  consist  of  a  localized  double  bond  (C5-C6)  and  a  conjugated  four-membered 
Moebius  ring  (Ci,C2,C3,C4) \   the  transition  state  27  also  contains  transannular 
bonds  between  Ci  and  C3  and  between  C2  and  C4. 

Irradiation  at  2.537  A  of  benzene  in  rigid  organic  glasses  at  liquid  nitro- 
gen temperature  has  been  studied  by  many  groups.42  54  A  product  proposed  by 
Anderton  and  coworkers  to  be  a  substituted  hexatriene  was  detected  by  its  uv 
spectrum.43  The  position  of  the  three  observed  peaks  varied  with  the  solvent 
used  to  form  the  rigid  glass.43  46  Migirdicyan  and  coworkers  reported  that  the 
product  from  the  reaction  in  ethanol  28  (R*  CHOHCH3)  dehydrated  on  preparative 


Cg)   mSd->  H  '•       iR  //(f^C^j/i 


alass 


-261- 

gas  chromatography  to  produce  1,3,5,7-octatetraene  identified  by  its  uv  spec- 
trum. 45'j46  After  gas  chromatography  the  product  from  the  reaction  in  methanol 
28  (R=  -CH2OH)  still  showed  the  same  three  uv  peaks.   The  formation  of  the  sub- 
stituted hexatriene  has  been  reported  to  be  first  order  in  the  intensity  of  the 
exciting  light  by  Migirdicyan  and  coworkers.47"49  There  is  no  evidence  which 
allows  the  distinction  between  a  four-centered  reaction  and  one  involving  a  hex- 
atriene diradical. 

Several  groups  have  observed  the  esr  spectra  of  solvent  radicals  upon  ir- 
radiation of  benzene  in  rigid  glasses.50  53  The  esr  spectra  of  the  radicals 
from  the  2537  A  irradiation  of  benzene  in  the  glass  and  esr  spectra  of  the  rad- 
icals produced  by  the  y-radiolysis  of  the  pure  solvent  glass  were  the  same. 
The  formation  of  molecular  hydrogen  with  a  yield  greater  than  twice  the  yield 
of  radical  formation  was  reported  by  Shelimov  and  coworkers.51  The  source  of 
most  of  the  hydrogen  was  the  solvent  since  the  use  of  benzene-d6  led  to  a  ratio 
of  H2/HD  of  9»5  a"t  benzene-d6  concentration  of  1.8X10  3  M  and  8.1  at  benzene-d6 
concentration  of  6.0X10  2  M  as  determined  by  mass  spectroscopy.  The  formation 
of  both  hydrogen  and  solvent  radicals  was  reported  to  be  second  order  in  the 
intensity  of  irradiation. 50>52j'53  The  reaction  is  interpreted  in  terms  of  trip- 
let benzene  absorbing  a  second  quantum  of  energy  to  produce  a  doubly  excited 
triplet.   The  doubly  excited  triplet  transfers  its  energy  to  a  solvent  molecule 
(RH  =  3-methylpentane ,  methylcyclohexane,  2-methyltetrahydrofuran,  methanol, 
isopropyl  alcohol,  or  cyclohexane)  producing  ground  state  benzene  and  a  solvent 
triplet  which  dissociates  into  two  radicals,  H*  and  R° .   The  hydrogen  radical 
abstracts  a  hydrogen  from  another  solvent  molecule  and  produces  a  second  trapped 
solvent  radical  R°.52  Support  for  the  biphotonic  process  involving  a  doubly 
excited  triplet  comes  from  the  recent  observation  by  Godfrey  and  Porter54  of 
the  absorption  spectrum  of  triplet  benzene  which  was  observed  after  flash  photol- 
ysis of  benzene  in  a  rigid  matrix  at  77°  K. 

SUMMARY 

Benzene  irradiated  in  the  vapor  phase  at  wavelengths  shorter  than  2200  A 
produces  polymer ,  f ulvene ,  an  unidentified  intermediate „  and  decomposition  prod- 
ucts. Fulvene  and  benzvalene  have  been  isolated  from  irradiated  benzene  solutions 
and  the  presence  of  benzvalene  and  dewar  benzene  in  solution  has  been  detected 
by  formal  addition  reactions  to  both  benzvalene  and  dewar  benzene.  Two  mechan- 
isms have  been  proposed  to  account  for  the  benzene  photoreactions.  The  Bryce- 
Smith  and  Longuet-Higgins  mechanism  is  based  on  the  correlation  of  the  lowest 
benzene  singlet  and  triplet  states  with  singlet  or  triplet  states  of  diradicals 
which  can  lead  to  the  observed  products.   The  Farenhorst  mechanism  is  based  on 
a  trans-benzene  intermediate  which  has  pi-orbitals  topologically  equivalent  to 
a  Moebius  ring.  In  rigid  media  at  77°  K  benzene  forms  substituted  hexatrienes 
by  a  monophotonic  process  and  leads  to  the  formation  of  solvent  radicals  and 
hydrogen  by  a  biphotonic  process. 

BIBLIOGRAPHY 

1.  N.  J.  Turro,  "Molecular  Photochemistry,"  W.  A.  Ben(iamin,  Inc.  ,  New  York, 
N.  Y. ,  I965. 

2.  W.  A.  Noyes  Jr.  and  I.  linger,  Adv.  in  Photochemistry,  hs   h$   (1966). 

3.  F.  Mellows  and  S.  lipsky,  J. '  Phys.  Chem. ,  JQ,   i+076  ( 1966). 

k.      J.  K.  Foote,  M.  H.  Mallon,  and  J.  N.  Pitts  Jr.,  J.  Am.  Chem.  Soc. ,  88, 
3698  (1966). 

5.  H.  R.  Ward,  J.  S.  Wishnok,  and  P.  D.  Sherman  Jr.,  J.  Am.  Chem.  Soc,  89, 
162  ( 1967) . 

6.  K.  Shindo  and  S.  Lipsky,  J.  Chem.  Phys.,  4£,  2292  (I966). 

7.  J.  N.  Pitts  Jr.,  J.  K.  Foote,  and  J.  K.  S.  Wan,  Photochemistry  and  Photo- 
biology,  k,   323  (1965)0 

8.  J.  E.  Wilson  and  W.  A.  Noyes  Jr.,  J.  Am.  Chem.  Soc,  6^3,  3025  (1941). 


-262- 

9.  C.  L.  Braun,  S.  Kato,  and  S.  Lipsky,  J.  Chem.  Phys. ,  ^,  1645  (1963). 

10.  H.  J.  F.  Angus,  J.  Mo  Blair,  and  Do  Bryce- Smith,  J.  Chem.  Soc,  200.3  (i960). 

11.  Jo  Mo  Blair,  and  D.  Bryce-Smith,  Proc.  Chem.  Soc,  287  (1957). 

12.  M.  McDaniel,  U.  of  I.  Organic  Seminar,  59  (Summer  1965) . 

13.  I.  Haller,  J.  Am.  Chem.  Soc,  88,  2070  (1966). 

14.  G.  Camaggi,  F.  Gozzo,  and  G.  Cevidalli,  Chem.  Comm.  ,  313  (1966). 

15.  D.  M.  Lemai  and  J.  P.  Lokensgard,  J.  Am.  Chem.  Soc,  88 ,  593^  (1966). 
I60  K.  E.  Wilzbach  and  L,  Kaplan,  J. '  Am.  Chem.  SocM  87,  4~004  (I965). 

17.  H.  Go  Viehe,  Angew.  Chem.  Internat.  Ed. ,  4,  746  (1965). 

18.  L,  Kaplan,  J.  Ritscher,  and  K.  E.  Wilzbach,  J.  Am.  Chem.  Soc,  88,  2881 
(1966). 

19.  E.  Farenhorst  and  A.  F.  Bickel>  Tetrahedron  Letters.,  59H  (1966). 

20.  K.  E.  Wilzbach  and  L.  Kaplan,  J.  Am.  Chem.  Soc,  88,  2066  (1966). 

21.  D.  Bryce-Smith,  A,  Gilbert,  and  B.  H.  Orger,  Chem,  Comm. ,  512  (1966). 

22.  R.  Srinivasan  and  K.  A.  Hill,  J.  Am.  Chem,  Soc,  87,  4653  (1965). 

23.  D.  Bryce- Smith  and  A.  Gilbert,  Chem  Comm.,  643  (i960). 

24.  G.  Koltzenburg  and  K.  Kraft,  Tetrahedron  Letters,  389  (1966) . 

25.  G.  Koltzenburg  and  K.  Kraft,  Angew.  Chem.  Internat.  Ed.,  hj   981  (1965). 

26.  D.  Bryce-Smith  and  J.  £.  Lodge,  J.  Chem.  Soc,  2675  (I962) 

27.  Do  Bryce-Smith  and  A.  Gilbert,' J.  Chem.  Soc,  918  (1965). 

28.  D.  Bryce-Smith  and  M.  A.  Hems,  Tetrahedron  Letters,  1895  (1966). 

29.  J.  S.  Bradshaw,  Tetrahedron  Letters,  2039  (1966). 

30.  E.  Grovenstein  Jr.  and  D.  V.  Rao,  Tetrahedron  Letters,  148  (1961). 

31.  D.  Bryce-Smith  and  J.  E.  Lodge,  J.  Chem.  Soc,  695  (1963). 

32.  D.  Bryce-Smith  and  J.  E,  Lodge,  Proc.  Chem.  Soc,  333  (196l). 

33.  J.  G.  Atkinson,  D.  Eo  Ayer,  G.  Biichi,  and  E,  W.  Robb,  Jo  Am.  Chem.  Soc, 
85,  2257  (1963). 

34.  D.  Bryce-Smith  and  H.  C.  Longuet-Higgins,  Chem.  Comm.,  593  (1966). 
35 »  S.  F.  Mason,  Quarterly  Reviews,  15,  287  (1961). 

36.  H.  C.  Longuet-Higgins  and  E.  W.  Abrahamson,  J.  Am.  Chem.  Soc,  87,  2045 

(1965). 

37.  G.  W.  King  and  E.  H.  Pinnington,  J.  Mol.  Spec  try. ,  15,  39^+  (1965)* 
380  E.  Farenhorst,  Tetrahedron  Letters,  6465  (1966). 

39.  E.  E.  van  Tamelen,  Angew.  Chem.  Internat.  Ed.,  4,  738  (1965). 

40.  Ro  B.  Woodward  and  R.  Hoffmann,  J.  Am.  Chem.  Soc,  87,  395  (1965). 

41.  E.  Hielbronner,  Tetrahedron  Letters,  1923  (1964). 

42.  G.  E.  Gibson,  N.  Blake,  and  M.  Kalm,  J.  Chem.  Phys.,  21,  1000  (1953)- 

43.  E.  J.  Anderton,  H.  T.  J.  Chilton,  and  G.  Porter,  Proc  Chem.  Soc,  352 
(I960). 

44.  I.  Norman  and  G.  Porter,  Proc  Roy.  Soc,  A230,  399  (1955). 

45.  E.  Migirdicyan,  J.  Chim.  Phys,,  6j>v  520  ( 1966) . 

46.  E.  Migirdicyan  and  S.  Leach,  J.  Chim.  Phys.,  _5j3,  762  (1961). 

47.  S.  Leach,  E.  Migirdicyan,  and  L.  Grajcar,  J.  Chim.  Phys.,  56,  748  (1959). 
48o  E.  Migirdicyan  and  S.  Leach,  J.  Chim.  Phys.  ,  58,  4l6  (I96TJ". 

h-9.     E.  Misrirdicyan ,  J.  Chim.  Phys.  .  6j5,  535  (1966XT 

50.  V.  G.  Vinogradova,  B.  N.  Shelimo^rLV.Fok,  and  V,  V,  Voevodskii,  Dokl.  Akad. 
Nauk.  S.  S.  S.  R.  ,  154,  i88  (1964). 

51.  B.  N.  Shelimov.  N.  V.  Fok,  V.  V.  Voevodskii,  Dokl.  Akad.  Nauk.  S.  S. 
S.  R. ,  144,  59o  (1962). 

52.  B.  Brocklehurst ,  W.  Ao  Gibbons,  F.  T.  Lang,  G.  Porter,  and  M.  I.  Savadatti, 
Trans,  Faraday  Soc,  62,  1793  (1966). 

53.  E.  Migirdicyan,  J.  Chim.  Phys.,  6^,  5^3  (1966). 

54.  T.  S.  Godfrey  and  G.  Porter,  Trans.  Faraday  Soc,  62,  7  (1966). 

55.  L.  Kaplan  and  K.  E.  Wilzbach,  J.  Am.  Chem.  Soc,  '8£,  1030  (1967). 

560   K,  E.  Wilzbach,  J.  S.  Ritscher,  and  L.  Kaplan,  J.  Am.  Chem.  Soc,  8g,  1031 

( 1967) • 
57.  D.  J.  Patel,  M.  E.  H.  Howden,  and  J.  D.  Roberts,  J.  Am.  Chem.  Soc,  85, 

3218   ( 1963) 0 

560      w.    A.    JMoyes  Jr,    and  D.    A.    Barter,    J.    Chem.    Phys.,    46,   674  (I967). 


-263- 
THE  FHOTODIMERIZATION  OF  THYMINE 


March  1.6,  1967 


Reported  by  Sheldon  A.  Schaffer 

INTRODUCTION 

It  has  long  been  hypothesized  that  ultraviolet  (UV)  radiation  damage  to  micro- 
organisms has  involved  alterations  in  the  deoxyribonucleic  acids  (DNA)  of  the 
organism,1  but  until  fairly  recently  a  molecular  interpretation  of  the  radiation 
damage  was  unavailable.  Since  the  UV  spectrum  of  DNA  is  due  almost  entirely  to  the 
absorptions  of  its  purine  and  pyrrolidine  base  components,  most  of  the  investigations 
of  UV  photodamage  have  involved  studies  of  the  photochemistry  of  the  most  common 
purines,  adenine  (I)  and  guanine  (II),  and  the  most  common  pyrimidines,  thymine  (III), 
cytosine  (IV)  and  uracil  (V)  (Note  that  thymine  is  found  only  in  DNA  and  uracil  is 
found  only  in  ribonucleic  acids  (RNA)). 


NHp 


H=K 


NHo 


0 


II 


IT 
H 

IV 


HN 


O^^N- 
H 

V 


Early  studied2"4  indicated  that  the  pyrimidine  bases  were  much  more  sensitive 
to  UV  radiation  than'  the  purines,  and  therefore  most  investigations  have  been 
directed  toward  the  photochemistry  of  the  pyrimidines. 

ISOLATION  OF  THE  THYMINE  PHOIODIMER 

The  first  photoproduct  isolated  from  UV  irradiation  of  solutions  of  pyrimidines 
was  the  hydration  product  of  uracil,  5,6-dihydro=.6=hydroxyuracil.2j,5-?6  Berends  and 
co-workers7  pointed  out  that  the  slow  reversible  hydration  of  uracil  could  not 
account  for  the  rapid  lethal  effect  of  UV  radiation  on  bacteria.  Their  investigations 
showed  that  besides  the  reversible  hydration  of  uracil  an  irreversible  reaction  takes 
place  when  the  irradiation  is  carried  out  on  an  aqueous  solution  at  neutral  pH. 
Later  work  on  the  irradiation  of  pyrimidine  bases  in  frozen  aqueous  solutions  showed 
that  thymine,  which  exhibits  only  weak  susceptibility  to  UV  radiation  in  liquid 
aqueous  solutions,  was  highly  sensitive  to  such  radiation  in  frozen  solutions. 
Thus,  irradiation  of  a  frozen  aqueous  solution  of  thymine  with" a  4w  germicidal  lamp 
(with  maximum  intensity  at  25^  mu)  for  15  minutes  resulted  in  a  kOfo   loss  of  thymine 
absorptivity  at  264  mu.  Uracil  under  the  same  conditions  lost  only  6$  (corrected 
for  the  formation  of  reversible  photoproduct)  of  its  absorptivity  at  260  m^i,  while 
cytosine  was  totally  unaffected.  The  reason  for  the  not  too  obvious  experiment  of 
irradiation  of  frozen  solutions  of  pyrimidines  was  that  DNA  in  aqueous  solution  is 
known9  to  be  surrounded  by  an  ordered  lattice  of  water,  and  it  was  thought  that 
frozen  aqueous  solutions  of  the  pyrimidine  bases  would  approximate  the  native  state 
of  DNA  better  than  liquid  solutions  and  thus  provide  a  better  model  for  the  inves- 
tigations  of  the  photochemistry  of  DNA. 


Irradiations  of  DNAX  and  apurinic  acid   (DNA  which  has  undergone  a  mild  acid 
hydrolysis  to  remove  all  the  purine  bases)  gave  products  chromatographicaliy 
identical  to  the  so-called  "first  irreversible  product"7  obtained  from  the  irradia- 
tion of  thymine  in  frozen  solutions. 

Beukers  and  Berends12^13  reported  the  first  physical  data  on  the  thymine  photo- 
product. Their  compound  gave  an  elemental  analysis  for  CsHeO^T^,  the  molecular 
formula  of  thymine,  thus  showing  that  it  was  either  a  dimer  or  polymer  of  thymine. 
A  molecular  weight  determination  by  the  method  of  isothermic  distillation14  gave 
values  between  24-0  and  270  indicating  that  the  molecule  was  a  dimer  of  thymine 
(calc.  mol.  wt.  ■  252).  The  compound  lacked  UV  absorption  at  264  mu,  indicating 
the  lack  of  the  C5-Cs  double  bond  of  thymine,  and  its  infrared  spectrum  showed  weak 
peaks  at  960  cm  -1  which  were  taken  to  be  characteristic  of  a  cyclobutane  ring  system 


■cut- 


(in  the  light  of  recent  evidence16  this  assignment  is  somewhat  in  doubt).  The  nmr 
spectrum17  (DMSO»d6)  showed  signals  (relative  to  external  water)  at  -326  (2H, 
singlet,,  -W)  ,   -179  (2H*  singlet,  -NH)  ,  4±3   (2H,  singlet,  -CH)  ,  and  +185  cps  (6H, 
singlet,  C-CH3) .   On  the  basis  of  these  data  Beukers  and  Berends13  suggested 
structure  VI  for  the  thymine  photodimer.  Wulff  and  Fraenkel17  pointed  out  that 
there  are  four'  structures,  VI-IX,  which  fit  the  data  for  the  thymine  photodimer. 


CH3  H 


H 


cAn 


ss 


%/ 


H  CH3 
0 

VI 


H 


CH3  CH3 


M 


/ff 


0 


H   H 
H        H 

VIII 


nh 


0        H 


M 


^s 


.  H  CH; 
IX 


•nh 


They  also  showed  that  molecular  models  indicate  that  only  the  cis-syn  Isomer  (VII) 
could  reasonably  be  expected  to  be  formed  from  the  dimerization  of  adjacent  thymines 
on  the  same  DNA.  strand,  and  therefore  the  cis-syn  structure  should  be  assigned  to  the 
dimer  isolated  from  DBA.  An  enzymatic  digestion  of  a  sample  of  UV  irradiated  DNA 
gave  various  trinucleotides  of  the  type  pXpTpT17  (where  X  is  any  one  of  the  four 
bases  occurring  in  DMA  and  TpT  is  the  thymine  dimer  joined  by  the  normal  phospho- 
diester  bond  between  the  ribose  moieties  of  the  thymine  nucleosides) .  This  implies 
that  at  least  some  of  the  thymine  dimers  are  formed  between  adjacent  thymines  on 
the  same  DNA  strand..  Furthermore,  they  suggested  that  one  of  the  trans  isomers 
(VIII  or  IX)  might  be  formed  in  cross-linked  dimers  between  two  DNA  strands.18*19 

TEE  STRUCTURE  DETERMINATION  OF  THE  THYMINE  PHOTODIMER 

In  1965  Blackburn  and  Davies20  offered  the  first  chemical  proof  of  structure 
for  the  thymine  dimer.  They  found20*21  that  the  treatment  of  the  thymine  dimer 
(formed  by  the  irradiation  of  a  frozen  solution  of  thymine)  with  10$  NaOH  gave  a 
disodium  salt  believed  to  have  structure  X  because  it  exhibited  a  maximum  UV 
absorption  at  237  eu«  The  monoanion  of  thymine  has  its  maximum  absorption  at  23C  rnu. 
Increasing  the  NaOH  concentration  to  k-Cffo   resulted  in  the  formation  of  a.  white 
precipitate  which  showed  only  end  absorption  in  the  UV  and  whose  IR  spectrum  showed 
bands  at  1310,  I56O  (COi)  ,  1655  ( -NHCQNH2) ,  315 0,  3280  and  3370  cm"1  (amide  NH's). 
This  material  was  assinged  the  structure  of  the  disodium  salt  of  the  bisureido  acid 
(XI).  Such  ureido  acids  are  known  to  be  formed  from  5^6-dihydrouracils  by  treatment 

CH3  CH3 
Na  02&„  V        17- .  €  02Na 


NaOH 


°^<   H 

H 


H 


NaOH 


0 


H2N 


^Sr^S 


H 


H 


1 


0 

H 


X 


XI 


with  base.22*23  When  XI  was  dissolved  in  water  it  reverted  to  X,  and  X  could  be 
converted  back  to  the  thymine  dimer  by  treatment  with  acid.   The  facile  recyclization 
of  XI  in  water  is  in  contrast  to  the  behavior  or  the  p-ureido  acid  salts  or  the 


-2op- 


5,6-dihydrouracils  which  recyclize  only  in  the  presence  of  acid22-"23  and  is 
probably  due  to  the  all  cis  arrangement  of  groups  on  the  cyclobutane  ring. 

Treatment  of  XI  with  bromine  (an  attempted  Hoffmann  rearrangement)  resulted  in 
the  formation  of  a  rearranged  product,  isomeric  with  the  thymine  dimer.  This  com- 
pound exhibited  only  UV  end  absorption  in  both  alkaline  and  neutral  solutions 5,  its 
IR  spectrum  showed  bands  at  iGjk   (WHCGNH2) ,  1723,  1770  (COOTCO) ,  32^0,  337Q  and 
3^20  cm'1  (amide  NH's) j  the  nmr  spectrum  (in  trifluoroacetic  acid)  gave  signals  at 
tO.02  (IB,  singlet,  W)  ,   2.55  (IH,  singlet,  MH)  ,  h.96   (IE,  doublet,  J  =  9  cps, 
cyclobutane  hydrogen),  5.55  (XH,  quartet,  J±   =  9  cps,  J2   ■  2  cps,  cyclobutane 
hydrogen),  8.38  and  8. if  5  ( 3H  each,  singlets,  C-CH3)  .  The  weak  coupling  of  the 
signal  at  T5.55  could  be  eliminated  by  running  the  spectrum  in  deuterated  TFA.  In 
DMSO  solution  an  additional  broad  band  appeared  at  t2„9  (2H,  33H)  .   On  the  basis  of 
these  data,  Blackburn  and  Davie s  offered  structure  XII  as  the  rearranged  compound. „ 

9  CH3  H  H 


The  chemical  proof  offered  to  support  structure  XII  is  as  follows :  treatment 
of  the  rearranged  product  with  nitrous  acid  results  in  the  loss  of  the  carbamoyl 
moiety  with  a  simultaneous  collapse  in  the  nmr  AB  pattern  to  an  A2   system.  This 
requires  the  location  of  the  carbamoyl  group  on  one  of  the  nitrogens  of  the 
imidazole  ring  to  provide  the  molecular  dissymmetry  which  makes  the  cyclobutane 
hydrogens  magnetically  non-equivalent. 

Pyrolysis  of  XII,  or  its  decarbamoylated  derivative,  gave  2,3-dimethylmaleimide, 
XIII,  and  2-imidazolone,  XIV.  The  production  of  XIII  shows  that  both  the  rearranged 


xrv 

product  and  the  thymine  dimer  had  methyl  groups  on  vicinal  carbons,  and  therefore 
the  dimer  must  have  one  of  the  syn  structures  VII  or  VIII.  A  syn  structure  is  also 
indicated  by  the  Jc13H  satellite  spectrum  in  the  nmr  of  the  dimer.24 

The  formation  of  2-imidazolone  (XIV)  from  the  pyrolysis,  taken  with  the  nmr 
data  of  XII,  confirm  that  the  cyclobutane  hydrogens  of  the  dimer  were  vicinal.  The 
coupling  constant  of  9  cps  for  these  hydrogens  is  consistent25  with  their  being  cis 
oriented,  but  their  orientation  cannot  be  rigorously  proved  in  this  way  as  the 
Karplus  relation26  is  not  strictly  applicable  to  systems  containing  strong  electro- 
negative groups.27  Further  proof  for  the  cis  ring  junction  was  based  on  the  failure 
of  similar  systems  to  give  trans  fused  rings28  and  the  fact  that  only  cis  fused 
dimethyl  bicyclo[4.2.0]octane-7,8-dicarboxylate  gives  cyclohexene  on  pyrolysis.29 

Chemical  proof  of  structure  for  the  DMA  derived  thymine  photodimer  was  obtained 
by  Blackburn  and  Davie s30  by  growing  E^  coli  on  thymidine -6T,  irradiating  the 
bacteria  with  UV  light,  and  isolating  the  photoproducts  chromatographically.  The 
photoproducts  were  mixed  with  carrier  thymine  "ice "■dimer"  and  repeatedly  recrystallized 
without  loss  of  any  radioactivity.  Treatment  of  an  NaOH  solution  of  the  dimer  with 
bromine  gave  the  rearrangement  product  XII  with  ^Ojo   of  the  initial  activity.  Since 
Blackburn  and  Davies  feel  that  the  oxidative  rearrangement  is  only  allowed  for  the 
cis-syn  isomer,  they  conclude  that  the  DHA.  derived  dimer  must  then  have  structure  VII. 


:66~ 


STUDIES  ON  SUBSTITUTED  THYMINES 

Studies  on  the  UV  Induced  photodimerizations  of  substituted  thymines 9X7 s 31 
thymidine,32  and  thymidyl  (31  "*"  5')  thymidine33  have  shown  that  more  than  one 
cyclobutane  type  thymine  dimer  may  be  formed.  The  identification  of  the  photodimers 
from  substituted  thymines  has  been  presented  by  Blackburn  and  Davies31  and  rests  on 
the  following  evidence ;  irradiation  of  a  frozen  solution  of  1,3-dimethylthymine 
leads  to  two  cyclobutane  type  photodimers17  the  higher  melting  of  which  was  shown 
to  be  identical  with  the  tetramethyl  derivative  of  the  thymine  "ice -dimer",17  and 
thus  has  structure  XXb.   Similarly,  1-methylthymine  gives  two  cyclobutane  photo- 
dimers when  irradiated  in  a  frozen  aqueous  solution.31  Both  of  these  photodimers 
are  alkylated  smoothly  with  dimethylsulfate  to  give  the  tetramethyl  compounds.   One 
of  these  compounds  was  shown  to  be  identical  to  the  tetramethyl  derivative  of  the 
thymine  "ice-dimer"  and  thus  the  original  compound  had  structure  XXc.  The  other 
dimethylated -derivative  of  the  1-methylthymine  dimer  was  different  from  the 
unas signed,  lower  melting  dimer  from  1,3-dimethylthymine  and  thus  represents  a  third 
isomer. 

0  0  0  0 


(a:R=R'=H) 

(b:R=R'=CH3) 

(c;R=CH3,R»=H) 


(asR=R»=H) 
(bsR=R»=CH3) 
(c:R=CH3,R»=H) 
(dsR=H>R'=CH3) 


XXI 


(a:R=R'=H) 
(b;R=R»=CH3) 


XXII 


(a;R^R5=H) 
(b;R=R«=CH3) 


Blackburn  and  Davies  pointed  out  that  1-methylthymine  exists  in  two  crystal 
modifications34  and  that  topochemical  arguments  (vide  infra)  presented  by  Stewart35 
show  that  irradiation  in  the  solid  state  of  the  more  stable  of  the  crystal  forms 
should  give  the  dimer  XIXc.   On  this  basis  the  second  dimer  of  1-methylthymine  was 
assigned  the  trans -syn  structure  XIXc.  This  implies  that  the  unassigned  dimer  from 
1,3-dimethylthymine  must  have  one  of  the  anti  structures,  either  XXIb  or  XXIIb. 
An  assignment  cannot  be  made  for  this  compound  at  this  time. 

The  structure  of  the  photodimer  of  3~niethylthymine  was  deduced  from  the  partial 
methylation  of  the  thymine  "ice-dimer"  to  give  a  symmetrical  dimethyl  derivative 
(determined  by  nmr)  which  undergoes  photoreversion  (vide  infra)  to  3-methylthymine 
and  is  identical  in  its  spectroscopic  and  chromatographic  properties  with  the  3  = 
methylthymine  dimer.  This  establishes  structure  XXd  for  the  3-ine'thylthymine  dimer. 

Weinblum  and  Johns32  have  carried  out  a  similar  treatment  to  prove  the 
structures  of  four  different  thymine  dimers  obtained  from  thymine,  thymidine,  DNA  and 


J& 


u 


thymidyl  (31  •»,5")  thymidine. 
MECHANISM  OF  THE  FHOTODIMERIZATION 

Phot odimerizat ions  of  the  type  undergone  by  thymine  have  been  observed  for 
other  pyrimidine  bases.36"38  Waeker38  has  attempted  to  correlate  the  tendency  of 
pyrimidines  to  undergo  a  phot odimerizat ion  with  the  polarity  of  the  C5-C6  double 
bond.  Mantione  and  Pullman39  have  pointed  out  that  this  argument  considers  only 
the  ground  state  properties  of  the  molecules  and  does  not  take  into  account  the 
role  of  the  phot oactivat ion.  An  alternate  explanation  based  on  the  unpaired 
electron  density  in  the  C5-C6  double  bond  in  the  excited  triplet  state  of  the 
pyrimidines  was  offered  by  Mantione  and  Pullman  to  explain  the  relative  order  of 
dimerization  of  the  pyrimidine  bases. 

The  calculations  of  Mantione  and  Pullman,  presented  in  Table  1,  were  made 
using  simple  Mckel  molecular  orbital  theory.  They  acknowledge  the  fact  that  this 
simple  theory  does  not  distinguish  between  excited  state  singlets  or  triplets,  but 
they  assume  that  the  calculations  nearly  represent  the  thymine  triplet. 

Table  I39 

Cone,  of  odd  electrons  in 
Ability  to  the  C5~C6  bond  in  the  excited 

Compound  dimerize  triplet  state. 

Orotic  acid  Good  1.120 

Thymine  "  1.207 

Uracil  "  1.252 

6~Methyluracil  "  1.208 

Isocytosine  Fair  1.159 

5-Aminouracil  "  1.053 

5-Methylcytosine  Weak  O.879 

Cytosine  "  O.857 

2-Thiothymine  Ml  0. 95  3 

5-Nitrouracil  "  O.639 

6  -Azathymine  1 .  142 

The  data  in  Table  1  show  good  qualitative  agreement  with  experimental  fact 
(except  for  those  compounds  which  contain  an  additional  heteroatom)  and  while  the 
results  indicate  that  high  electron  spin  density  in  the  C5-C6  bond  promotes 
dimerization ^  they  do  not  prove  that  the  triplet  excited  state  is  a  precursor  to  the 
photodimer. 

The  reason  that  the  triplet  state  is  favored  by  Mantione  and  Pullman  seems  to 
be  linked  to  the  observation  by  Beukers  and  Berends7^37  that  oxygen  and  paramagnetic 
salts  decrease  the  production  of  photoproducts  in  the  irradiation  of  liquid  solu- 
tions of  uracil  and  increase  the  amount  of  phot ore version  of  previously  formed 
dimers.  It  is  known  that  paramagnetic  substances  increase  the  number  of  singlet  ■> 
triplet  transitions  allowed  by  increasing  the  spin-orbit  coupling  in  the  molecules 
undergoing  the  transitions.40 

Recently  Lamola  and  MLttal41  have  studied  the  dimerizations  of  thymine  and 
uracil  in  acetonitrile.  They  found  that  the  dimerization  of  thymine  could  be 
completely  quenched  by  the  addition  of  isoprene  (a  triplet  quencher  for  thymine). 
The  uracil  dimerization  was  only  partially  quenched  by  isoprene.  This  work  implies 
that  the  dimerization  of  thymine  in  acetonitrile  proceeds  entirely  through  a 
triplet  state  while  the  dimerization  of  uracil  proceeds  partially  through  a  triplet. 
Chromatographically  the  dimer  formed  from  thymine  in  acetonitrile  is  different  from 
the  thymine  "ice -dimer"  and  was  tentatively  assigned  a  trans -anti  structure. 

Lamola  and  Mittal  also  investigated  the  UV  irradiation  of  uracil  in  liquid 
aqueous  solution.  The  major  products  formed  from  such  irradiations  are  the  uracil 
hydration  product  and  uracil  dimers.  Upon  adding  2,4- hexadienol  (HDE),  an  H20 
soluble  triplet  quencher,  to  the  solution,  they  observed  an  increase  in  the  ratio  of 
hydrate  to  dimers.  They  could  identify  two  dimers  among  the  photoproducts,  one  of 
which  was  identical  with  the  uracil  "ice -dimer"  while  the  other  was  a.  new  dimer. 


-268- 

In  the  presence  of  HDE  the  ratio  of  "ice-dimer"  to  new  diraer  increased.  Thus  it 
appears  that  the  new  uracil  dimer  is  formed  through  a  triplet  species  while  the 
uracil  "ice -dimer"  and  the  hydration  product  are  not. 

An  analogous  situation  occurs  in  the  dimerization  of  coumarin42  where  singlet 
state  excited  coumarin  is  thought  to  lead  to  a  cis~"head -to-head"  dimer  while 
triplet  excited  coumarin  gives  the  trans -"head -to-head"  and  "head -to-tail"  dimers. 

D6nges  and  Fahr43  have  recently  studied  the  structure  of  the  uracil  "ice-dimer" 
and  have  assigned  to  it  a  cis - "head -to -head "  structure.  Thus  it  appears  that  the 
dimerization  of  uracil  in  liquid  aqueous  media  is  controlled  in  the  same  way  as  the 
dimerization  of  coumarin,  and  the  dimerization  of  thymine  in  liquid  solutions  may 
be  under  this  same  control. 

Excited  state  triplets  have  been  observed  both  by  optical44  and  ESR45546 
measurements  on  UV  irradiated  thymine  and  DMA  samples  in  H20 methylene  glycol  glasses 
at  77°K.  The  triplet  species  found  in  all  the  cases  studied  could  be  assigned  to 
the  conjugate  base  of  thymine. 

Free  radicals  have  also  been  in  UV  irradiated  samples  of  thymine  and  DNA47=49 
at  77°K,  and  here  again  the  radical  could  be  assigned  to  a  thymine  species,  XXIII. 

0 


HI 


'CH3 


^H 

H 

XXIII 

These  data  show  that  the  thymine  triplet  and  free  radical  are  important  species 
present  in  UV  irradiated  samples  of  thymine  and  DNA,  but  they  do  not  implicate 
either  species  directly  in  the  photodimerization  reaction. 

CONTROL  OF  THE  STEREOCHEMISTRY  OF  THE  THYMINE  "ICE-DIMER" 

It  was  mentioned  above  that  one  of  the  isomeric  dimers  obtained  from  the 
irradiation  of  a  frozen  solution  of  l~methylthymine  was  assigned  its  structure  on 
the  basis  of  the  crystal  structure  of  1-methylthymine.35  Schmidt  and  co-workers50 
have  shown  that  the  photodimerizations  of  olefins  in  the  solid  state  depends  on  the 
orientation  of  the  molecules  in  the  crystal  lattice  and  occur  with  a  minimum  amount 
of  atomic  and  molecular  motion.  Among  the  examples  of  this  topochemical  control  is 
the  dimerization  of  trans -cinnamie  acid.  This  compound  exists  in  two  crystal  forms , 
a  and  p.  Irradiation  of  the  more  stable,  a,  form  gives  only  a-truxillic  acid,  XXIV, 
while  irradiation  of  the  p  form  leads  only  to  p-truxinic  acid,  XXV.51'52 


COaH 


Fh 


a„form_ 
C02H 


Jb£orm 
hv 


CO^H 


Wang53"55  has  suggested  that  thymine  in  frozen  aqueous  solutions  exists  in  a 
high  state  of  aggregation,  being  excluded  from  the  ice  crystals  and  forced  into 
clusters  of  solid  thymine.  The  UV  spectrum  of  thymine  in  a  frozen  solution56  most 
nearly  resembles  the  spectrum  of  solid  thymine  and  helps  to  confirm  Wang's  theory. 


-2o9- 

If  thymine  does  exist  in  a  high  state  of  aggregation  in  a  frozen  solution, 
then  the  arguments  of  Schmidt  and  his  co-workers55  would  require  that  the  stereo- 
chemistry  of  the  dimer  formed  from  the  UV  irradiation  of  the  frozen  solution  should 
be  determined  by  the  crystal  structure  of  solid  thymine. 

An  X-ray  analysis  of  thymine  monohydrate57  shows  it  to  have  a  structure  which 
can  be  represented  as  in  Figure  1.   If  we  restrict  the  photodimerization  of  thymine 
to  that  course  which  requires  the  least  amount  of  molecular  movement,  it  is  clear 
that  the  cis-syn  isomer  will  be  the  only  dimer  formed.  Evidence  that  thymine  exists 

Figure  1 


.0 

5-f 

.-  -H-^ 

0= 

O 

H 

u 

-CH3 

0= 

Q 

H 

-ch3  y 

A\ 

H            H 

K 


H3C-U 

K 


Crystal  Structure  of  Thymine  Monohydrate58 

as  the  monohydrate  in  frozen  aqueous  solutions  comes  from  experiments  by  Wang58 
where  thin  films  of  thymine  were  irradiated  under  conditions  of  varying  humidity. 
Table  2  illustrates  the  importance  of  humidity  on  the  thymine  dimerization,  and 
Wang  regards  these  data  as  proof  that  thymine  monohydrate  is  the  species  undergoing 
the  photodimerization. 

Table  2 

Humidity  (#) 

Time  (hr)  98         71         30        PgOs 

1  15.0      17.3      11.4      7.5 

2  31.0      30.9      21.0      16.3 

3  55.0      55.O      27.O      21.7 

io  Dimerization  of  Thymine  Under  Conditions  of 
Varying  Humid ity( 58) . 

FHOTOREVERSION 

One  of  the  major  reasons  for  the  interest  shown  in  the  photodimerization  of 
thymine  is  its  connection  with  the  biological  inactivation  of  bacteria  and  phages 
by  UV  light. 19,59~63  It  has  been  well  established  that  some  of  the  effects  of  UV 
radiation  may  be  reversed  by  some  photochemical  process,64  and  one  of  the  first 
properties  of  the  thymine  photodimer  to  be  determined  was  its  ability  to  revert 
back  to  thymine  when  it  was  reirradiated  with  UV  light  in  a  liquid  aqueous  solution. 
In  fact,  it  was  shown  that  irradiation  of  thymine  led  to  the  establishment  of  a 
photostationary  state  between  the  photodimer  and  the  monomer.65  At  short  wavelengths 
the  equilibrium  favors  the  monomer  while  long  wavelengths  favor  the  dimer. 


It  has  also  been  found  that  a  crude  enzyme  extract  from  yeast  catalyzes  the 
splitting  of  the  thymine  dimer63'66  or  the  splitting  out  of  the  dimer  from  the  DNA 
chain.67 

The  thymine  dimer  now  has  a  firmly  established  place  in  the  area  of  photo- 
damage  and  photorepair  in  biological  systems. 

BIBLIOGRAPHY 

1.  M.  R.  Zelle  and  A.  Hollaender  in,  "Radiation  Biology,"  ed.  by  A.  Hollaender, 
McGraw,  New  York,  1955^  vol.  2. 

2.  R.  L.  Sinsheimer  and  R.  Hastings,  Science,  110,  525  (1949). 

3.  C.  E.  Carter,  J.  Am.  Chem.  Soc,  72,  I835  (1950). 

4.  E.  Christensen  and  A.  C„  Giese,  Arch.  Biochem.  Biophys.,  £1,  208  (195*0  . 

5.  R.  L.  Sinsheimer,  Radiation  Research,  1,  5O5  (1954). 

6.  A.  M.  Moore  and  C.  H.  Thomson,  Science  122,  594  (1955). 

7.  A.  Rorsch,  R.  Beukers,  J.  Ijlstra,  and  W.  Berends,  Rec.  trav.  chim. ,  77,  423 
(1958). 

8.  R.  Beukers,  J.  Ijlstra,  and  W.  Berends,  Rec.  trav.  chim.,  77,  729  (I958). 

9.  B.  Jacobson,  J.  Am.  Chem.  Soc,  77,  2919  (1955). 

10.  R.  Beukers,  J.  Ijlstra,  and  W.  Berends,  Rec.  trav.  chim.,  79,  101  (i960) . 

11.  R.  Beukers,  J.  Ijlstra,  and  W.  Berends,  Rec.  trav.  chim.,  7H,  247  (1959). 

12.  R.  Beukers  and  W.  Berends,  Biochim.  Biophys.  Acta,  4l,  550  (i960). 

13.  R.  Beukers  and  W.  Berends,  Biochem.  Biophys.  Acta,  ¥9,  101  (1961). 

14.  R.  Signer,  Ann.,  4j8,  2h6   (19.30). 

15.  L.  W.  Barrison,  J.  Chem.  Soc,  l6lh   (I95I). 

16.  H.  E.  Ulery  and  J.  R.  McClenon,  Tetrahedron,  1£,  749  (I963). 

17.  D.  L.  Wulff  and  G.  Fraenkel,  Biochim.  Biophys.  Acta.,  £1,  332  (I961). 
17b.  H.  Dellweg  and  A,  Wacker,  Photochem.  Photobiol. ,  5,  119  (I966) . 

18.  D.  Shugar  and  G.  Baranowska,  Nature,  185 ,  33  (i960). 

19.  J.  Marmur  and  L.  Grossman,  Proc.  Nat.  Acad.  Sci. ,  47,  778  (196l). 

20.  G.  M.  Blackburn  and  R.  J.  H.  Davies,  Chem.  Commun. ,  215  (1965). 

21.  G.  M.  Blackburn  and  R.  J.  H.  Davies,  J.  Chem.  Soc.  (C),  2239  (1966). 

22.  R.  D.  Blatt,  J.  K.  Martin,  J. Mo  T«  Ploesser,  and  J.  Murray,  J.  Am.  Chem.  Soc., 
76,  3663  (195^). 

23.  D.  Keglevic  and  A.  Kornhauser,  Croat.  Chem.  Acta,  31,  47  (1959). 
2k.     R.  Anet,  Tet.  Letters,  3713  (1965), 

25.  R.  Steinmetz,  W.  Hartmann,  and  G.  0.  Schenk,  Ber. ,  £8,  3854  (1965). 

26.  Mo  Karplus,  J.  Chem.  Phys.,  30,  11  (1959). 

27.  M.  Karplus,  J.  Am.  Chem.  Soc.,  8£,  2870  (I963). 

28.  E.  R.  Buchman,  A.  0.  Reims,  T.  Skir,  and  M.  J.  Schatter,  J.  Am.  Chem.  Soc,  64, 
2696  ( 19^2) . 

29.  P.  deMayo,  S.  T.  Reid,  and  R.  W.  Yip,  Can.  J.  Chem.,  42,  2828  (1964). 

30.  G.  M.  Blackburn  and  R.  J.  H.  Davies,  Biochem.  Biophys.  Res.  Commun.,  22,  704 
(1966).  — 

31.  G.  M.  Blackburn  and  R.  J.  H.  Davies,  J.  Chem.  Soc.  (C),  1342  (1966). 

32.  D.  Weinblum  and  H.  E„  Johns,  Biochim.  Biophys.  Acta,  114,  450  (1966). 

33»  H.  E.  Johns,  M.  L.  Pearson,  J.  C.  LeBIanc,  and  L.  W.  Helleiner,  J.  Mol.  Biol., 
9,   503  (1964). 

34.  K.  Hoogsteen,  Acta  Cryst.,  16,  28  (1963). 

35.  R.  F.  Stewart,  Biochim.  Biophys.  Acta,  7^,  129  (I963). 

36.  D.  Shugar  and  E.  Chargaff  in  "The  Nucleic  Acids,"  ed.  by  J.  N.  Davidson, 
Academic  Press,  New  York,  i960,  vol.  3,  p.  40. 

37.  R.  Beukers,  J.  Ijlstra,  and  W.  Berends,  Rec.  trav.  chim.,  78,  883  (1959). 

38.  A.  Wacker  in  "Progress  in  Nucleic  Acids  Research,"  ed.  by  J.  N.  Davidson  and 
W.  E.  Cohn,  Academic  Press,  New  York,  1963,  vol.  1,  p.  369. 

39.  J.  M.  Mantione  and  B.  Pullman,  Biochim.  Biophys.  Acta,  91,  387  (1964) „ 

40.  N.  J.  Turro,  "Molecular  Photochemistry,"  W.  A.  Benjamin  Inc.,  New  York,  I965, 
chapter  4. 

41.  A.  A.  Lamola  and  J.  Mittal,  Science,  1^4,  1560  (1966). 

42.  G.  S.  Hammond,  C.  A.  Stout,  and  A.  A.  Lamola,  J.  Am.  Chem.  Soc,  86,  JJ03  (1964). 


~271- 

43.  K.  H.  Dtfnges  and  E.  Fahr,  Z.  Naturforsch.  ,  21b,  87  (1966). 

44.  J.  W.  Longworth,  R.  0.  Rahn,  and  R.  G.  ShuLman,  J.  Chem.  Rhys.,  4£,  2930  (1966). 

45.  R.  0.  Rahn,  R.  G.  Shulman,  and  J.  W.  Longworth,  J.  Chem.  Phys.,  4j,  2955  (1966). 

46.  R.  G.  Shulman  and  R.  0.  Rahn,  J.  Chem.  Phys.,  4£,  2940  (1966). 

47.  J.  Eisinger  and  R.  G.  Shulman,  Proc.  Nat.  Acad.  Sci.,  50,  69k   (1963)  . 

48.  P.  S.  Pershan,  R.  G.  Shulman,  B.  J.  Wyluda,  and  J.  Eisinger,  Physics,  1,  163 
(1964). 

49.  R.  G.  Shulman,  R.  0.  Rahn,  P.  S.  Pershan,  J.  W.  Longworth,  and  J.  Eisinger  in 
"Electron  Spin  Resonance  and  the  Effects  of  Radiation  on  Biological  Systems," 
National  Academy  of  Sciences-  National  Research  Council,  Washington,  D.  C. , 
1966,  p.  46. 

50.  M.  D.  Cohn  and  G.  M.  J.  Schmidt,  J.  Chem.  Soc.,  1996  (1964). 

51.  M.  D.  Cohn,  G.  M.  J.  Schmidt,  and  F.  I.  Sonntag,  J.  Chem.  Soc.,  2000  (1964). 

52.  G.  M.  J.  Schmidt,  J.  Chem.  Soc,  2014  (1964). 

53.  S.  Y.  Wang,  Nature,  190,  690  (I96I). 

54.  S.  Y.  Wang,  Photochem.  Photobiol.  ,  3,  395  (1964). 

55.  S.  Y.  Wang,  Fed.  Proc.,  24,  Part  III,  Supp.  15,  S-71  (1965). 

56.  S.  L.  Davis  and  I.  Tinoco,  Jr.,  Nature,  210,  1286  (I966). 

57.  R.  Gerdil,  Acta  Cryst.,  14,  333  (I961). 

58.  S.  Y.  Wang,  Nature,  200,~%79  (19^3). 

59„  A.  Wacker,  H.  Dellweg,  and  E.  Lodeman,  Angew.  Chem.,  73,  64  (I961). 

60.  A.  Wacker,  H.  Dellweg,  and  D.  Weinblum,  Naturwiss.,  4j_,  477  (1961). 

61.  A.  Wacker,  J.  chim.  phys.,  58,  ll4l  (1961). 

62.  R.  B.  Setlow  and  J.  K.  Setlow,  Proc.  Nat.  Acad  Sci.,  48,  1250  (1962). 

63.  D.  L.  Wulff  and  C.  S.  Rupert,  Biochem.  Biophys.  Res.  Commun. ,  7,  237  (I962). 

64.  C.  S.  Rupert  in  "Photophysiology,"  ed.  by  A.  C.  Geise,  Academic  Press, 
New  York,  1964,  p.  283. 

65.  R.  Beukers,  J.  Ijlstra,  and  W.  Berends,  Rec.  trav.  chim.,  78,  883  (1959). 
GG0     A.  A.  Lamola,  J.  Am.  Chem.  Soc.,  88,  813  (1966). 

67.  R.  B.  Setlow  and  W.  L.  Carrier,  Proc.  Nat.  Acad.  Sci.,  £1,  226  (1964). 


March  20,  1967 


_07O .. 

THERMAL  REARRANGEMENTS  OF  CYCLOHEPTATRIENES1 

Reported  By  W.  D.  Shermer 

Introduction: 

Cycloheptatrienes,  which  also  have  the  common  name  tropilidenes »  have  been  found 
to  undergo  several  thermal  and  photolytic  rearrangements.   These  isomerizations  can  be 
divided  into  two  genral  classes  -  hydride  shifts  and  skeletal  rearrangements.   The  lat- 
ter rearrangements  are  exemplified  by  the  reactions  shown  in  Figure  1.   One  example  of 
the  photochemical,  rearrangements  will  be  discussed  briefly  but  it  is  not  intended  to 
be  a  complete  review. 


^X 


Figure  1 


(Eq.  1) 


^•73 


r-^PK 

H  ^3^   \_J/     kl3 
II 


a.)  R=H 

b)  R=D 

c)  R=4> 

d)  R=0CH3 


Chemical  studies  involving  thermal  reactions  of  7-substituted  cycloheptatrienes 
have  often  been  complicated  by  the  production  of  unexpected  side-products  resulting 
from  the  isomerization  of  the  substituted  tropilidene . 2 f 3  This  isomerization  could 
conceivably  occur  by  either  an  intermolecular  or  an  intramolecular  process  or  both. 
Since  equation  2  Is  a  known  reaction,4  a  possible  intermolecular  reaction  would  be  the 
initial  formation  of  a  small  amount  of  tropylium  ion,  followed  by  hydride-ion  transfer 

from  another  cyclohepta 


(Eq.  2) 


triene  molecules.   To 

distinguish  between  the 
inter-  and  intramolecular 
processes,  ter  Borg  and 

his  coworkers  undertook  a  study  of  7-deuterocycIoheptatrIene  (l-b)  by  mass  spectroscopy. 

Table  1  gives  the  results  of  their  work.   The  invar iance  of  the  isotopic  distribution 

over  the  range  of  experimental  conditions  rules  out  the  possibility  of  an  intermolecular 

rearrangement . 

Table  1 
Distribution  of  Deuterium  After  Heating  7-Deuterocycioheptatriene 

(percentages) 
Starting  Material        Heated  Products 

98   98  121  140  140 

1700  3297  W)   270  kl£ 

6.0  6.1  5.9  6.0  6,2  6,1 

93.9         93-9  9^.1  93.9  93.7  %•■ .3 


Temperature 
Time  hours 
C7H8 
C7H7D 
C7KSD2 
C7H5D3 


Calculated  Binomial 
Distribution  Due  To 
Intermolecular  React. 

36.8 

39.2 


-.  Q 

JLO< 


4.9  +  D4,  D5,etc, 


Four  intramolecular  shifts  are  possible:   ( l)  from  C-7  to  C-l  or  C-6  (a  position) 
(2)  from  C-7  to  C-2  or  C-5  (3  position),  (3)  from  C-7  to  C=3  or  C-k   (y  position), 
and  (4)  a  random  shift  from  the  7  position  to  the  a,  &,  and  tf  positions.   To 
elucidate  which  of  the  four  possible  mechanisms  is  in  operation,  Nozoe  and  Tak- 
ahasi2  followed  the  isomerization  of  7-methoxycycloheptatriene.   The  peaks  in 
the  NMR  spectrum  due  to  the  methyl  and  C-7  protons  (6.65T  and  6.83T  respectively) 
decrease  with  time  at  103°C  and  eventually  disappear  completely.   Simultaneously, 
a  singlet  at  6.^3T(3H)  and  a  triplet  at  7»77f(2H)  appear  and  increase  in  intensity. 
On  continued  heating,  a  singlet  at  6.50r(3H)  and  a  doublet  at  7«55t(2H)  appear 
and  increase  in  intensity  and,  on  prolonged  heating,  increase  at  the  expense 
of  the  signals  at  G.kyx   and  7«77^«  As  equilibrium  nears,  two  new  signals  appear, 
a  singlet  at  6.58t(3H)  and  a  triplet  at  7«83t(2H),>  At  equilibrium  the  areas 
under  the  methyl  peaks  is  as  follows:  6. ^3t:6. 50r:6. 58t=13. 6:80.5:6.0.  Analysis 
of  the  four  proposed  mechanisms  shows  that  only  the  7  to  tf  hydride  shift  will 
give  the  sequence  of  peaks  observed  for  the  C-7  hydrogens  of  multiplet  -  triplet  - 
doublet  -  triplet. 

A.  P.  ter  Borg  and  coworkers3  made  a  kinetic  study  of  the  isomerization 
of  I-b.   It  was  noted  that  a  shift  of  a  deuterium  atom  from  the  7  position  yields 
the  same  compound  and,  consequently,  no  first  order  isotope  effects  are  observed. 
Neglecting  second  order  effects,  which  are  generally  small,  only  one  rate  constant 
is  involved  in  the  reactions.   This  constant  is  defined  as  the  rate  constant 
for  the  reaction  in  which  a  hydrogen  of  the  CH2  group  shifts.;  i.e.  k73=k37=k31 
etc.  (for  I-b  a  statistical  factor  of  1/2  is  necessary) .  Differential  rate 
equations  were  developed  to  describe  the  rearrangements  taking  place  for  the  7 
to  a'  shift  as  well  as  for  each  of  the  other  three  possible  mechanisms.   Theoretical 
calculations,  based  upon  the  experimentally  determined  number  of  hydrogens  in 
the  7  position  and  the  length  of  time  of  heating,  were  carried  out  to  determine 
the  number  of  protons  in  the  a,  (3,  and  tf  positions.   These  calculated  values 
were  then  compared  to  the  experimental  values  to  determine  which  mechanism  best 
fit  the  experimental  data.   Table  2  gives  the  results  of  one  such  set  of  calculations. 
For  broad  peaks,  the  error  in  integration  of  NMR  spectra  can  be  as  high  as  10$  > 

Table  2 
Proton  Distribution  for  7-Deuterocycloheptatriene 
(After  heating  at  i*fO°C  for  kl:2   hours) 
.1      Calculated  For  Mechanism 


Position 

Experime 

7 

1.66 

a 

1.82 

3 

1.86 

K 

1.72 

7"># 

7->6 

7-X2 

7-*Random 

(1.66) 

(1.66) 

(1.66) 

(1.66) 

1.81 

1.86 

1.73 

1.80 

1.86 

1.73 

1.81 

1.80 

1.73 

1.81 

1.86 

1.80 

decreasing  as  the  peaks  get  sharper  and  more  intense  . 7  No  mention  is  made  in 
the  paper  of  what  was  done  to  minimize  this  error,   Consequently,  the  significance 
of  these  numbers  is  in  doubt  and  the  tables  can  not  be  accepted  at  full  face 
value.  In  their  NMR  studies,  ter  Borg  and  KLoosterziel  found  the  same  sequence 
of  signals,  mentioned  above  in  Nozoe1 s  work,  develop  for  the  proton  in  the  7 
position  in  I-b  and  ter  Borg  cites  this  as  evidence  for  the  7  to  5"  shift.   The 
data  given  in  Table  2  appear  to  substantiate  this  conclusion,  but  a  more  positive 
statement  cannot  be  made.  In  accordance  with  the  f-  3"  hydrogen  shift,  ter  Borg 
proposed  the  mechanism  shown  in  Figure  2.   The  multicentered,  intramolecular 
nature  of  the  mechanism  is  further  substantiated  by  an  absence  of  solvent  effects 
on  the  rates  of  rearrangement. 

B 

^ 


V 

Fieure  2 


The  rate  constant  for  the  isomerization  of  I-b,  calculated  from  the  observed 
number  of  protons  in  the  7>  Q>  P>  and  £f  positions ,  was  k=6.0  x  10  7  sec  x   at 
121°C.   It  has  been  observed  that  the  rearrangement  of  I  to  II  is  always  accelerated 
by  substitution.  For  I-d  at  121°C ,  Nozoe  and  Takahasi2  found  k73=2.48  x  ICf5 
sec"1  (extrapolated),  approximately  100  times  as  fast  as  the  unsubstituted  tropilidene. 
At  120. 2°C,  Nozoe  reports  k73=2»31  x  1.0  5  sec  x   (extrapolated)  for  I-d  and  ter  Borg 
reports  )^73-J).G0   x  10  b  sec  -1  for  I-c,6  7~phenylcycloheptatriene,   The  rate  constants 
reported  for  I-d  were  based  on  MR  measurements  and,  since  no  standard  deviation 
is  reported,  these  numbers  should  be  qualified  in  the  same  manner  as  the  MR  work 
by  ter  Borg.  The  rate  constant  for  I-c  was  based  on  UV  measurements  and  should 
be  quite  accurate  but,  again,  no  standard  deviation  is  reported. 

Further  Information  on  these  three  compounds  may  be  gained  by  comparing  their 
activation  parameters  (Table  3)  •  These  values  are  based  on  MR  calculations  with 
no  standard  deviation  given  and,  therefore,  can  only  be  taken  as  approximate 
values.   The  large  negative  AS-  indicates  that  this  reaction  may  proceed  through 
a  transition  state  such  as  that  pictured  by  ter  Borg.   The  decrease  in  activation 
energy  in  going  from  I-b  to  I-c  possibly  reflects  the  amount  of  conjugation 
energy  due  to  the  trienic  and  aromatic  systems  in  the  transition  state.   If  this 

Table  3 
Activation  Parameters 

Compound  AEa(Keal. /mole)   AH-(Kcal./mole)  AS-(e.u.) 

I-b3         31             30.2  =8.2 

I-c          27.6            26.9  -11.7 

I-d9         26.4            25.7  =15.0 

conjugation  exists,  it  would  require  some  co-planarity  between  the  phenyl  group 
and  the  trienic  system  in  the  transition  state.  This  should  result  In  a  more 
negative  AS-^  which  is  observed  experimentally.   Nozoe  ascribes  the  more  negative 
AS^  for  I-d  as  compared  to  I-c  to  some  contribution  of  the  lone  pair  of  electrons 
to  the  ragidity  of  the  transition  state.   However,  Nozoe  is  trying  to  explain 
an  extremely  small  difference  between  two  numbers  ,  of  which  only  the  value  for 
I-c  can  be  considered  significant  since  it  is  based  on  UV  measurements  and  not 
on  MR  data  as  is  the  AS-  for  I-d.   Consequently,  this  explanation  is  highly 
speculative  and,  as  Nozoe  admits,2  more  work  is  necessary.   This  same  problem 
exists  in  trying  to  explain  the  AAS-  for  I-b  and  I~c.   While  the  difference  can 
only  be  regarded  as  the  approximate  value,  when  considered  in  the  light  of  the 
mechanism  proposed  by  ter  Borg  and  the  values  of  the  Arrhenius  activation  energies, 
it  is  at  least  indicative  of  the  difference  which  does  exist  between  the  two 
entropy  values. 

The  equilibrium  concentrations  of  a  series  of  R  groups  was  studied  by  ter  Borg 
and  coworkers.   In  compounds  III,  II,  IV,  and  1  the  R  group  is  in  conjugation 
with  3;  2,  1,  and  0  double  bonds  of  the  trienic  system  respectively  and  their 
respective  concentrations  reflect  the  ability  of  the  isomer  to  be  stabilized  through 
conjugation  as  the  electron  donating  tendency  of  R  increases,  isomer  III  predominating. 
If  R  is  other  than  an  electron-donating  group,  no  overwhelming  preference  of  one 
isomer  over  another  is  shown. 

Equilibrium  Concentrations8 


R 

III 

11 

IV   I 

T  °C 

N(CH3)2 

100 

„_ 

__ 

100 

OCH3 

88 

9 

3   » 

120 

SCH 

76 

16 

8   — 

11.4 

CH3 

57 

2k 

17    2 

140 

C6H5 

64 

18 

1.8   - 

136 

CN 

52 

2k 

2k        — 

142 

The  mechanism  proposed  by  ter  Borg  for  the  rearrangement  involves  a  1,5 
hydrogen  shift  with  migration  of  two  double  bonds,  the  remaining  double  bond  not 
participating.   If  this  hypothesis  is  true,  1,3-cycloheptadienes  should  also 
undergo  rearrangement.   Kloosterziel  and  ter  Borg10  prepared  2^7-aihydrotopone  (v)and 


!Eq.  3) 


•275- 

found  that  it  rearranged  to 
2 , 3-dihydrotropone  ( VI ) .   The 
rate  of  isomerization  was  followed 
via  UV  and  the  equilibrium  constant 
calculated.  At  60°C  ;.  K=k1/k„1=l,86- 
kj^l.texlO""6  sec"1  and  k,.1=7»6  x  10  7 
sec  1.   These  rate  constants  include 
no  significant  solvent  effect  s 
at  60°c  ki  +  k~i=2.l8  x  10  6sec  1   in  n-heptane  and  4. 5  x  10  6sec  -1  in  ethanol. 
The  slight  solvent  effects  which  do  appear  may  be  due  to  the  solvation  of  the 
carbonyl  group.   At  101°C,  ki=l.l6  x  10  4sec  1  or  approximately  104  times  k73 
for  tropilidene  itself.   There  is  no  spectral  evidence  for  the  presence  of  an 
enol  and,  therefore,  a  maximum  of  only  a  few  percent  may  be  present.  As  a  result 3 
formation  of  the  enol  and  subsequent  hydride  transfer  is  ruled  out  by  ter  Borg, 
who  proposes  that  the  small  amount  of  enol  which  might  be  formed  would  be  insufficient 
to  account  for  the  large  rate  enhancement.   Without  knowing  any  rate  data  for  the 
enol,  this  is  not  a  well-grounded  assumption.   If  the  rate  of  reaction  of  the 
enol  was  very  rapid,  in  conjunction  with  a  steady  state  approximation,,  a  few  percent 
enol  could  easily  account  for  the  rate  enhancement.   It  would  be  necessary  to 
show  that  enolization  catalysts ,  such  as  acids,  do  not  affect  the  rate  of  reaction 
to  prove  that  enolization  does  not  cause  the  rate  enhancement. 

The  predictions  of  Woodward  and  Hoffmann11  also  support  the  1,5  hydride  shift. 
Due  to  the  constraints  imposed  by  the  ring,  suprafacial  hydride  shifts  are  the 
most  likely  transfer  mechanism  and,  thermally,  only  the  1,5  shift  would  be  allowed 
to  proceed  in  a  suprafacial  manner;  the  1,3  and  1,7  shifts  would  be  antarafacial. 
On  the  other  hand,  photochemically,  the  1,3  and  1,7  shifts  are  allowed  to  be 
suprafacial  and  the  1,5  shift  would  be  allowed  in  an  antarafacial  manner, 
Razenberg  and  ter  Borg12^13  found  that  the  7-substituted  tropilidenes  do  indeed 
'undergo  a  1,7  shift  in  photolytically  induced  rearrangements.   Thus  all  the  work 
carried  out  substantiates  the  proposed  1,5  hydride  shift  in  thermal  isomerlzations. 
Skeletal  Reorganizations 
A)   Norbornadiene-Cycloheptatriene  Isomerization 

W.  G.  Wood14  studied  the  isomerization  of  norbornadiene  to  form  cycloheptatriene 
and  found  that  toluene  and  the  products  of  the  reverse  Diels-Alder  reaction, 
acetylene  and  cyclopentadiene ,  were  also  formed.   The  mechanism  shown  in  Figure  3 
was  considered  because  small  amounts  of  benzene  and  ethylene  were  found  as  side= 
products  and  carbene  is  known  to  add  to  benzene  to  give  tropilidene  and  toluene. 


+  HoC: 


Figure  3 


VII 


The  norbornadiene  was  pyrolysed  in  the  presence  of  a  large  excess  of  n-butane 
in  an  attempt  to  trap  the  carbene.  However,  no  trace  of  pentanes  or  benzene 
were  found  while  the  isomerization  and  the  reverse  Diels-Alder  proceeded  to 
give  the  expected  yields.   Two  controls  were  run  to  demonstrate  that  carbene  does 
react,  under  the  experimental  conditions,  with  n-butane  to  give  pentanes  and 
with  benzene  to  give  toluene.   Wood,  therefore,  concluded  that  this  was  not  the 
correct  mechanism  and  proposed  the  mechanism  shown  in  Figure  h. 


<r—> 


<r 


il     ^ 


Figure  h 

Lustgarten  and  Richey15  studied  the  rearrangement  of  7-phenyl  and  7-alkoxy- 
norbornadienes  to  cycloheptatrienes  and  found  only  the  1,  2,  and  ^-substituted 
tropilidenes,  no  aromatic  isomers  where  found.   No  significant  solvent  effects 
were  observed  which  tends  to  rule  out  charge  separation  in  the  transition  state 
and  indicates  that  an  intramolecular  process  is  possible,, 


(Eq.  k) 


VII 


(in  n-decane; 


-$■ 


a) 

R=H 

b) 

R=D 

c) 

R=<1> 

d) 

R=OMe 

e) 

R=OC(CH3)3 

f) 

R=0C2H.4QC2H5 

AHi=34,5  Kcal./mole,  A&^=1  e,u.  for  Vll-e) 


When  R=H,  the  isomerization  occurred  at  ^75°C  and  toluene,  cyclopentadiene , 
and  acetylene  were  produced  as  well  as  tropilidene.  When  R  was  phenyl  or  alkoxy, 
the  isomerization  was  carried  out  at  only  175°C  and  only  the  substituted  tropilidenes 
were  observed.  In  contrast,  when  the  methylene  carbon  had  a  ketal  or  thioketal 
substituent  attached  to  it,  the  isomerization,  which  can  be  carried  out  in  the 
same  temperature  range,  produced  only  benzene  and  compounds  derived  from  the  ketal 
or  thioketal  function  and  no  tropilidenes  were  found.   Using  a  capillary  furnace 
mounted  at  the  inlet  of  a  mass  spectrometer,  Lemal  and  coworkers16-5'1'7  detected 
the  presence  of  *CH3,  C02,  and  CH3C02CH3  when  7>7-&Imethoxynorbornadiene  was 
pyrolysed.  They  proposed  that  benzene  and  dimethoxy  carbene  were  initially 
produced  and  the  carbenes  then  decomposing  by  a  radical  chain  mechanism  to  give 
the  products  observed.   To  account  for  the  formation  of  benzene  and  the  dimethoxy 
carbene  in  a  manner  consistent  with  a  general  mechanism  for  the  various  substituted 
norbornadienes  discussed  above,  Lustgarten  and  Richey  proposed  the  intermediate 
shown  in  equation  5«   Using  this  general  mechanism,  the  equilibrium  shown  in 
equation  6  should  exist  for  the  mono substituted  norbornadiene  but  the  fragmentation 


H3CO 


(Eq.  5) 


(Eq.  6) 


+  (CH30)2C; 


-2(7- 

of  the  norcaradiene  intermediate  would  result  in  a  less  stable  carbene  than  in 
the  dimethoxy  case  and,  therefore  might  be  slower  than  the  competing  1.5-hydrogen 
transfer.  However,  direct  loss  of  dimethoxy  carbene  without  intermediates  cannot 
be  ruled  out,  and  since  tropilidenes  are  not  known  to  decompose  to  give  carbenes, 
Lustgarten  and  Richey  have  no  evidence  for  the  intermediates  in  equation  5. 

Herndon  and  Lowry18  studied  the  kinetics  of  the  isomerization  of  norbornadiene 
to  cycloheptatriene  to  determine  if  the  toluene  produced  was  generated  directly 
from  the  norbornadiene  or  from  the  cycloheptatriene  or  both.   (See  Figure  5.) 
They  used  a  gas  phase  stirred  flow  reactor  designed  so  that  the  contents  of  the 
reactor  are  completely  mixed  by  diffusion,  thereby  leading  to  uniform  concentrations 
which  become  time -invariant  within  the  reactor. 


( BCH) 


( CPD) 


The  experimental  results 
show  that  toluene  is  produced 
from  both  compounds.   This 
result  was  indicated  in  some 
previous  work  by  Klump  and 
Che sick19  but  not  proven 
conclusively.   Table  k   gives 
the  data  obtained  by  Her,don 
and  Lowry. 


Figure  5 


Table  k 
Rate  Constants  and  Arrhenius  Activation  Energies 


( 1)  BCH  ->  CPD  +  A 

(2)  BCH  ->  CHT 

(3)  BCH  -»  T 
( h)  CHT  ->  T 


k  (sec  x) 

(400.6°C) 
2.31  x  10~2 
1.73  x  I0~2 

1.04  x  l(f3 

1.05  x  10" 3 


££a   (calo/mole) 


50,190  +  76O 
50,610  +  78O 

53,1^0  ±  730 
52,100  +  820 


Herndon  and  lowry  claim  all  of  the  reactions  are  first  order  and 

unimolecular  and  present  the  reaction  scheme  shown  in  Figure  6  as  the  most  likely 

mechanism.   The  symbol  [I]  represents  a  common  intermediate  to  tropilidene  and 

toluene.  Birely  and  Chesick20  have  also  examined  these  reactions  and  have  obtained 


^  CHT 


BCH  >  [Ij 

I 

CPD  +  A  "  T 

Figure  6 

almost  identical  values  for  the  various  rate  constants  and  energies  of  activation. 
Since  the  values  of  the  activation  energies  for  all  of  the  equations  in  Table  k 
are  almost  identical,  Birely  and  Chesick  claim  that  this  is  evidence  for  a  common 
intermediate  to  all  three  products:   cyclopentadienes  (+  acetylene) ,  cycloheptatriene, 
and  toluene.  Herndon  and  Lowry18  dispute  this  conclusion  and  argue  that  this 
is  evidence  against  a  common  Intermediate  and  is  evidence  only  for  similar  initial 
steps.  If  a  common  intermediate  for  the  reverse  Diels-Alder  reaction  and  the 
isomerization  did  exist,  the  ratio  of  the  rate  constants  can  be  shown,  to  be  the 
ratio  for  the  reactions  which  take  place  after  the  common  intermediate  is  formed. 
An  Arrhenius  plot  of  this  ratio  gives  the  difference  in  activation  energies.  In 
this  case  the  difference  is  extremely  small,  400  cal. ,  and  this  would  Indicate  that 
the  two  processes  occurring  after  the  intermediate  is  formed  are  essentially  similar. 
But  the  isomerization  and  reverse  Diels-Alder  reactions  are  not  similar.   Thus,  no 
common  intermediate  exists  and  it  is  simply  fortuitous  that  the  activation 


-270- 
energies  are  almost  identical.   Using  this  same  approach,  it  might  also  be  argued 
that  the  similarity  of  the  activation  energies  after  the  common  intermediate  is 
fortuitous,  thereby  allowing  no  distinction  to  be  made.  Tropilidene  has  not, 
as  yet,  beerj  reported  as  a  side  product  in  the  Diels-Alder  synthesis  of  norbornadiene 
as  might  be  expected  if  a  common  intermediate  did  exist  and  this  has  been  offered 
in  support  of  the  theory  of  Herndon  and  Lowry.  However,,  based  upon  the  difference 
in  the  heats  of  formation ^35^36  formation  of  norbornadiene  is  favored  by  approximately 
8  Kcal./moie  and,  consequently,  it  may  be  formed  with  complete  exclusion  of  topilidene. 

B)  Norcaradiene  -  Cycloheptatriene  1 somer  iz ation  ° 

Some  chemical  reactions  of  tropilidene  often  give  products  which  appear  to 
come  from  norcaradiene^9 °21  while  others  give  products  involving  the  cycloheptatriene 
structure.   This  ambiguous  nature  would  make  it  appear  that  tropilidene  is  in 
equilibrium  with  its  valence  tautomer  -  norcaradiene. 

Anet22  studied  the  temperature  dependent  MR  spectra  of  tropilidene  to  determine 
if  it  is  planar  or  nonplanar.   At  -150°C,  the  methylene  protons  give  rise  to  two 
chemically  shifted  bands  with  a  separation  of  76  cps  and  increasing.   The  mean. 
chemical  shift  is  7»8t,  essentially  unchanged  from  that  observed  at  higher 
temperatures.   Since  the  methylene  protons  are  nonequivalent  at  low  temperatures, 
cycloheptatriene  is  nonplanaro  In  similar  work,  Jensen  and  Smith23  were  able  to 
get  down  to  -170°C  and  the  separation,  between  peaks  was  86  cps.  They  were  not 
able  to  find  any  evidence  for  the  presence  of  norcaradiene.  An  electron  diffraction 
study  by  Traettenberg24  showed  that  tropilidene  was  indeed  nonplanar  with  the 
plane  comprised  of  the  1,2,5^  and  6  carbons  making  an  angle  of  36. 5°  with  the 
plane  of  the  1,6,  and  7  carbons  and  an  angle  of  40.5°  with  the  plane  of  the 
2,^,k,   and  5  carbons. 

An  equilibrium  between  tropilidene  and  norcaradiene  should  be  detected  by 
variations  in  the  coupling  constants  with  changes  in  temperature,,  Roberts  and 
coworkers25  carried  out  a  complete  analysis  of  cycloheptatriene  at  -70°C  and 
+115° C  but  no  change  in  coupling  constants  could  be  observed.  Consequently, 
there  is  less  than  5$  norcaradiene  present  even  at  -7C°C.   When  7^7-bistrifluor- 
omethylcycloheptatriene  was  studied,25  the  trifluoromethyl  groups  remained  equivalent 
down  to  -l8.5°C  and  there  is  no  evidence  of  any  of  the  norcaradiene  structure. 

Both  the  trifluoromethyl  group  and  the  cyano  group  are  strongly  electron 
withdrawing,  having  negative  resonance  effects  and  negative  inductive  effects.26 
The  Hammett  para-sigma  constants  are  0..54  and  0.66  respectively.2^  Yet,  when  one 
of  the  CF3  groups  is  replaced  by  a  CN  group,  some  norcaradiene  is  observed.28  At 
-85°C,  the  relative  concentration  of  the  substituted  tropilidene  to  substituted 
norcaradiene  is  80s 20  as  calculated  from  WMR  spectra  peaks.  When  both  trifluoromethyl 
groups  are  replaced  by  cyano  groups,  only  the  7>7~dicyanonorcaradiene  (VIII) 
is  observed.29  At  100°C,  VTII  rearranges  to  phenylmalonitrile  and  3,7-dicyano- 
cycloheptatriene  (IX)  which  in  turn  undergoes  a  series  of  1,5  hydrogen  shifts 
to  give  1,4-  and  l,5~dicyanocycloheptatriene  upon  further  heating. 

C)  7  >7-Di substituted  Cycloheptatriene s ; 

Berson  and  Willcott'-'    examined  the  norcaradiene-diradical  intermediate 
equilibrium  proposed  by  Wood  (Fig.  k)    in  an  attempt  to  determine  if  the  rate  of 
recyclization  of  the  diradical  intermediate  to  form  norcaradiene  occurred  at 
a  rate  competitive  with  hydrogen  transfer  to  form  toluene  from  the  diradical 
intermediate.  Blocking  the  1,5-hydrogen  transfer  by  di substitution  at  the  7 
position,  3,7^7-trimethyltropilidene  (x)  was  pyrolysed  at  300°C  to  produce  the 
profusion  of  products  shown  in  equation.  7;  the  percentages  given  refer  to  the  com- 
position of  the  reaction  mixture  after  pyrolysis  for  40  minutes,  but  do  not  include 
1  -  2$  of  unidentified  products. 


-27S- 


H3<^3?5<CH3       H3VCH2     H3°        CH2  H3Cf,CH3  Ife=  ^CH3 

it.  f\  +  fi  ;aQ^  ^ 


1=/        HsC^^^R 


H*C 


X(35^)  Xl(25#)  Hl(2l)  XIII(2<#)    XIV(l6$)        XV(2$)      (XVI)  (2^) 

a)    R=H  b)    R=D 

The  mechanism  of  the  isomerization  has  been  shown  by  deuterium  labeling  to 
be  a  true  skeletal  rearrangement  of  the  ring  carbons  rather  than  a  superficial 
series  of  hydrogen  shifts.  Examination  of  equation  7  shows  that  the  1-6  carbon 
chain  maintains  the  same  sequence  in  each  compound ,  while  the  C-7  carbon  and 
its  geminal  methyls  are  allowed  to  wander  and  reattach  between  any  pair.   Both 
nonaromatizing  rearrangements  to  form  XI  and  XIV  are  reversible.   When  either  XI 
or  XIV  is  isolated  and  resubjected  to  the  reaction  conditions,  a  typical  mixture 
of  pyrolysis  products  results.  Mass  spectroscopy  studies  have  shown  that  the 
rearrangements  do  not  take  place  by  an  intermolecular  process.  The  mechanism 
shown  In  Figure  7 ,   proposed  by  Berson  and  Willcott,32  incorporates  the  major  structural 
changes,  the  intramolecularity,  and  the  reversibility.   This  mechanism  indicates 
that,  compounds  XVII,  XVIII,  and  XIX  should  be  formed.   So  far  these  products 
have  not  been  found  but  may  be  present  in  the  1-2$  of  unidentified  material 
or  may  be  formed  after  a  longer  period  of  time  than  Berson  and  Willcott  ran  their 


XVIII 


Figure  7 


XIX 


experiments.   Production  of  the  dienes  XIII  and  XIV  is  not  without  precedent. 
Employing  the  norcaradiene  intermediate,  this  isomerization  can  be  classified 
as  a  Cope  type  rearrangement  with  one  double  bond  being  replaced  by  a  cyclopropyl 
ring  and  a  hydrogen  from  one  of  the  geminal  methyls  being  transfered  to  the  six 
member  ring.37-*38  There  is  no  way  to  distinguish  a  diradical  mechanism  from 
a  concerted  1,5-carbon  shift  without  intermediates.   This  latter  process  is 
permitted,  but  not  required,11  and  requires  the  C-7  migration  from  C-l  to  C-5 
to  be  suprafacial.  An.  examination  of  properly  substituted  optically  active 
tropilidenes  should  distinguish  this  from  an  antarafacial  process  and  from  a 
mechanism  involving  a  diradical  intermediate.   This  work  is  currently  in  progress. 


32 


-280- 

Fhotochemical  Isomerization: 

Chapman  and  Borden^'J  found  that  Irradiation  of  neat  7-alkoxycycloheptatriene 
produced  mainly  the  substituted  bicyclo[3.2.0]heptatriene  (XX-b1)  and  only  trace 
amounts  of  toluene.  If  this  irradiation  is  carried  out  in  the  vapor  phase ,  both 
XX-b1  and  1-alkoxycycloheptatriene  (ill-b')  are  produced.   Pyrolysis  of  XX-b' 
produces  Ill-b'  and  irradiation  of  Ill-b1  converts  it  to  XX-b1.   Srinivasan34 
found  that  irradiation  of  tropilidene  in  the  vapor  phase  produces  mainly  toluene 
and  a  maximum  of  5$  bicyclo[3.2.0]heptadiene  (XX-a).   The  formation  of  neither 
toluene  nor  XX-a  was  quenched  by  addition  of  oxygen  or  nitric  oxide ,  which  indicates 
that  the  isomerization  is  intramolecular  and  that  it  does  not  arise  from  a  triplet 
state  of  tropilidene.   Srinivasan  has  postulated  that  toluene  is  formed  from  a 
vibrationally  excited  ground  state  and  not  from  an  upper  electronic  state.  The 
production  of  the  isopropyl-toluenes  upon  pyrolysis  of  3,7,7-trimethyltropilidenes 
shows  that  this  reaction  also  occurs  thermally  and  this  particular  photochemical 
example  was  included  because  of  this  correspondence. 

Bibliography 

1)  W.  G.  DeWitt,  Univ.  of  111.  Organic  Seminar  Abstracts,  1963,  P.  19. 

2)  T.  Nozoe  and  K.  Takahashi,  Bull.  Chem.  Soc  Japan,  ^8,  665,  (1965). 

3)  A.  P,  ter  Borg,  H.  KLoosterziel,  and  N.  van  Meurs ,  Rec.  Trav,  Chim.,  82,  717,  (1963), 
k)    Z.  N.  Parnes,  M.  E,  Volpin,  and  D.  N.  Kursanov,  Tet.  Let.,  No.  2.1,  20,  (i960). 

5)  A.  P.  ter  Borg,  H.  KLoosterziel,  and  N.  van  Meurs,  Pro.  Chem.  Soc,  359,  (1962). 

6)  A.  P.  ter  Borg  and  H,  KLoosterziel,  Rec.  Trav.  Chim.,  82,  74l,  (1963). 

7)  J.  A.  Pople,  W.  G.  Schneider,  and  H.  J.  Bernstein,  "High  Resolution  Nuclear 
Magnetic  Resonance,"  McGraw-Hill,  New  York,  N.Y. ,  1959,  P.  78« 

8)  A.  P.  ter  Borg,  E.  Razenberg,  and  H.  KLoosterziel,  Rec.  Trav,  Chim.,  82,  1230,(1963) 

9)  E.  Weth  and  A.  S.  Dreiding,  Pro.  Chem.  Soc,  59,  (1964). 

10)  A.  P.  ter  Borg  and  H.  KLoosterziel,  Rec  Trav.  Chim.,  82,  II89,  (1963). 

11)  R.  B.  Woodward  and  R.  Hoffmann,  J.  Am.  Chem.  Soc,  87,  2511,  (1965). 

12)  A.  P.  ter  Borg  and  E.  Razenberg,  Rec.  Trav.  Chim.,  B4,  24l,  (1965). 

13)  A.  P.  ter  Borg  and  E.  Razenberg,  Rec  Trav.  Chim.,  BE,  245,  (1965). 

14)  W.  G.  Wood,  J.  Org.  Chem.,  23,  ilO,  (1958). 

15)  R.  Ko  Lustgarten  and  H.  G.  Richey,  Jr.,  Tet.  Let.,  4655,  (1966). 

16)  D.  Mo  Lemal,  R.  A.  Lovald,  and  R. '  W.  Harrington,  Tet.  Let.,  2779,  (1965). 

17)  D.  M.  Lemal,  E.  P.  Gosselink,  and  S.  D.  McGregor,  J.  Am.  Chem.  Soc,  88,582,(1966). 

18)  W.  C.  Herndon  and  L.  L.  Lowry,  J.  Am.  Chem.  Soc,  86,  1922,  (1964). 

19)  K.  N.  KLump  and  J.  P.  Chesick,  J.  Am.  Chem.  Soc,  8£,  130,  (1963). 

20)  J.  H.  Birely  and  J.  P.  Chesick,  J.  Phy.  Chem.,  66^   568,  (1963). 

21)  W.  von  E.  Doering,  G.  Laber,  R.  Vonderwahl,  N.  F.  Chamberlain,  and  R.  B. 
Williams,  J.  Am.  Chem.  Soc,  78,  5448,  (1956). 

22)  F.  A.  L.  Anet,  J.  Am.  Chem. " Soc. ,  86,  458,  (1964). 

23)  F.  R.  Jensen  and  L.  A.  Smith,  J.  Am.  Chem.  Soc,  86,  956,  (1964). 

24)  M.  Traettenberg,  J.  Am.  Chem.  Soc,  86,  4265,  (19^5). 

25)  J.  B.  Lambert,  C.  J.  Durham,  P.  Lepoutere,  and  J.  D.  Roberts,  J.  Am.  Chem. 
Soc,  87,  3896,  (1965). 

26)  E.  S.  Gould,  "Mechanism  and  Structure  In  Organic  Chemistry,"  Holt,  Rinehart, 
and  Winston,  New  York,  N.  Y, ,  1959,  P.  218. 

27)  J.  Hine,  "Physical  Organic  Chemistry",  McGraw-Hill,  New  York,  N.  Y. ,  I962,  P.  87. 

28)  E.  Ciganek,  J.  Am.  Chem.  Soc,  87,  1149,  (1965). 

29)  E.  Ciganek,  Private  communication, 

30)  J.  A.  Berson  and  M.  R.  Willcott,  III,  J.  Am.  Chem.  Soc,  87,  2751,  (1965). 

31)  J.  A.  Berson  and  M.  R.  Willcott,  III,  J.  Am.  Chem.  Soc,  B?,  2752,  (I965). 

32)  J.  A.  Berson  and  M.  R.  Willcott,  III,  J.  Am.  Chem.  Soc,  B8,  2494,  (1966). 

33)  0.  J.  Chapman  and  G.  W,  Borden,  Proc  Chem.  Socl,  221,  (1963). 

34)  R.  Srinivasan,  J.  Am.  Chem.  Soc,  84,  3432,  (1962). 

35)  A.  F.  Bedford,  A.  E.  Beezer,  C.  T.  Mortimer,  and  H.  D.  Springall,  J.  Chem.  Soc, 

3823,  (1963). 

36)  H.  L.  Finke,  D.  W.  Scott,  M.  E,  Gross,  J.  E.  Messerly,  and  G.  Waddington, 
J.  Am.  Chem.  Soc,  78,  5469,  (1958). 

37)  P.  J.  Ellis  and  II.  M.  Frey,  Proc  Chem.  Sou.  221,  (1964). 

38)  G.  Ohloff,  Tet.  Let.,  3795,  (1965). 


CYCLIZATION  REACTIONS  OF  N-HALOAMINES,  -AMIDES,  AND  --IMINES 

Reported  by  Daniel  R,  Bloch  March  2j5,  1967 

INTRODUCTION 

There  have  been  many  reports  in  the  recent  literature  concerning  the  use  of 
nitrogen  free  radicals  in  organic  synthesis.1  Although  cyclization  reactions  of 
N-halogenated  nitrogen  compounds  have  been  known  for  a  long  time,1  only  recently 
has  the  mechanism^2"6  the  influence  of  solvent  and  structure  on  the  efficiency  of 
the  process, 1,2>7   9  and  addition  of  nitrogen  radicals  to  unsaturated  compounds  been 
studied.7"13  Preference  for  formation  of  pyrrolidines  and  7 -lactones  from  the 
photolysis  of  N-halo  compounds  has  been  well  documented. 1"4>8  Readily  obtainable 
starting  materials,  good  yields  and  relatively  simple  reaction  conditions  are 
characteristic  of  these  reactions. 

N-HALOAMINES 

The  first  cyclization  reaction  involving  N-haloamines  to  appear  in  the  lit- 
erature was  reported  by  Hofmann  in  I883.13  The  reaction  of  N-bromoconiine  (I)  in 
hot  sulfuric  acid  and  subsequent  treatment  with  base  afforded  B-coneceine  (II). 


1)  H5SQ4.UK)0 

2)  OH"        ' 


I  II 

No  further  work  appeared  until  the  early  1900' s  when  Loeffler  and  co-workers  reported 
further  examples  of  cyclization  reactions  of  N-haloamines  including  an  elegant 
synthesis  of  nicotine  (III).14  Thus,  reactions  of  this  nature  have  been  called 


1)  H2S04,100C 


2)  OH' 


III 


Hofmann-Loeffler  (H-L)  reactions,  although  other  names  have  been  used.1  In  a  review 
by  Wolff,1  a  table  of  reported  H-L  reactions  has  been  compiled  which  is  complete 
through  the  1950' s  and  describes  reactants,  products  and  reaction  conditions. 

Wawzonek3*5  was  the  first  to  study  the  mechanism  of  the  H-L  reaction  and  later 
work  by  Lukes6  and  Corey4  have  shown  the  reaction  to  be  a  free  radical  chain  process. 
The  following  mechanism  has  been  proposed. 

1+    RpNHCI    (5)   ?©       +' 
— »   -RI-H    ^    >   Cl-R-N-H  +  RsNH 

or  R2NCI 


+ 

RsNHCl 


+  • 

RsNH 


+  CI' 


OH 
Corey4  has  shown  that  a  0. J>hU  solution  of  N-ehlorodi-n-  (I) 

butylamine  (NCBA)  in  85/0  sulfuric  acid  at  25°  is  stable 

in  the  dark  for  285  minutes,  although  decomposition  R-N-R 

could  be  induced  by  the  addition  of  ferrous  ion.  {5) 

Photolysis  of  the  NCBA  solution  showed  an  induction 

period  which  could  be  essentially  removed  by  purging  the  solution  with  nitrogen. 

The  reaction  could  be  interrupted  by  removing  the  light  source  and  started 

immediately  by  irradiating  again.  These  features  are  characteristic  of  radical 

chain  react! ^1= 


-282  - 

Corey4  prepared  the  N-chloro  compounds  by  passing  chlorine  gas  over  a  ligroin 
(60°-90°)  solution  of  the  amine.  The  resulting  solution  was  washed  with  dilute 
acid  and  dilute  base  before  extracting  the  chloroamine  into  85/0  sulfuric  acid.  For 
reactions  in  anhydrous  solvent,  the  ligroin  solution  was  washed,  dried,  and  con- 
centrated in  vacuo.  An  aliquot  of  the  residue  was  taken  up  in  absolute  acetic  acid. 
The  amount  of  N-chloroamine  assumed  present  was  based  on  the  total  amount  of  active 
chlorine  as  determined  by  titration. 

H-L  reactions  involving  acyclic  reactants  generally  lead  to  pyrrolidines1'4'8 
which  are  formed  (eq  1)  by  a  cyclic  mechanism  involving  the  nitrogen  and  a  hydrogen 
on  a  8 -carbon.  The  predominance  of  hydrogen  removal  at  the  6 -carbon  favors  an 
intramolecular  hydrogen  abstraction  which  occurs  preferentially  thru  a  quasi - 
six-membered  transition  state.  If  the  process  were  intermolecular,  one  would  expect 
more  random  hydrogen  abstraction  giving  a  greater  variety  of  products. 

Radicals  which  abstract  hydrogen  atoms  from  carbon  generally  show  a  preference 
for  hydrogen  in  the  order  tertiary  />   secondary  )>  primary.  This  same  order  of 
reactivity  is  followed  for  H-L  reactions.  In  the  free  radical  decomposition  (85$ 
sulfuric  acid,  95°)  of  N-chlorobutylamylamine  (IV),  two  products  are  possible  from 
hydrogen  abstraction  at  a  6-carbon.   l-n-Butyl-2-methylpyrrolidine  (V)  would  result 
from  secondary  hydrogen  abstraction  and  1-n-amylpyrrolidine  (VI)  would  be  formed 
from  primary  hydrogen  abstraction.  The  fact  that  (V)  was  the  only  tertiary  amine 

CH3CH2(CH2)35J(CH2)3CH3         I   H  [~  ~J 

CI  w    CH3  N 

n-C4H9  n-C5Hi:L 

IV  V  VI 

isolated  shows  the  preference  of  secondary  over  primary  hydrogen  abstraction.  For 
a  comparison  of  the  reactivity  of  tertiary  and  secondary  hydrogen  and  of  tertiary 
and  primary  hydrogen,  the  N-chloro 

derivatives  of  n-butylisohexylamine  and  n-amylisohexylamine  were  prepared  and 
subjected  to  H-L  conditions.  No  tertiary  amine  could  be  isolated  although  the 
disappearance  of  N-chloroamine  was  very  rapid  and  accompanied  by  the  evolution  of 
hydrogen  chloride.  It  was  suggested  that  the  tertiary  chloro  compound  was  formed 
and  rapidly  solvolyzed  in  strong  sulfuric  acid.  It  was  also  found  that  t-butyl 
chloride  liberated  hydrogen  chloride  when  shaken  with  85$  sulfuric  acid. 

The  stereochemistry  of  the  H-L  reaction  was  studied  by  thermally  decomposing 
the  N-chloro  derivative  of  ( -) -methylamylamine-4-d  (VII)  in  sulfuric  acid  at  95°. 
The  products,  1,2-dimethylpyrrolidine  (Villa)  and  l,2-dimethylpyrrolidine-2-d 
(VHIb)  ,  isolated  in  kyf>   yield  were  optically  inactive.  An  isotope  effect  (%/kjj) 


Jt-h(d) 
:h, 

CH3 


"T^fe 


VII  Villa  ,b 

of  3.5^+  was  observed  for  the  reaction.  This  result  strongly  suggests  the  decom- 
position involves  an  intermediate  in  which  the  8 -carbon  is  trigonal. 

As  evidence  that  a  6-chloro  compound  is  an  intermediate,  Corey4  treated  the 
solution  resulting  from  photolysis  of  NCBA  with  silver  ion.  Practically  no  silver 
chloride  precipitated.  When  the  resulting  solution  was  made  basic,  silver  chloride 
was  obtained  in  ca.  99$  yield.  Reaction  under  thermolytic  conditions  gave  a  65/0 
yield  of  silver  chloride.  This  suggests  that  the  unreactive  chlorine  was  bound  to 
a  carbon  atom  prior  to  hydrolysis.   Since  basification  resulted  in  cyclization  to  the 
5-carbon  and  freeing  of  the  chloride,  it  is  reasonable  that  the  chloride  was  bound 
to  the  5-carbon.  Wawzonek5  was  successful  in  isolating  the  4-chloro  derivative  in 
377o  yield  from  the  decomposition  of  NCBA  in  sulfuric  acid. 


-28j- 

Recent  work  by  Neale8  has  shown  that  side  reactions  may  also  be  important  in  the 
photolytic  decomposition  of  NCBA.  He  studied  the  reaction  with  respect  to  a)  acidity, 
b)  degree  of  purity  of  chloroamine,  c)  applied  irradiation  and  d)  the  rate  at  which 
nitrogen  swept  the  reaction  mixture.  Most  decompositions  were  run  at  20°  in  acetic 
acid  1..5M  in  water  while  the  molarity  of  sulfuric  acid  was  varied  (Table  I). 

Table  I  Photolytic  Rearrangement  of  N-Chlorodi-n-butylamine 


■ry  ' 

Molarity 

Source 

H2SO4 

BusjNCI 

Bu^NCl 

1 

0 

0.223 

2 

0.5 

0.260 

5 

1.0 

0.255 

k 

2.5 

0.243 

5 

1.5 

0.44 

U-l 

6 

1.5 

o.46 

U-2 

7 

1.5 

0.46 

U-l 

8 

1.9 

0.46 

U-l 

9 

3.9 

0.46 

U-l 

10 

1.5 

0.46 

U-2 

1.1 

1.9 

0.46 

U-2 

12 

3.9 

0.46 

U-2 

13 

1.9 

0.46 

D 

14 

3.9 

0.46 

D 

15 

7.7 

0.46 

D 

16 

1.0 

0.46 

D 

17 

0.97 

0.48 

U-l 

18 

1.0 

0.25 

U-l 

19 

2.5 

0.24 

U-l 

a^ 

Entries 

3    1=4, 

ref  4^  entries 

5-19,  : 

Decomposition 


2910 
62 
52 

^7 
0.81 
1.62 

O.96 
0.37 
1.08 
O.98 
O.54 

0.43 
0.24 
0.16 
0.05 
0.10 
0.22 
0.14 


Yield  of  N-butyl- 
pyrrolidine 

0 

42 

59 
80 

75 

42 
41 
56 
88 

65 

8~7 

95 
49 
17 
60 

75 


ref  8.   Irradiations  entries  1-4,  quartz 
lamp,  range  200-400  mu^  entries  5-19*  Hanovia  mercury  arc  lamp  with  filter, 
transmition  300-400  rnu.  ^Nitrogen  flow?  entries  1-4,  reaction  run  under  nitrogen ^ 
entries  6-19>  nitrogen  bubbled  slowly  through  solution j  entry  5*  nitrogen  rapidly 
bubbled  through  solution.   Solvents  entries  1-4,  18  and  19  run  in  anhydrous  acetic 
acid 5  entries  5-17  run  in  acetic  acid  1.5M  in  H20.   Entries  1-4,  prepared  as  in  ref 
15.  ^Determined  by  loss  of  active  chlorines  entries  1-4  are  half -life  values *j 
entries  5-19  are  in  mmoles  of  chloroamine  consumed  per  min. ,  which  is  constant  for 
O-80/0  of  the  reaction,  except  entry  17:0-30$  (ref  27). 


The  chloroamine  was  prepared  from  NCS  and  amine  in  ether  by  stirring  for  one  hour, 
washing  with  water  and  dilute  sulfuric  acid,  drying  and  evaporating  the  solvent. 
Reagent  grade  NCS  and  amine  gave  product  ( U-I) ,  recrystallized  NCS  and  amine  gave 
product  (U-II)  and  distilled  U-I  gave  product  (D) .   Increased  chloroamine  purity 
increased  yields  of  N-n-butylpyrrolidine  (NBP)  and  decreased  the  rate  of  reaction 
(Table  I). 

Wnen  the  intensity  of  the  irradiation  was  decreased  or  lower  wave  lengths 
filtered  out,  the  rate  of  reaction  was  suppressed  and  the  yields  of  NBP  were  decreased. 
The  rate  of  flow  of  nitrogen  bubbled  through  the  reaction  mixture  also  determined  the 
rate  of  reaction.   Increased  rate  of  flow  increased  the  yield  of  KBP  but  decreased 
the  rate  of  reaction.  Trapping  spent  nitrogen  showed  volatile  substances  were  being 
removed  from  solution  by  the  nitrogen. 

In  the  dark,  acid  solutions  of  distilled  and  undistilled  chloroamine  were 
unstable  and  UV  spectra  showed  formation  of  a  new  compound  with  X^^  at  306  m^..  At 
a  given  sulfuric  acid  concentration  in  acetic  acid  1.5M  in  water,  the  rate  of  formation 
of  this  new  compound  in  the  dark  was  the  same  for  distilled  and  undistilled  chloroamine. 
The  rate  was  also  found  to  be  dependent  on  sulfuric  acid  concentration.   In  a  solution 
1M  in  sulfuric  acid  decomposition  is  rapid  whereas  in  a  solution  4M  in  sulfuric 
decomposition  is  very  slow.8  The  absorbing  species  was  proven  to  be  volatile  by 

*in  min. 


sweeping  it  in  a  nitrogen  stream  from  a  weakly  acidic  solution  of  undistilled  chloro- 
amine  into  an  acetic  acid  trap.  Extraction  with  pentane  gave  a  solution  with 
absorption  at  307  mn.  The  species  responsible  for  this  absorption  was  shown  to  be 
N,N-dichloro-n-butylamine  by  comparison  with  a  sample  formed  from  NCS  and  butylamine 
(  ether      g      ^ 
v  max   ^  '  max  "^ 

Photolytic  decomposition  of  NCBA  was  followed  by  periodically  taking  the  UV 
spectra  of  aliquots.  When  chloroamine  of  purity  U-I  and  U-II  was  photolyzed  in 
aqueous  acetic  acid,  new  variable  absorption  appeared  in  the  region  312-320  mu  which 
grew  to  a  miximum  near  60-70%  reaction  and  disappeared  when  the  active  titer  fell  to 
zero.  A  less  pronounced  maximum  appeared  in  the  region  320-340  mu  in  anhydrous 
acetic  acid.  When  distilled  chloroamine  was  used  the  new  absorption  which  appeared 
(X1Tiax  3O6  mu)  remained  constant.  In  solutions  of  excess  strong  acid  (7.7M  acid:0.46M 
chloroamine)  no  new  absorption  appeared  during  reaction.   It  is  apparent  that 
dichloroamine  is  formed  during  photolysis  of  chloroamine  solutions  although  it  is 
unstable  under  the  reaction  conditions  and  does  not  accumulate  in  solution.  At 
lower  sulfuric  acid  concentrations,  dichlorobutylamine  formation  may  compete  favorably 
with  the  H-L  reaction  which  could  account  for  decreasing  yields  of  pyrrolidines  with 
decreasing  acid  concentration  (Table  I). 

Corey  found  that  the  rate  of  decomposition  of  chloroamine  increased  with 
increasing  sulfuric  acid  concentration  (Table  I,  entries  1-4)  while  Neale  stated  his 
results  (Table  I,  entries  8-1J?)  were  nin  direct  contrast"  to  Corey's  and  that  the 
rate  decreased  with  increasing  acid  concentration.  Neale  also  reported  that  in 
solutions  c_a.  0. 5M  in  chloroamine,  reactions  are  quite  slow  when  the  molar  ratio  of 
sulfuric  acid  to  chloroamine  was  1:1  or  2:1  (Table  I,  entries  16  and  17).  At  low 
acid -amine  ratios,  it  appears  that  both  researchers'  data  indicate  an  increase  in 
rate  with  an  increase  in  sulfuric  acid  concentration.  Neale' s  values  support  this 
conclusion  up  to  an  acid -chloroamine  ratio  of  3:1. 

Although  the  mechanism  of  the  H-L  reaction  is  generally  agreed  upon,  the 
initiating  species  is  subject  to  controversy.  Wawzonek  proposed  that  the  protonated 
N-chloroamine  is  the  initiating  species  since  the  chloroamine  exists  mostly  in  the 
protonated  form  in  sulfuric  acid  solutions.  Corey4  suggested  that  unprotonated  N- 
chloroamine  is  the  initiator.  Protonated  NCBA  shows  no  appreciable  absorption  above 
225  mu,  whereas  free  amine  absorbs  at  higher  wave  length  (A^£c267,  e   .  320) 17  (Fig  I). 


Fig  I.  N-chlorodi-n-butylamine: 
(1)  2.65  x  10~3M  in  HOAc,  cell 
1.0  cm j  (2)  I.98  x  10~3M  in  CCI4, 
cell  1.0  cm  j  (3)  0.46M  in  H2SO4- 
1.5M  H20-H0Ac;  cell  0.105  cm;  (4) 
0.46M  in  I.5M  H2SO4-I.5M  H20- 
HOAc,  cell  0.105  cm. 
N,N-Dichloro-n-butylamine :  ( 5) 
1.86  x  10~3M  in  ether,  cell  1.0  cm, 
from  NCS  and  butylamine;  (6) 
I.56  x  10~3M  in  pentane,  cell 
1.0  cm,  isolated  from  reaction 
mixture;  (7)  absorbance  of  2  mm. 
thickness  Pyrex  glass  vrs.  air. 


Absorbance 


325  mu 


-285- 
Corey's  data  show  that  the  rate  of  reaction  increases  with  increasing  sulfuric  acid 
concentration.   If  free  chloroamine  is  the  initiating  species,  its  concentration 
should  decrease  with  increasing  acidity  and,  hence,  decrease  the  initiation  rate. 
Thus  acid  catalysis  must  involve  acceleration  of  the  propagation  process  and/or 
retardation  of  chain  termination.   It  seemed  likely  to  Corey  that  strong  acid  should 
inhibit  chain  termination.  The  interaction  of  two  protonated,  positively  charged 
radicals  by  coupling  or  atom  transfer  would  be  slower  than  for  neutral  species, 
especially  if  the  ions  were  solvated.  The  photolytic  decomposition  of  NCBA  in  carbon 
tetrachloride18  and  anhydrous  acetic  acid,2'4  where  amine  is  unprotonated ,  is  very 
slow  relative  to  the  acid  catalyzed  reaction  (Table  I,  entry  1).  There  are  at  least 
three  possible  explanations  why  the  reaction  is  slow:  a)  Neale's  suggestion  that  the 
unprotonated  chloroamine  is  a  poor  initiator,  b)  Corey's  proposal  that  chain  termina- 
tion is  more  favorable  for  unprotonated  nitrogen  radicals,  and  c)  an  acid  catalyzed 
propagation  sequence  which  has  not  yet  been  defined. 

Neale8  suggested  that  the  N,N-dichloro-n-butylamine  is  the  initiating  species 
in  the  H-L  reaction,  since  it  is  formed  and  decomposed  during  photolysis  (^g^  3^6 > 
e  320).  He  further  argued  that  if  free  N-chloroamine  were  the  initiating  species, 
its  low  concentration  and  extinction  coefficient  would  require  it  to  absorb  light 
and  dissociate  very  efficiently  (Fig  I). 

N-HALOAMIDES  AND  N-HALOIMIDES 

The  mechanism  proposed  for  the  photolytic  decomposition  of  N-haloamides  and  N- 
haloimides  is  analogous  to  that  proposed  for  the  H-L  reaction  (eq  1).11,19,2°  These 
reactions  are  initiated  by  light  and  peroxides  and  are  inhibited  by  bubbling  oxygen 
through  the  reaction  solution.   One  notable  exception  is  that,  in  general,  acid  is 
not  needed  to  catalyze  the  reaction.  Although  reaction  conditions  vary,  hydrogen 
abstraction  at  the  /-carbon  is  predominant,  with  subsequent  formation  of  7 -lactones. 
Barton19  has  done  the  most  extensive  mechanistic  study  to  date.   In  a  search  for  a 
general  method  of  forming  saturated  lactones  from  saturated  acids,  photolysis  of  N- 
iodoamides  afforded  7-iminolactones  which  could  be  hydrolyzed  to  7-lactones. 
lodination  of  33-acetoxy-ll-oxo-5a:-pregnane-20-carboxamide  (IX)  with  lead  tetra- 
acetate and  iodine  in  benzene  gave  c_a.  55/^  yield  of  lactone  upon  alkaline  hydrolysis. 
In  a  similar  manner,  orthotoluamide  gave  phthalide  and  stearamide  gave  7-stearo- 
lactone.  The  following  mechanism  was  proposed:  /) 


CH3CO 


IX 


r 


Fb(0Ac)4-I. 


hv 


H 


f 


MH 

* 

I' 


m 

H 


0 


base 


HNI 


m 

H 


(2) 


X0^         XN^ 


Of  course,  one  cannot  rule  out  the  possibility  in  any  of  the  known  N-haloamide  or  N- 
haloimide  rearrangements  that  intramolecular  hydrogen  abstraction  is  performed  by 
amide  oxygen  rather  than  nitrogen,  followed  by  tautomeric  regeneration  of  the  normal 
amide  group  (eq  3) •   This  would  involve  the  same  size  cyclic  transition  state  as 
preferred  by  alkoxy  radicals  in  the  light  induced  radical  chain  decomposition  of 


-286- 

RtW  < >     RC=NR'  >    .RC=NR'   >   .RCNHR1  (3) 

t-butylhypochlorite s  with  side  chains  of  three  carbons  or  longer.  These  reactions 
occur  largely  via  an  intramolecular  path  (1,5  hydrogen  shift)  to  give  S-chloro- 
alcohols.24  An  intermediate  N-iodoamide  was  isolated  by  reacting  lead  tetraacetate 
and  iodine  with  benzamide.  N-Iodoamides  could  also  be  prepared  by  reacting  amides 
with  t-butylhypochlorite  and  iodine  in  various  solvents  where  the  iodinating  species 
was  shown  to  be  t-butylhypoiodite. 19  Crystalline  N-iodo  compounds  were  obtained  from 
benzamide,  succinimide,  n -butyr amide ,  n-hexanamide  and  n-octadecanamide  using  hypo- 
halite  as  the  iodinating  agent.  N-Iodo-n-octadecanamide  was  isolated  in  two  forms 
with  melting  points  of  Hk     and  120°.  Whether  these  were  two  crystal  forms  or  two 
geometrical  isomers  has  not  been  determined. 

Solutions  resulting  from  photolysis  of  N-iodoamides  showed  a  strong  infrared 
band  at  1680  cm"1.  This  suggested  the  presence  of  an  amide  or  iminolactone.  Washing 
these  solutions  with  sodium  hydrogen  sulfite  caused  the  appearance  of  a  /-lactone 
band  in  the  infrared  spectrum.  The  following  results  were  obtained  in  order  to 
determine  which  of  the  two  species  was  present  after  photolysis  and  before  hydrolysis. 
7-Iodobutyramide  was  found  to  be  unstable  at  room  temperature  in  a  "humid"  atmosphere 
and  cyclized  spontaneously  to  give  7 -lactone.  Photolysis  of  the  higher  melting  N- 
iodo-n-octadecanamide  showed  that  all  the  iodine  originally  present  as  N-iodo  was 
found  as  molecular  iodine  at  the  end  of  the  reaction.   Yields  were  never  greater  than 
50/0  unless  excess  iodinating  agent  was  used.  These  results  suggested  the  following 
sequence : 


DEI 


_>    ( 

-y 

>KH- 

HI        K-^ 

tt       1       f 

■\^NH 

r 

R 

+ 

V 

\y 

Ia 

During  the  initial  period  of  darkness  there  was  no  appearance  of  iodine.   Irradiation 
caused  a  rapid  development  of  iodine.  During  a  further  period  of  darkness  only  a 
small  amount  of  iodine  appeared  which  was  about  10$  of  the  amount  of  iodine  liberated 
during  the  light  period.  Further  periods  of  light  and  darkness  produced  the  same 
effects. 

The  solution  resulting  from  photolysis  of  N-iodo-n-octadecanamide  was  divided 
into  two  equal  portions.   One  portion  was  hydrolyzed  in  the  normal  manner  while  the 
other  was  treated  with  zinc  dust  and  acetic  acid  before  hydrolysis.  Both  solutions 
gave  the  same  yield  of  lactone  upon  hydrolysis.   It  has  been  shown19  that  7-iodo- 
butyramide  is  smoothly  converted  to  butyramide  under  identical  treatment  with  zinc 
and  acetic  acid.  These  results  suggest  that  iminolactone  is  responsible  for  the 
infrared  band  at  1680  cm"1.   Isolation  of  a  derivative  of  the  postulated  iminolactone 
was  finally  achieved  during  the  photolysis  of  7-phenylbutyramide  in  the  presence  of 
an  excess  of  t-butyl  hypochlorite  and  iodine.  The  crystalline  compound  which  formed 
during  the  photolysis  was  regarded  as  the  iodine  chloride  complex  (X)  of  N-iodo-7- 
phenylbutyroiminolactone.  Excess  iodinating  agent  reacting  with  7 -iminolactone 
could  be  responsible  for  the  formation  of  X.   Infrared,  NMR,  and  microanalytical 
data  support  the  structure  assigned  to  (X) . 


ft  J      Uni.ici 


0 

X 

Like  the  H-L  reaction,  optically  active  compounds  with  an  asymmetric  7-carbon 
rearrange  to  give  racemic  product.  Photolysis  of  optically  active  (+) -^-methylhexan- 
amide  (XI)  in  the  presence  of  t-butylhypochlorite  and  iodine  gave  racemic  k-methyl-k- 
hexanolactone  (XII) .  An  insertion  mechanism  was  thus  excluded. 

*    v> 0 

iuj.2 

XI  XII 


N-Bromoamides  and  N-chloroamides  can  also  be  phot olytic ally  rearranged  to  give 
7-lactones.  Rearrangement  is  most  efficient  with  N-t -butyl  derivatives  of  the  amides, 
Treatment  of  a  N-t-butylamide  with  10$  excess  t-butylhypobromite  in  carbon  tetra- 
chloride at  room  temperature  gave  the  N-bromo  compound.  On  subsequent  irradiation, 
active  bromine  was  lost  within  ten  minutes.  The  infrared  spectra  of  the  resulting 
solution  showed  a  strong  absorption  characteristic  of  secondary  amides.  Brief 
heating  afforded  the  iminolactone  (eq  5)  which  could  be  precipitated  by  dilution  with 
anhydrous  ether.  The  N-chloroamides  could  not  be  rearranged  as  readily  as  their  N- 


RCH2CH2CH 


Jj-t- 

Br 
XIII 


R 


C4H9 


*   W 


HBr 
N-t-C4H9 


(5) 


bromo  counterparts.  Heating  the  7-chloro  compound  in  sulfuric  acid  was  required  for 
ring  closure.  Table  II  lists  reaction  conditions  for  photolytic  rearrangement  of 
N-iodo-,  N-bromo-,  and  N-chloroamides. 


Table  II.  Photolytic  Rearrangement  of  N-Haloamides  and  N-Haloimides 


Compound 


AcO 


CONHR 


Temp 

Light 

Time, 

Substituent 

Solvent 

°C 

Source 

min 

io   7 -Lactone 

-H 

benzene 

15 

a 

300 

k6.b 

-CH3 

HCCI3 

reflux 

b 

90 

None 

-C6H5 

HCCI3 

reflux 

b 

2^0 

17.2 

CH3(CH2)3' 


CH3(CH2)3' 


CH3CH2CH 


J 


CH2CH2CH2' 
C&R5 


•H 

benzene 

2k 

c 

120 

37 

■t-C^Hg 

benzene 

26 

c 

10 

71 

CC14 

26 

c 

10 

79 

•C(C6H5)  3 

benzene 

25 

c 

10 

None 

•CH3 

benzene 

25 

c 

150 

None 

CCI4 

25 

c 

150 

None 

■C6H5 

benzene 

25 

c 

120 

None 

■CH3 

benzene 

30 

c 

20 

h3 

■t-C4Hg 

benzene 

23 

c 

25 

53 

•H 

CFC13 

0 

d 

18  hrs 

3 

■CCH3 

CFCI3 

0 

d 

8-9  hrs 

17 

-H 


-CCH< 


CFC13 


CFC13 


25 


8  hrs 


11  hrs 


19 


37 


125 -watt  high -pressure  mercury  arc, lamp,  tungsten  lamp,  Hanovia  100-watt 
medium  pressure  mercury  arc  lamp,  and  Rayonet  3500  A0  lamp  augmented  by  a  Victor 
500-watt  mercury  vapor  lamp. 

The  lower  yields  for  N-methyl  compounds  (Table  II)  could  be  attributed  to  dehydro- 
halogenation  resulting  from  loss  of  hydrogen  a  to  nitrogen  (N-C-H).   Irradiated  N- 
chloro-N-methylacetamide  (XIV)  reacted  completely  within  30  minutes  while  N-t-butyl- 
N-chloroacetamide  (XV)  was  unreactive  for  180  minutes  under  identical  conditions 


■288- 


£ 

XIV 


CH3CN-CH3 


CH 


3oij*-t-C4H9 

XV 


pentanoamide ,  which  in  common  with  the  N-t-butylamides  lack  a  hydrogens,  were 
unsuccessful.   While  N-bromo  compounds  could  be  rearranged  in  benzene  or  carbon 
tetrachloride,  N-chloro  rearrangements  were  successful  only  in  benzene  and  pyridine, 
This  solvent  effect  is  presently  being  investigated,25'26 

The  rearrangement  products  of  N-chloro-N-t-butylpentanoamide  and  N-chloro-N- 
t-butylhexanoamide  were  reduced  to  the  4-chloroamines  with  diborane,  Refluxing  for 
three  hours  gave  the  corresponding  pyrrolidine  (eq  6) .  This  is  another  way  of 


RCHCH2CH2CNH-t -C4H9 

ii 


BpH 


2n6 


[RCHCH2CH2CH2NH-t-.C4.H9] 

CI 


base 


9- 

t-C4,Hg 


R 


(6) 


preparing  H-L  rearrangement  products  without  using  strong  acids. 

The  hydrogen  abstracting,  chain-carrying  species  in  aliphatic,22  allylic  and 
benzylic23  halogenations  by  N-halosuccinimides  is  normally  the  halogen  atom,  rather 
than  the  succinimidyl  radical.   Since  intramolecular  rearrangements  are  generally 
more  rapid  than  the  corresponding  intermolecular  reactions,  it  is  possible  that 
acyclic  imidyl  radicals  might  rearrange  at  rates  fast  enough  to  permit  selective 
introduction  of  functional  groups  at  the  7-position  of  imides  (eq  7).20 


R— i 

R 


R 


N- 


R 


V    I 


R" 


R' 


T^X 


(7) 


R" 


XVI 


Petterson  has  shown  that  7-chloro-N-acetylamides  (XVI,  X=C1,  R"=CCH3)  are  formed 
from  compounds  having  primary,  secondary,  or  benzylic  7-hydrogen.  N-Chloroimides 
were  readily  made  Q>  9Cffo)   by  reacting  the  parent  imide  with  t-butylhypochlorite  in 
methanol.  4  Photolysis  under  helium  produced  ^--chloro  derivatives  which  were  con- 
verted to  lactones  by  acid  hydrolysis.  Table  II  shows  that  under  similar  conditions 
N-acetylamides  give  better  yields  of  7-lactone  than  do  the  corresponding  N-hydro- 
amides.  The  reason  why  the  second  carbonyl  group  on  nitrogen  promotes  rearrange- 
ment is  not  yet  known.   One  might  expect  from  Neale's  work  that  benzene  would  be  a 
better  solvent  for  imide  rearrangements.   Preliminary  studies  show  that  reaction 
times  in  benzene  are  decreased  but  yields  are  also  diminished.28 

CONCLUSION 

Respectable  yields  from  N-halo  cyclization  reactions  and  the  short  series  of 
operations  involved  in  the  reaction  further  spotlights  the  increasing  usefulness  of 
nitrogen  radicals  in  organic  synthesis.  These  reactions  serve  as  an  example  of  a 
free  radical  synthesis  which  may  be  difficult  or  even  impossible  by  a  nonradical 
approach.  The  mechanism  of  these  reactions  is  generally  agreed  upon,  although  the 
species  responsible  for  initiating  the  reactions  must  be  subject  to  further 
investigation.  Rearrangement  conditions  and  yields  are  dependent  upon  the  solvent, 
groups  substituted  on  nitrogen  in  amides  and  the  substituents  at  the  5-carbon 
(amines)  or  7 -carbon  (amides  and  imides). 


-289- 
BIBLIOGRAPHY 

1.  M.  E.  Wolff,  Chem.  Rev.,  6£,  55  (1963). 

2.  G.  R.  Wright,  J.  Am.  Chem.  Soc,  JO,   I958  (1948). 

3.  S.  Wawzonek  and  T.  P.  Culbertson,  J.  Am.  Chem.  Soc,  8l,  5367  (1959). 
k.  E.  J.  Corey  and  W.  R.  Hertler,  J.  Am.  Chem.  Soc.,  82,  1657  (i960). 

5.  S.  Wawzonek  and  P.  J.  Thelen,  J.  Am.  Chem.  Soc,  72,  2118  (I95O) . 

6.  R.  Lukes  and  M.  Ferles,  Coll.  Czech  Chem.  Comm. ,  20,  1227  (1955). 

7.  R.  S.  Neale,  Tetrahedron  Letters,  483  (1966). 

8.  R.  S.  Neale  and  M.  R.  Walsh,  J.  Am.  Chem.  Soc,  87,  1255  (1965). 

9.  F.  Minisce  and  R.  Galli,  Tetrahedron  Letters,  167  (1964). 

10.  F.  Minisci  and  R.  Galli,  Chim.  Ind.  (Milan),  46,  546  (1964). 

11.  R.  S.  Neale,  N.  R.  Marcus,  and  R.  G.  Schepers,  J.  Am.  Chem.  Soc,  88,  305I 
(1966).  ~ 

12.  R.  S.  Neale,  M.  R.  Walsh,  and  N.  L.  Marcus,  J.  Org.  Chem.,  £0,   3683  (1965). 

13.  A.  W.  Hofmann,  Ber. ,  16,  558,  586  (I883). 

14.  K.  Loeffler  and  S.  Kober,  Ber.,  42,  3427  (1909)  ',   K.  Loeffler  and  C.  Freytag, 
ibid.,  42,  3^27  (1909). 

15.  G.  H.  Coleman,  G.  Nichols,  and  T.  F.  Martens  in  "Organic  Synthesis,  Coll. 
Vol.  Ill,"  John  Wiley  and  Sons,  Inc.,  New  York,  N.  Y. ,  1955,  p.  159. 

16.  C.  Walling,  "Free  Radicals  in  Solution,"  John  Wiley  and  Sons,  Inc.,  New  York, 
N.  Y.,  1957,  Chapt  8. 

17.  W.  S.  Metcalf,  J.  Chem.  Soc,  148  (1942). 

18.  S.  Wawzonek  and  J.  D.  Nordstrom,  J.  Org.  Chem.,  27,  3726  (1962). 

19.  D.  H.  R.  Barton,  A.  L.  J.  Beckman,  and  A.  Gossman,  J.  Chem.  Soc,  181  (I965). 

20.  R.  C.  Petterson  and  A.  Wambsgans,  J.  Am.  Chem.  Soc,  86,  1948  (1964). 

21.  R.  S.  Neale  and  R.  L.  Hinman,  J.  Am.  Chem.  Soc,  8£,  2o*66  (1963))  R.  S.  Neale, 
J.  Am.  Chem.  Soc,  86,  5340  (1964). 

22.  P.  S.  Skell,  D.  L.  Tuleen,  and  P.  D.  Read,  J.  Am.  Chem.  Soc,  8£,  285O  (I963). 

23.  R.  E.  Pearson  and  J.  C.  Martin,  ibid.,  8£,  354  (1963). 
2k.     C.  Walling  and  A.  Padwa,  ibid.  ,  8£,  1597  (I963). 

25.  R.  C.  Petterson,  A.  Wambsgans,  and  R.  S.  George,  151st  National  Meeting  of  the 
American  Chemical  Society,  Pittsburgh,  Pa.,  March  1966,  Paper  No.  53. 

26.  G.  A.  Russell,  J.  Am.  Chem.  Soc,  80,  4987  (1958). 

27.  R.  S.  Neale,  private  communication. 

28.  R.  C.  Petterson,  private  communication. 


-290- 

THE  THERMAL  END0-EX0  ISOMERIZATION  OF  SOME  DIELS-ALDER  ADDUCTS 

Reported  by  Tommy  L.  Chaff in  April  6,  1967 

According  to  the  Alder  Rule1  of  Diels-Alder  additions,  of  the  two  stereo- 
isomeric  adducts  of  a  cyclic  diene  with  a  dienophile,  that  one  which  is  formed 
with  the  maximum  accumulation  of  double  bonds  will  preponderate.  Although  there 
are  many  exceptions  to  this  rule,  there  has  been  controversy  as  to  whether  they 
result  from  the  isomerization  of  the  initially  formed  endo  adduct  directly  to  to 
the  exo; 

Addends^  Endo^Exo 

or  whether  the  endo  product  is  reversibly  formed  from  the  addends  which  can 
slowly  form  the  thermodynamic ally  more  stable  exo  isomer; 

Endc^  Addends"^  Exo 

Any  type  of  direct  isomerization  which  does  not  involve  dissociation  into  kin- 
etically  free  addends  will  be  termed  an  "internal"  mechanism  in  contrast  to  an 
"external"  mechanism  such  as  dissociation  and  recombination.  This  seminar  will 
present  the  evidence  for  these  two  possibilities. 
ORIGIN  OF  THE  PROBLEM 

The  first  thermal  isomerization  of  this  type  was  reported  in  1933  by  Alder 
and  Stein2.   They  observed  the  isomerization  of  endo-dicyclopentadiene  (la) 
to  exo-dicyclopentadiene  (lb)  at  170°.   They  felt  that  this  was  due  to  the  existence 


lb 


of  an  equilibrium  with  a  small  concentration  of  the  monomer  under  the  reaction 
conditions . 

It  has  also  been  observed1 >3   that  addition  of  maleic  anhydride  to  6,6-pent- 
amethylenefulvene  at  room  temperature  produces  the  endo  isomer  while  at  higher 
temperatures  mixtures  of  the  isomeric  adducts  are  found.   This,  too,  was  explained 
on  the  basis  of  the  reversible  formation  of  small  concentrations  of  the  addends. 
Woodward  noted3  that  the  dihydro  derivatives  did  not  isomerize  and  that  a  solution 
of  the  adducts  in  benzene  or  ethyl  acetate  turned  yellow  when  warmed,  presum- 
ably from  the  colored  fulvene.   The  exo  isomer  was  considerably  more  stable  and 
therefore  the  reaction  was  thought  to  be  kinetically  controlled  at  low  temperatures. 
This  dissociation-recombination  mechanism  has  been  proposed4  to  be  a  general  one 
for  Diels-Alder  adducts  of  cyclic  dienes. 

Craig,  in  1951  >  reported5  that  heating  the  endo  adduct  of  maleic  anhydride 
and  cyclopentadiene  (Ila)  to  I9O0  produced  the  exo  isomer  (lib)  and  not  the  ad- 
dends as  had  been  previously  reported.6  He  concluded  that  the  isomerization 
proceeded  by  means  of  a  non-isolable  intermediate  and  noting  that  the  dibromo 
compound  did  not  rearrange,  he  proposed  the  following  mechanism  involving  the  double 
bond: 


H   0 


-291- 


-> 


Ha 


'.lib 


C02H 
C02H 


III 


This  mechanism  can  explain  why  the  addends  -./ere  observed  for  the  rearrange- 
ment of  the  fulvene -maleic  anhydride  adducts,  since  this  type  of  mechanism  would 
be  impossible  for  fulvene.7 

Schroder  has  proposed8'9  that  the  endo-exo  isomerization  of  dicyclopentadiene 
is  an  intramolecular  rearrangement. 
EXPERIMENTAL  EVIDENCE 

Berson  has  studied10  the  rate  of  formation  of  the  endo  and  exo  adducts 
(III)  from  furan  and  maleic  acid.   Contrary  to  a  previous  report1 x,  the  reaction 

does  not  produce  only  the  endo  isomer  .   Since  the  adducts  are 
quite  unstable12  they  were  isolated  by  saturating  the  double 
bond  with  bromine  to  prevent  retrogression.  If  both  isomers 
are  formed  directly  from  the  addends ,  then  the  rates  of  formation 
of  both  should  be  at  their  maxima  when  the  concentration  of 
the  addends  is  highest,  at  the  beginning.  This  is  assuming 
second-order  kinetics  for  both  isomers.   On  the  other  hand, 
if  the  exo  isomer  is  formed  directly  from  the  endo  without  forming  the  kinet- 
ically  free  addends,  then  the  endo  isomer  would  have  its  maximum  rate  of  formation 
at  the  beginning  and  the  exo  isomer  should  have  its  maximum  rate  at  some  later 
time  corresponding  to  the  maximum  concentration  of  the  endo  isomer.   The  data 
are  shown  graphically  in  Figure  1,  and  although  it  seems  to  support  the  direct 

isomerization,  the  analytical 
method  could  not  be  tested  on 
mixtures  of  known  composition 
due  to  the  instability  of  the 
endo  adduct,  and  the  authors 
chose  not  to  distinguish  between 
the  proposed  mechanistic  paths. 

It  has  been  reported5 
that  the  adduct  of  cyclopenta- 
diene  and  maleic  anhydride  (II) 
undergoes  diene  interchange  at 
200°  with  2,3-dimethylbutadiene 
and  dienophile  interchange  with 
fumaric  acid.   This  would  seem 
to  indicate  that  at  least  part  of  the  rearranged  product  results  from  retrogression,, 

Berson13-'14  studied  the  isomerization  of  the  endo  adduct  of  maleic  anhydride 
and  cyclopentadiene  (II)  with  C14  labeled  carbonyl  groups  in  the  presence  of 
an  equimolar  amount  of  unlabeled  maleic  anhydride  in  boiling  decalin  (I88.50) . 
Since  the  endo  adduct  exchanged  rather  rapidly  under  the  reaction  conditions, 
it  was  necessary  to  determine  the  activity  of  the  formed  exo  adduct  at  short 
reaction  times.  This,  in  turn,  meant  that  small  amounts  of  the  exo  adduct  were 
formed.  Therefore  the  amount  and  activity  of  formed  exo  adduct  were  determined 
by  isotopic  dilution.   If  the  isomerization  proceeds  by  purely  an  internal  or 
direct  isomerization,  the  activity  of  the  exo  adduct  formed  at  any  time  will 
be  the  same  as  the  endo  activity  at  that  time.   If  the  isomerization  proceeds 
by  an  external  path,  that  is,  by  retrogression,  then  the  activity  of  the  exo 
adduct  formed  at  any  time  will  be  the  same  as  the  activity  of  maleic  anhydride 


50 


100   150   200 
TIME,  hours 
Figure  1 


250   300 


-292- 
at  that  time.   By  dividing  the  reaction  into  arbitrarily  small  time  increments 
the  theoretical  activity  to  be  expected  of  the  exo  adduct  by  each  path  was  cal- 
culated by  graphical  integration.   The  activity  of  the  exo  adduct  isolated  was 
considerably  higher  than  that  expected  by  a  purely  external  path  and  therefore 
i t  vas  concluded  that  a  significant  part  of  the  isomerization  occurred  by  a  direct 
path  not  involving  kinetically  free  fragments.  They  felt  that  the  most  likely 
mechanism  involved  cage  recombination  or  some  intermediate  complex  of  the  ad- 
dends. It  has  since  been  demonstrated15  that  the  rate  of  addition  of  cyclopent- 
adiene and  maleic  anhydride  is  too  slow  to  compete  with  diffusion  and  therefore 
cage  recombination  is  not  a  reasonable  possibility, 

Baldwin  and  Roberts16  conducted  the  isomerization  of  the  endo  isomer  in 
fche  presence  of  tetracyanoethylene  (TCNE)  which  had  been  reported  to  be  a  very 
good  dienophile15.   The  assumption  was  that  TCNE  would  react  with  cyclopentadiene 
much  faster  than  would  maleic  anhydride.   Only  partial  inhibitiion  of  the  for- 
mation of  the  exo  isomer  was  observed  which  was  taken  to  indicate  that  an  inter- 
nal and  an  external  mechanism  were  in  competition. 

A  number  of  possible  internal  mechanisms  have  been  proposed  in  addition 
to  the  7^5-hydrogen  shift  suggested  by  Craig5.   They  fall  basically  into  two 
categories:  mechanisms  in  which  both  diene -dienophile  bonds  are  broken  with  the 
fragments  contained  in  a  solvent  cage  or  as  a  complex;  and  mechanisms  in  which 
any  other  bonds  are  broken.  In  the  second  category  are  included:  (l)   an  acid 
catalyzed  Wagner °Meerwe in  rearrangement17;  (2)   a  base  catalyzed  inversion 
(enoiization)18;  (5)   formation  of  a  nortricyclyl  derivative  (IV)  or  cyclopent- 
adienyl  succinic  anhydride  ( V)  5"19j2°-  (4)   and  cleavage  of  the  2^,3-carbon-carbon 
bond  to  form  a  common  intermediate-  possibly  a  double  cyclopropane  structure  (  vT*1! 


0^°  ^0 

IV  V  VI 

has  been  observed  that  the  rearrangement  is  not  catalyzed  by  acids  or  bases5 
which  would  seem  to  eliminate  the  first  two  proposals.   Roberts  and  co-workers21 
planned  to  study  the  internal  mechanism  further  by  rearranging  the  endo  adduct 
which  was  stereospecifically  labeled  in  only  one  carbonyi.   They  planned  to  use 
TCNE  as  a  scavenger  for  cyclopentadiene  so  that  all  of  the  exo  isomer  formed 
would  result  from  the  internal  mechanism.  It  was  hoped  that  the  position  of  the 
label  in  the  exo  isomer  would  eliminate  some  of  the  possible  internal  mechanisms. 
However,  it  was  found  that  maleic  anhydride  reacts  with  cyclopentadiene  at  a 
rate  comparable  to  that  of  TCNE  under  the  reaction  conditions.   An  alternative 
would  be  to  run  the  reaction  in  an  excess  of  maleic  anhydride  so  that  essentially 
all  of  the  exo  product  formed  by  the  internal  process  would  be  labeled  and  none   of 
that  by  the  external  process.   First^  however,  the  isomerization  was  carried 
out  again  on  the  adduct  with  uniformly  labeled  carbonyls22  to  re-examine  the  case 
for  the  internal  mechanism.   This  was  prompted  by  the  fact  that  TCNE  was  found 
not  to  be  an  effective  diene  scavenger  for  this  case  and  by  the  observation  that 
in  Berson's  exchange  experiment14  in  boiling  decalin  not  all  of  the  maleic  an- 
hydride was  in  solution.   Furthermore^  a  considerable  amount  of  the  maleic  an- 
hydride sublimes  out  of  the  reaction  within  a  few  minutes.   Both  of  these  exper- 
iments ,  then^,  prejudice  the  results  toward  an  internal  process.   When  essentially 
the  same  experiment  was  run  with  an  equimolar  amount  of  the  labeled  endo  isomer 
and  unlabeled  maleic  anhydride  in  t-pentylbenzene ,  which  gave  a  homogeneous 
solution,  the  results  indicated  that  no  internal  mechanism  was  involved  in  the 


„29> 
isomerization. 

In  other  work,  Miranov  has  shown23  that  isomerization  of  the  7-methyl 
(VII)  ,  1 -methyl  (VTIl),  and  6-methyl  (IX) -5-norhornene-2,3-endo-cis-dicarboxylic 
anhydrides  gives  a  2:1  ration  of  1-  to  6-methyl  adducts  and  only  a  trace  of  the 
7 -methyl  adduct.  Even  though  the  endo-exo  ratio  was  not  determined  this  is 
interpreted  solely  in  terms  of  dissociation  to  the  addends  since  substituted 
cyclopentadienes  have  been  shown  to  undergo  double  bond  migration  under  these 
conditions24.  Also  the  endo  adduct  of  maleic  anhydride  and  1,4-diphenyIcyclo- 


04 


V^ 


-L:;-^\- 

0 

1-4 

Q  I  >j 

0 

\>  V^ 

0 

VTI 


VIII 


IX 


X 


xr 


pentadiene  (Xj  has  been  reported25  to  give  a  1:3  ratio  of  1 -.'••■  exo  isomer  and  1,5- 
diphenyl-5-nor.oornene-endO'°cis°2,3'  dicarboxylic  anhydride  (XJ.  Although  the 
author  postulated  that  the  products  resulted  from  two  competing  mechanisms >  it 
appears  that  they  can  be  explained  in   the  same  way  as  those  of  Miranov, 

Baldwin26  has  studied  the  isomerization  of  specifically  deuterated  dicyc- 
lopentadienes  (i).  An  external  mechanism  should  produce  scrambling  of  the  deu- 
terium.  Statistically,  one  would  expect  25$  unlabeled,  25$  doubly  deuterated 
and  50/-'  monodeuterated  product  from  monodeuterated  starting  material.  An  inter- 
nal mechanism  would  preduct  all  monodeuterated  product.   This  will  obviously 
be  complicated  by  reversible  formation  of  the  monomer  from  one  or  both  of  the 
isomers  since  the  reaction  conditions  are  essentially  those  used  to  generate 
the  monomer  from  the  dimer27.   Therefore  if  any  labeling  specificity  is  found 
in  the  product  it  can  be  taken  as  evidence  for  an  internal  mechanism,  but  scram- 
bling of  the  label  could  be  interpreted  as  evidence  against  an  internal  mechanism 
only  if  it  can  be  demonstrated  that  the  specifically  deuterated  product  does 
not  equilibrate  under  the  reaction  conditions  and  that  the  starting  material 
has  not  equilibrated  before  reaction.  A  specifically  deuterated  exo  dimer  was 
recovered  after  90  minutes  at  196°  during  which  time  no  scrambling  had  occurred. 
The  endo  dimer  appears  to  have  undergone  about  60$  equilibration  at  the  end  of 
7  minutes  under  the  reaction  conditions.  The  results  of  the  isomerization  of 
the  endo  dimer  are  shown  in  Table  I, 


Table  I 


Deuterium  Distributions  in  Dicjyc lopentad iene s  upon  endo  to  exo 

Re  ar r  a njeme  ntatl96cr~" 


Time ,     Rearrangement , 

min.       °/o 


^endo  Dimer -\    /- 
'd0  dx  d2  d3  '    1 


exo  Dimer  — \ 
dQ  di  d2  d3  I 


0 
7 

60 
90 


o  6  8o  k  10 

1,1  26  57  16  l 

11  34  42  18  5 

16  67  20  9  3 


30  hi  19  4 
28  k-9   22  2 

31  47  2.1  1 


The  theoretical  distribution  of  deuterium  in  the  product  was  calculated 
on  the  assumption  that  both  the  doubly  and  the  triply  deuterated  endo  starting 
material  had  the  deuterium  in  one  cyclonentadiene  unit.   This  leads  to  a  cal- 
culated distribution  of  29.4^  d0.»  44.4$  d2, ,  \$?o   d2,  and  1%  d3.  This  is  in  good 
agreement  with  the  experimental  results,  and  for  short  reaction  times }   where 


equilibration  is  incomplete  in  the  starting  material,  would  seem  to  suggest 
little  if  any  contribution  from  an  internal  mechanism,,  Herndon  and  co-workers30 
have  arrived  at  the  same  conclusions  from  a  kinetic  study  of  the  thermal  decom- 
position of  the  dimers  in  the  gas  phase, 

Berson  studied  the  isomerization  of  optically  active  adducts  of  cyclo- 
pentadiene  with  methyl  acrylate  and  methyl  methacrylate19-'28.  Heating  the  op- 
tically active  exo  adduct  of  cyclopentadiene  and  methyl  methacrylate(XIIb)  at 
170°  in  decalin  for  3°  5  hours  gave  5°6/»  conversion  to  the  racemic  endo  adduct 
(XIIa)o   The  recovered  starting  material,  was  only  T/°  racemized  and  it  was  shown 
that  the  optically  active  endo  isomer  does  not  racemize  under  the  reaction 
conditions.  F>jrthermore ,  the  extent  of  conversion  corresponds  to  that  antic- 
ipated on  the  assumption  that  the  addends  are  common  intermediates  for  the  race- 
mization of  the  exo  isomer  and  its  conversion  to  endo.   Under  these  circumstances 
the  percent  conversion  should  have  proceeded  to  an  extent  equal  to  the  percent 

racemization  of  exo  times  the  kinetic  ratio(~^— )  for  formation  of  adducts  from 
addends.  exo 


C02CH3 

R 

C02CH3 

Xllaj  R=CH3 

Xllb; 

R=CH3 

XIIIa$  R=H 

Xlllb; 

R=H 

XI  Vb 


Similar  results  were  also  found  for  the  isomerization  of  the  optically 
active  endo  adduct  of  cyclopentadiene  and  methyl  acrylate (XIIl) .   These  results 
strongly  indicate  the  lack  of  an  appreciable  contribution  from  an  internal  mech- 
anism in  this  case. 

In  the  systems  considered  thus  far  the  isomers  have  been  chemically  dif- 
ferent and  it  has  been  necessary  to  follow  the  behavior  of  both  species,  how- 
ever for  the  optically  active  adduct  of  9-phenylanthracene  and  maleic  anhydride 
(XIV)  this  is  not  the  case  since  the  isomers  are  also  enantiomers15.   If  an 
internal  mechanism  exists  which  allows  interconversion  of  the  enantiomers ,  loss 
of  optical  activity  should  exceed  dissociation  into  kinetically  free  fragments. 
If,  however,  the  rate  of  loss  of  optical  activity  is  exactly  the  same  as  the  rate 
of  dissociation,  then  there  can  be  no  other  significant  path  for  racemization. 
The  loss  of  optical  activity  of  the  adduct  was  followed  as  a  function  of  time  and 
tee  first  order  rate  constants  determined  at  three  different  temperatures. 
The  rates  of  dissociation  were  determined  spectrophotometrically  by  means  of  the 
diene  ultraviolet  absorption  and  found  to  be  identical  with  those  for  loss 
of  optical  activity. 

One  final  example  of  this  Isomerization  has  been  studied29.  When  either 
the  endo  or  exo  adduct  of  cyclopentadiene  and  l,4-benzoquinone=2,3-epoxide  is 
heated  to  220°~for  10  minutes  the  resulting  mixture  consists  of  an  approximately 
equal  mixture  of  the  two  isomers  as  evidenced  by  a  comparison  of  the  infrared 
spectra  with  those  of  known  mixtures.   When  either  isomer  is  heated  under  the 
same  conditions  in  the  presence  of  an  equimolar  amount  of  TONE  only  1,4-benzo- 
quinone-2,3=epoxide,  292,3,3~tetracyano-5-norbornene,  and  starting  material  are 
produced. 
CONCLUSION 

In  the  light  of  these  studies  there  seems  to  be  no  firm  evidence  for  any 
mechanism  for  the  thermal  endo°exo  isomerization  of  Diels-Alder  aducts  other 
than  dissociation  and  recombination. 


-295- 

BIBLIOGRAPHY 

1.  K.  Alder  and  Go  Stein,  Angew.  Chem.,  50,  514  (1937)' 

2.  K.  Alder  and  G.  Stein,  Ann.,  504,  216(1933) • 

3.  R.  B.  Woodward  and  H.  Baer,  J.  Am.  Chem.  Soc.  ,  66,   645  (1944). 

4.  K.  Alder  and  W.  Trimborn,  Ann. .  566 y  58  (I.950). 

5.  Do  Craig,  J.  Am.  Chem.  Soc,  73,~TO9  (1951). 

60   M.  Kloetzei,  Organic  Reactions ,  Vol.  IV,  R.  Adams,  Ed.,  Wiley,  New  York, 
1948,  p.  9. 

7.  D.  Craig,  J.  J.  Shipman,  J.  Kiehl,  F.  Widmer,  R.  Fowler  and  A.  Hawthorne, 
J.  Am.  Chem.  Soc. ,  76,  4573  (1954). 

8.  W.  Schroder,  Angew.  Chem.,  72,  865  (i960). 

9.  Wo  Schroder,  Angew.  Chem.,  jQ,  24l  (1961). 

10.  J.  A.  Berson  and  R.  Swidler,  J.  Am.  Chem.  Soc,  7£,  1721  (1953). 

11.  0.  Diels  and  K.  Alder,  Ann.,  490,  243  (193L). 

12.  R.  B.  Woodward  and  H.  Baer,  J.  Am.  Chem.  Soc,  70,  ll6l  (1948). 

13.  J.  A.  Berson  and  R.  D.  Reynolds,  J.  Am.  Chem.  Soc,  77  4434  (1955). 

1.4.   J.  A.  Berson,  R.  D.  Reynolds  and  W.  M.  Jones,  J.  Am.  Chem.  Soc,  j^,  6049 

(  I956)  c 

15.  J.  A.  Berson  and  W.  A.  Mueller,  J.  Am.  Chem.  Soc,  83,  4940  (1961). 

16.  J.  E.  Baldwin  and  J.  Do  Roberts,  J.  Am.  Chem.  Soc.  85,  11 5  (1963). 

17.  P.  Bartlett  and  A.  Schneider,  J.  Am.  Chem.  Soc,  68,  6  (1946). 

18.  H.  Kwart  and  I.  Burchik,  J.  Am.  Chem.  Soc,  74,  3094  (1952). 

19.  J.  A.  Berson,  A.  Remanick  and  W.  A.  Mueller,  J.  Am.  Chem.  Soc,  82 ,  5501 
(I960). 

20.  R.  B.  Woodward  and  T.  J.  Katz ,  Tetrahedron,  5,  70  (I.959)  . 

21.  U.  Schiedegger ,  J.  E.  Baldwin  and  J.  D.  Roberts,  J.  Am.  Chem.  Soc,  89, 
894  (1967). 

22.  C.  Ganter,  U.  Schiedegger  and  J.  D.  Roberts,  J.  Am.  Chem.  Soc  ,  87.  2771 

(1965). 

23.  Vo  A.  Miranov,  To  M.  Fadeeva ,  A.  U.  Stepanyants  and  A.  A.  Akhrem,  Bull. 
Acad.  Sci.,  U.  S.  S.  R.  ,  Div.  Chem.  Sci.,  (Eng.  Trans.),  293  (1966)  . 

24.  V.  A.  Miranov,  E.  V.  Sobolov  and  A.  N.  Elizarova,  Tetrahedron,  ig,  1939  (1963) 

25.  K.  Leppanen,  Ann.  Acad.  Sci.  Fennicae,  Ser.  A  II,  131  (1965)$  Chem.  Abstr. , 
64,  14112  ( 1966) . 

26.  Jo'e.  Baldwin,  J.  Org.  Chem.,  Jl,  244l  (1966). 

27.  R.  R.  Moffett,  Organic  Synthesis,  Vol.  32,  Wiley,  New  York,  1952,  p.  4l. 

28.  J.  A.  Berson  and  A.  Remanick,  J.  Am,  Chem.  Soc,  8_3,  4947  (1961). 

29.  M.  J.  Youngquist,  D.  F.  O'Brien  and  J.  W.  Gates,  Jr.,  J,  Am.  Chem,  Soc, 
88,  4960  (1966). 

30.  W.  C.  Herndon,  C.  R.  Grayson,  and  J.  M.  Manion,  J.  Org.  Chem.,  J52,  526 
(1967)o 


-2Q6- 

REACTIQNS  OF  NH  RADICALS 

Terry  G.  Burlingame  April  10,  1967 

INTRODUCTION 

Imidogen,  or  NH,  may  be  considered  the  simplest  of  the  nltrenesj  i.e.,  that 
class  of  reactive  species  containing  monovalent  nitrogen,  which  therefore  contains 
six  electrons  in  its  outer  shell.  Several  reviews  on  nitrene  intermediates,  R-N, 
in  general,  have  appeared  in  the  literature. 1,a  Imidogen  is  isoelectronic  with  the 
much-studied  reactive  species  carbene,  or  CHS„3  Both  of  these  may  exist  either 
as  the  triplet  or  as  the  singlet  specie?,  (in  which  case  carbon  and  nitrogen  are 
quite  electrophilic  due  to  their  electron  deficient  structure) „  The  most  commonly 
used  method  to  generate  NH  is  through  the  phct ©decomposition  of  hydrazoic  acid,  HN3, 
in  which  the  well-established  primary  process  is  as  follows : 

hV 
H-H-Ns-N    -=-»    NH  +  N2 

Free  NH  was  first  detected  in  l892„4  Since  thatjiime,  the  species  has  been  well 
characterized  spectroscopically  in  the  gas  phase5  (by  absorption  and  emission)  and 
in  solid  matrices  at  low  temperatures6  (by  UV  and  IR  absorption) .  Less  definitive 
evidence  has  been  obtained  for  existence  of  NH  in  the  liquid  phase. 

The  electronic  states  of  NH  have  been  determined  spectroscopically.7 

Ground  state ;  triplet ,  3IT 
Lowest  excited  state;  singlet,  XA 
Second  excited  state;  singlet,  127' 
The  exact  energy  separation  of  these  states  is  not  known,  although  the  3£"  and  ^ 
levels  have  been  estimated  to  differ  in  energy  by  2]   kcal.7 

SCOPE 

The  purpose  of  this  seminar  will  be  to  discuss  in  some  detail  the  gas,  solid, 
and  liquid  phase  reactions  of  NH  radicals  with  other  molecules 3  to  examine  the 
mechanisms  proposed  for  those  reactions  in  which  NH  is  postulated  as  an  important 
intermediate,  and  to  discuss  the  validity  of  such  mechanisms  in  the  light  of  the 

experimental  evidence  presented. 

REACTIONS  OF  NH  IN  THE  VAPOR  PHASE 

The  most  systematic  analysis  to  date  of  the  reactions  of  NH  radicals  with 
various  organic  molecules  in  the  gas  phase  has  been  carried  out  by  Lwowski  and  co- 
workers.8 Reactions  of  methane,  ethane,  ethylene,  butene-1,  heptene-3,  and  2,3 
dimethyIbutene-2  with  NH  radicals  generated  by  the  phot ode composition  of  hydrazoic 
acid,  Hl3,  were  studied.  Two  sets  of  experiments  were  performed;  those  in  which  NH 
was  generated  by  high  energy  flash  photolysis  of   <  _  hydrocarbon  mixtures,  and  those 
in  which  steady,  slow  irradiation  of  the  mixtures  was  carried  out.  Flash  kinetic 
spectroscopy  performed  at  various  time  intervals  after  irradiation  in  the  flash 
experiments  gave  absorption  bands  (A3rt <-X32T)  due  to  the  triplet  ground  state  of  the 
.NH  radical  and  bands  from  highly  vibrationally  excited  °C«N  radicals.  The  C2   and  CH 
transients  were  detected  in  the  ethane  and  ethylene  reactions)  CH  was  also  detected 
in  the  methane  reaction.  There  is  a  close  correlation  between  the  decay  time  of  the 
NH  absorption  and  the  appearance  time  of  the  °CSN  spectrum.  No  bands  due  to  singlet 
NH  were  observed.  The  principal  nitrogen-containing  products  in  all  the  flash- 
initiated  reactions  were  N2  and  HCN.  No  alkyl  cyanides  were  detected.  The  slew 
photolyses  yielded  in  general  HCN,  alkyl  cyanides,  saturated  hydrocarbons  and  hydrogen 
in  the  gas  phase.  No  analyses  for  NH3  or  KHL4.N3,  aim*  ,r  certainly  products  of 
secondary  NH  reactions,  were  made.  For  example,  in  the  case  of  ethylene-HN3  mixtures, 
products  obtained  were  HCN,  CH3CN,  CH*,  H2  and  an  amorphous  solid.  The  HCN/CH3CN 
ratio  of  .9  was  independent  of  ethylene  pressure  in  the  range  8O-56O  mm.  Photolysis 
of  a  0.5 ;1  mole  ratio  mixture  of  DN3  and  HN3  gave  CH^DCN  and  CH3CN  in  a  ratio  of 
O.25.  'The  following  scheme  was  proposed  in  view  of  the  above  results  to  occur  in 
"hot.h  t.Vip  flash  and  alow  reactions* 


32~)     +     CH^CHo 


H2C-CH< 


H2G~CH3 


'f 


Further  fragmentation  of  the  nitre ne  intermediate  then  occurred  according  to  its 
energy  content  % 


H 


^D°CH3 


flash 


H 


reaction 


-»■    C3U     +     H.     +'C^K 


N 


A"Ui3       reaction 


H 


•IS 


,^={H2   +   CH3CM 

^HCN  +  CH4 


If  the  above  scheme  is  reasonable ,  one  should  expect  to  find  CH3CN  but  no  HCN  in  the 
steady  photolysis  of  HN3  +  2,3»dimethylbutene-2; 


}E3       CHj 


CHa-C— ( 


CH3  CH3 

CHq-C  —  C — CH 


HH    CH3 


CH3CU  +  (CH3)^CHCH3 
CaH6  +  (CH3)^C-CNN 


This  was  found  to  be  the  case.  Furthermore,  in  the  reaction  between  NH  and  butene-1 
addition  at  the  2-position  produced  C2H5CH  and  CH3CN  in  almost  equal  amounts,  suggest- 
ing that  the  intermediate  formed  after  1,3  hydrogen  transfer  had  a  lifetime  long 
enough  to  permit  vibrational  relaxation.   In  addition,  flash  and  slow  photolysis  of 
ethyl  azide  vapor,  C^HsNs,  gave  similar  results  to  the  ethylene  -NH  reactions, 
suggesting  that  the  same  intermediate  nitre ne  was  being  formed  in  both  cases.   Flash 
photolysis  gave  HCN,  CH3CN  (ratio  ^.9),  H2,  CH4  and  a  white  solid.   Steady  irradia- 
tion gave  CH3CN,  CJU,  H2  and  a  polymeric  gum..  ,  HON  seems  to  react  in  forming  the  gum. 
No  ethyleneimine ,  or  ethylamine,  which  result  from  solution  photolysis  of  C^E5N3,9 
were  found, 

Lwowski  and  co-workers  have  considered  in  detail  several  alternative  explanations 
which  may  also  be  used  to  rationalize  their  results s 

The  possibility  exists  for  the  formation  of  nitrogen  atoms  from  HN3.  Indeed, 
the  energy  requirements  of  the  process  HN3  •>  H°  +  °N  +  N2  are  well  within  the  limits 
of  that  supplied  by  the  2537  A  radiation.  Winkler  et.al.10  have  proposed  a  "unified 
mechanism"  for  the  reactions  of  nitrogen  atoms  with  simple  organic  molecules, 
illustrated  here  for  C3  molecules  s 


CH3c=CH~Cii2  \ 
CH3CH2CH2C1  ) 
CH3CH2CH2CH3  > 
CH2— G.H2   J 


+  N° 


XCH^ 


N«  +  "CaHs 
N-  +  CH3» 


CH3='CH  «'Cfl2 


HCN  +  CH3' 
HCN  +  2H« 


-»  HCN  +  C^H5< 


+  H< 


In  addition,  Dubrin  et.al.11  have  studied  the  reactions  of  methane  and  ethylene 
with  13N  atoms  produced  by  nuclear  techniques  and  have  concluded  that  the  reaction 
with  ethylene  follows  the  paths 


CsH^  +  N(TD) 


>>  HCN  +  °CH3 


Cyanide  radicals  were  ruled  out  on  the  basis  of  results  different  from  the  expected 
H2C-CH-C=N/HCN  ratio  of  4~7ol  obtained  on  addition  of  1XCN  to  ethylene.12  Lwowski 
and  co-workers  rule  intermediacy  of  nitrogen  atoms  in  their  system  out  since  (1)  no 
emission  or  absorption  bands  for  triplet  N2  resulting  from  If  atom  recombination  are 
observed.   (2)  There  is  close  correlation  of  the  decay  and  appearance  times  of  NH 
and  CN  respectively,  making  further  dissociation  of  IfH  unlikely. 

One  could,  argue  also  that  the  flash  and  slow  photoiyses  proceed  by  entirely 


-2Q8- 
different  mechanisms.  The  only  rationale  for  a  common  intermediate  given  by  Lwowski 
is  the  similarity  of  products  obtained  in  the  flash  and  slow  photolyses  and  the 
corresponding  photolysis  of  ethyl  azide,  differences  being  primarily  due  to  dif- 
ferences in  the  energy  content  of  said  intermediate  depending  on  its  mode  of  formation. 
However,  there  seems  to  be  no  evidence  at  present,  to  rule  out  the  formation  of 
vibrationally  excited  ethylene imine  folio-wed  by  decomposition; 

NH  4  H2C*CH2  _»  HsQ^—  CH2    *  >     CH3CN  +  HCN 
( singlet  J 

triplet)  (singlet  or  triplet) 

However,  in  analogy  to  the  reaction  of  methylene  with  ethylene  in  the  gas  phase  to 
produce  excited  cyclopropane  which  then  isomerizes  to  propylene ,  we  would  also  expect 
excited  ethylene  inline  to  isomerize. 

The  independent  work  of  two  other  research  groups  bears  an  important  relation- 
ship to  that  of  Lwowski,  Miller  and  Rice13  studied  the  system  HN3-ethylene, 
analyzing  for  all  products  in  efforts  to  obtain  a  complete  mass  balance.  Products 
identified  were  KH4N3,  ethane,  and  HCN  in  comparable  amounts,  and  smaller  amounts  of 
H2  and  CH3CN.  Formation  of  NH4N3  in  reactions  of  NH  with  both  alkanes  and  alkenes 
suggests  the  competing  sequence; 

NH  4  CaHi  (or  other  H  atom,  donor)      ■>  NH2  +  0^3- 
NH2  +  C^  (or  other  donor)      ■>  NH3  +  CaH3- 
NH3  +  HN3     ■>  NH4N3 

In  a  series  of  three  papers  Back14"16  and  co-workers  have  analyzed  the  flash 
and  slow  photolysis  of  isocyanic  acid  vapor,  H~N=C=0,  in  the  presence  and  absence  of 
various  hydrocarbons.  A  priori  the  photolysis  of  HNCO  and  HN3  should  be  related  in 
the  same  way  that  reactions  of  the  two  methylene  precursors  ketene,  CH2=C=0,  and 
diazomethane  CH2N2  are  related.  At  low  pressures  the  photolysis  of  isocyanic  acid 
gave  CO,  N2  and  small  amounts  of  H2  as  non  condensable  products.14  Although  C02, 
HCN,  C2N2,  NO,  N20  and  N02  could  have  been  detected,  they  were  not  found.   Small 
amounts  of  NH3  and  N2H4  were  detected  also.  The  mechanism  proposed  accounted  for 
most  of  the  observations; 

HNCO  -£¥— »  m    +  co 

NH  +  HNCO  — >  NH2  4  NCO 

NH2  4-  HNCO  —  *  NH3  4  NCO 
2NC0  _>  N2  4-  2C0 

The  observation  that  added  ethylene  reduced  the  yield  of  CO,  N2  and  H2  was  accounted 
for  by  the  scavenging  of  the  initially  formed  NH  by  the  ethylene. 

Lwowski  has  given  three  alternative  explanations  in  the  light  of  his  studies  to 
account  for  Back's  failure  to  observe  HCN; 

(1)  HCN  produced  from  the  added  olefins  reacts  with  HNCO; 

-°\ 

C-H 

N 


oxadiazoles 
(2)  NH  reacts  rapidly  with  HNCO  to  give  »NCO  which  then  attacks  ethylene; 
•NCO  4  H2C=CH2  — »  0=C=N-CH2CH2- 

(5)  HNCO  forms  a  relatively  long  lived  excited  state  which  reacts  with  C^.4 
to  give  ethyl  isocyanate  faster  than  it  dissociates. 

Brash  and  Back16  have  carried  out  a  more  detailed  study  of  the  steady  irradiation 
of  HNCO  vapor  in  the  presence  of  olefins  and  paraffins.  Increased  amounts  of  olefins 
reduced  the  N2  and  H2  quantum  yields  to  zero  and  the  CO  yield  to  a  constant  value 
indicating  complete  scavenging  of  NH  radicals.  No  imines,  amines,  or  other  nitrogen 


-299 
containing  products  could  be  found,  HON  was  again  not  detected.  HNCO  irradiated 
with  up  to  500  mm  of  butene-2  gave  no  HCN  or  imines.  'The  results  were  explained  on 
the  basis  of  rearrangement  and  polymerization  of  highly  vibrationally  excited  inter- 
mediates although  no  specific  analysis  of  polymeric  products  was  made.  Photolysis 
of  HNCO  in  the  presence  of  ethane,  propane,  and  neopentane  showed  the  same  general 
behavior  except  that  the  hydrogen  yield  increases,  and  products  of  radical  coupling 
are  found.  Small  amounts  of  added  ethylene  reduce  the  H2  yield  drastically, 
presumably  by  efficient  scavenging  of  H  atoms.  Photolysis  of  DNCO-C3H8  mixtures  and 
HNCO-D3H8  mixtures  gave  primarily  HD,  showing  that  each  molecule  of  H2  contained  one 
hydrogen  atom  from  HNCO  and  one  from  propane.  Insertion  of  NH  into  a  C-H  bond  to 
give  a  vibrationally  excited  amine  was  proposed  as  the  most  likely  process  leading 
to  product  formation °9   however,  production  of  hydrogen  atoms  by  this  process  seems 
unlikely  and  other  alternative  mechanisms  cannot  be  ruled  out  by  the  data. 

REACTIONS  OP  NH  RADICALS  IN  THE  SOLED  PHASE 

The  stabilization  of  reactive  species  and  their  subsequent  reactions  in  solid 
matrices  at  low  temperatures  have  been  reviewed  in  a  recent  seminar.17  Briefly 
reviewing  some  of  the  main  differences  between  gas  phase  and  matrix  reactions,  we 
find  the  following: 

(1)  Severe  translational  limitations  exist  for  species  formed  in  a  matrix 9 
diffusion  out  of  the  matrix  "cage"  is  severely  limited  for  all  but  the  smallest  of 
molecules  or  radicals. 

(2)  Molecules  or  radicals  formed  in  excited  electronic  and/or  vibrational  states 
by  photolysis  or  combination  with  other  reactive  species  may  be  rapidly  converted  to 
lower  energy  electronic  and/or  vibrational  states  by  frequent  collisions  with  the 
inert  matrix  cage  molecules.  Alternatively,  the  matrix  cage  may  be  a  reactive 
molecule  which  can  efficiently  add  to  and  trap  a  reactive  intermediate. 

An  early  study  by  Milligan  and  Jacox18  attempted  to  correlate  reactions  of  CH2 
and  NH  in  inert  argon  matrices.   Infrared  spectroscopy  was  used  to  directly  analyze 
the  products  from  the  photolysis  of  mixtures  of  HN3-ethylene -argon  and  HN3-acetylene» 
argon  at  4°K.  The  spectral  analysis  of  the  HN3 -ethylene -argon  system  shows  that 
ethyleneimine  is  the  sole  product  in  the  matrix.  "This  product  could  conceivably 
arise  from  singlet  or  triplet  NH  reacting  in  the  matrix. 


matrix 


H2C=CH2  +  NHl^A)  — ■*  \J  deactivation 

H 
..  .  *  (triplet) 

HoC-ch^  +  m(3z~)     nigh  enew  »  <r-7  sPin 

h2^-Lh2    +    m(   2,  )  -  \y  inversion 


H2C<!H2  +  NH(32T) 


more 


favorable 

H2C-CH2   matrix  ^  H2C-CH2 
ta  tNH 

The  last  pathway  is  probably  more  likely  since  NH  should  readily  be  deactivated  to 
the  triplet  ground  state  by  matrix  collisions.  However,  there  is  no  experimental 
evidence  other  than  Lwowski's  spectroscopic  results  in  the  gas  phase  which  allows 
one  to  distinguish  unambiguously  between  the  three  pathways.  Matrix  photolysis  of 
HN3»eis-2-bute:ae  and/or  HN3 -trans -2-butene  might  help  establish  the  identity  of  the 
reacting  species.  Experiments  with  methylene  from  photolysis  of  diazomethane19 
have  shown  that  it  is  initially  formed  as  an  excited  singlet  which  is  subsequently 
deactivated  to  the  lowest  singlet  and  after  an  order  of  magnitude  more  collisions, 
to  the  ground  triplet.  Addition  of  triplet  CH2  to  ethylene  followed  by  spin  inversion 


-z.r\r\ 


accounts  for  formation  of  cyclopropane  in  matrix  reactions  to  be  contrasted  with 
exclusive  propylene  formation  in  the  gas  phase3  where  vibrationally  excited  singlet 
cyclopropane  is  initially  produced,  then  rearranges. 

A  somewhat  more  interesting  result  is  obtained  in  the  HN3~acetylene -argon 
photolysis.   Singlet  NH  would  be  predicted  to  react  as  follows: 


nh  +  hcech 


H~C 


=  C-H 


H 
H^N-ChC-H 


These  species  could  undergo  further  rearrangement  to  acetonitrile  or  methyl  iso- 
cyanide.  Triplet  NH  could  add  in  the  following  manner: 

HC-C 

t  im 

-C«H 


These  products  could  further  rearrange: 

CH.2=C=NH 


-»  HaC=C=MH 


azacyclopropene 

The  infrared  spectral  analysis  rules  out  aminoacetylene ,  acetyleneimine ,  and  azacyclo- 
propene 0  The  same  data  strongly  indicate  the  presence  of  acetonitrile  and  methyl ~ 
isocyanide  as  well  as  the  previously  unobserved  species  keteneimine,  suggesting  that 
both  singlet  and  triplet  NH  could  be  presnet.  The  analogous  matrix  reaction  of  CH2 
with  acetylene  produces  allene  almost  exclusively. 

The  photolysis  of  HNCO  and  DNCO  in  argon  and  nitrogen  matrices  at  k°   and  20°K 
has  been  carried  out.20  Infrared  analysis  reveals  results  quite  different  from 
those  obtained  in  the  gas  phase  in  that  little  NH  or  CO  is  spectroscopically 
detectable  and  assignment  of  new  bands  seems  consistent  with  the  species  H-O-C^N. 
Two  mechanisms,  each  assuming  a  different  primary  process,  are  proposed: 


( 1)      HNCO 


hv 


[H«     +     °NCO] 
matrix  cage 


diffusion 


H«      +      [°NCO] 


recombination 


»  HNCO  +  HOCN 


In  this  scheme,  enough  H  atoms  should  escape  the  cage  to  make  °NCO  observable  by  IR 
spectroscopy^  however,  no  °NCO  is  observed  by  infrared  spectral  analysis. 

(2)  HNCO  — »  [NH  +  CO] 

HNCO 
NH(32)  +  :c=o  »  :c — 6 


B 


-301- 

Evidence  for  the  second  pathway      ven  by  the  fact  that  in  a  separate  experiment 
photolysis  of  an  argon- C0-HN3  mixture  gave  good  yields  of  both  HNCO  and  HOCN. 
Analogous  to  this  reaction  in  carbene  chemistry  is  the  photolysis  of  N2  +  CO  +  CH2N2  in 
a  matrix  at  20°K  with  the  production  of  high  yields  of  ketene.3 

When  HN3  is  photolyzed  in  a  matrix  composed  of  solid  C02  and  the  product 
analysis  is  carried  out  by  direct  infrared  analysis  on  the  matrix,  two  distinct 
groups  of  bands  are  observed.21  One  group  increases  in  intensity  during  the 
irradiation,  the  second  group  rapidly  reaches  a  maximum,  then  decreases.  Control 
experiments  showed  that  NH  radicals  did  not  diffuse  through  the  C02  cage  to  give 
spectroscopically  observable  amounts  of  NHS.  Photolysis  of  RN3  in  an  N20  matrix 
gave  group  I  bands  but  none  of  the  second  group,  indicating  that  the  latter  arose 
from  an  intermediate  NH-C02  adduct.  Further  analysis  identified  group  I  bands  as 
H-N=0.  The  characteristic  absorptions  of  CO  were  also  identified  in  the  C02  matrix. 
Isotopic  substitution  using  C1302  and  C028  showed  definitely  that  a  carbon  containing 
species  was  responsible  for  group  II  bands 0  The  observation  of  pairs  of  bands  with 
different  growth  rates  in  similar  regions  of  the  transient  spectrum  strongly  suggests 
rapid  rearrangement  of  the  initial  intermediate  or  formation  of  an  intermediate 
capable  of  c is -trans  isomerization.  Possible  intermediates  are  the  following; 


0 


^o 


H 


/ 


BT 


-C-0 


Jj-N 


H' 


(a) 


(b) 


V 


(c) 


V 


0-0 

H     % 

(a) 


Analogies  for  (a)  exist  in  the  reaction  of  CH2  with  C02  in  a  matrix,3  (b)  could  under- 
go a  cis -trans  isomerization  about  the  N-0  bond  analogous  to  alkyl  nitrites.  Structure 
(cj  is  ruled  out  by  the  infrared  analysis  and  the  fact  that  there  was  no  comparable 
product  (glycxal)  produced  in  the  CH2-C02  reaction.  Rearrangement  to  (d)  would 
involve  unfavorable  movement  of  heavy  atoms  rather  than  simple  hydrogen  transfer. 

Interesting  insight  into  the  reacting  NH  species  in  matrix  reactions  may  be  given 
by  the  photolysis  of  mixtures  of  HN3  and  02  in  solid  nitrogen  at  20°Ko  22  Infrared 
analysis  of  the  products  formed  indicates  that  both  cis  and  trans  nitrous  acid, 
KO-N-0  are  initially  formed  which  undergo  further  trans ^cis  isomerization  by  UV 
radiation  and  cis^trans  isomerization  by  the  infra-red  team  of  the  spectrophotometer. 
Evidently  NH  readily  reacts  with  '      be  produce  HONO.  However,  the  alternate  sequence 

HN3*  +  02  — ¥    HONO  +  N2 
could  not  be  ruled  out  at  the  time0  Reactions  of  CH2  in  the  gas  phase  are  not 
affected  by  added  02  except  when  high  pressures  of  inert  gas  are  used,  indicating 
that  singlet  •»  triplet  deactivation  occurs  by  collision  with  the  gas,  the  reaction 
of  triplet  CH2  with  oxygen  then  occurring  readily. 

LIQUID  AT©  SOLUTION  PHASE  REACTIONS  INVOLVING  NH  RADICALS 

It  appears  that  the  evidence  involving  NH  formation  in  solution  reactions  of 
organic  molecules  is  less  conclusive  than  in  the  gas  and  solid  phase  reactions 
discussed  thus  far.  There  are  three  reagents,  the  decompositions  which  in  solution 
phase  are  proposed  in  some  cases  to  produce  NH  radicals.  These  are  (1)  hydrazoic 
acid  or  azide  ion  under  appropriate  photolytic  conditions  (2)  hydroxylamine-O- 
sulfonic  acid  H2N-0~S0,3H  under  alkaline  conditions,  and  (3)  chloramine,  H^NCl,  under 
alkaline  conditions.  We  shall  take  each  of  these  reagents  in  turn  and  compare 
examples  where  NH  has  been  considered  an  important  intermediate  with  those  where 
intermediacy  of  NH  radicals  has  more  or  less  conclusively  been  ruled  out. 

The  thorough  work  of  Burak  and  Treinin23  has  shown  that  NH  radicals  can  indeed 
be  produced  in  solution.  The  photolysis  of  degassed  aqueous  solutions  of  NaN3  with 
2537  A0  light  produces  N2,  NH20H,  H2,  NH3  and  N^Ii*.  Added  NH3  causes  a  large 
increase  in  the  N2H4  yield,  paralleled  by  a  corresponding  decrease  in  the  NH20H 
yield.  The  following  mechanism  was  proposed  by  the  authors  to  account  for  the  results: 


hv  . 


jfe 


»3 


+  H20 
Mo*  — 


* 
* 


-»  HN3"  4= 
NH  +  No 


OH 


-302  - 

The  ultimate  fate  of  the  NH  radicals  is  represented  in  the  following  scheme 

W    +     H20  — — ->  NH20H 
H20  +  Mi     +     N3  — >  N2  +  N2H2  +  OH" 


2E2H, 


N2H4  +  N2 


HN3  +  WH3 


The  authors  propose  that  singlet  NH  is  the  reactive  species,  acting  as  a  strong 
Lewis  acid  in  its  reactions  with  H20,  NH3,  and  Ng.  The  quantum  yield  of  N2  was  not 
affected  by  such  radical  scavengers  as  N20,  acetone  or  methanol -phosphate  buffer, 
which  results  rule  out  mechanisms  involving  chain  reactions  or  solvated  electrons. 
The  dependence  of  the  quantum  yields  of  N2  and  NH20H  on  azide  ion  concentration 
suggests  that  there  is  competition  between  Ng  and  H20  for  the  NH  radical.  Excess 
added  ammonia  also  exerts  a  marked  scavenging  effect  as  shown  by  its  effect  on  the 
quantum  yields.  Calculated  ratios  of  rate  constants  for  NH  scavenging  are  H20;NH3; 
NJ  -  l:l8;285«  The  observations  that  the  quantum  yields  are  independent  of  the 
light  intensity  and  that  smaller  concentrations  of  impurities  including  oxygen 
have  no  effect  on  ID  NH20H  show  that  the  NH  radicals  produced  are  quickly  scavenged 
by  the  large  excess  of  H20  present. 

An  early  study24  showed  that  reactions  suggestive  of  NH  formation  could  be 
carried  out  in  organic  solutions.  Hydrazoic  acid  was  irradiated  in  the  presence  of 
toluene  solvent  to  yield  presumably  mixed  toluidines.  The  isomer  ratio  was  not 
determined 3  derivative  formation  alone  was  used  to  confirm  the  presence  of  toluidines. 
The  only  control  reaction  run  was  that  of  a  dark  reaction  which  gave  no  product 
formation  over  a  12  hour  period.   In  a  later  more  systematic  study25  on  the  solution 
photolysis  of  azides,  hydrazoic  acid  was  irradiated  in  the  presence  of  benzene  to 
give  low  yields  of  aniline.   Photolysis  of  n-butyl  and  n-octyl  azides  in  benzene 
produced  comparable  amounts  of  N-n«butyl  and  N-n~octyl  anilines  respectively.  No 
mechanisms  for  these  reactions  were  proposed  although  presumably  if  NH  and  nitrenes 
were  involved,  reaction  would  proceed  by  insertion  of  N~H  or  N-R  into  a  C-H  bond. 
It  is  interesting  that  photolysis  of  diazomethane  in  benzene  solution  yields  32/0 
cycloheptatriene  and  9$  toluene;3 


+   :CH2    * >         L  \\  + 


However,  none  of  the  analogous  ring-expanded  product  1-H-azepine  was  searched  for 
in  the  corresponding  NH  experiment. 

Hydroxylamine-O-sulfonic  acid^,  H2N~0-S03H,  has  been  found  to  produce  some  very 
interesting  reactions  in  recent  years,  particularly  a  series  of  so-called  "imination" 
reactions  as  illustrated  by  the  following  schemes26 

. ,,,  ^H 

(   V=0   +   NH20S03H     °H  -> 

Formation  of  NH  followed  by  addition  to  the  00  bond  could  be  envisioned.  However, 
the  evidence  available  suggests  that  reaction  takes  place  through  undissociated 
H2NOSO3H. 


O 


0      +    HgN0S03g     -*£-*   (       V^J^L_  0     — >      (        Kj       +     S°4@ 

H 


Evidence  against  dissociation  of  H^OSO.aH  to  NH  radicals  is  the  observation  that  its 
rate  of  reaction  in  NaOH  with  various  added  nucleophiles  depends  on  the  nucleophile 
used  which  would  not  be  true  of  NH  formation  were  the  rate  determining  step.   Studies 
of  the  reaction  of  HI  with  H^OSGaH28  have  shown  that  nucleophilic  attack  on  nitrogen 


zrcz 


occurs  to  give  INH2  initially.  Alkyl  groups  on  the  nitrogen  atom  slow  the  reaction 

considerably .  Thus,  a  claim  that  HH  from  H2HOS03B'  was  trapped  by  reaction  with 

cyanide  ion  and  precipitated  as  silver  cyanamide27  may  be  better  explained  by  direct 

nucleophilic  attack  of  CN~  on  the  acid.  Furthermore,  H2NOSO3H  does  not  incorporate 

radioactive  sulfur  from  radioactive  sulfate  solutions  thus  ruling  out  an  initial 

equilibrium.  q 

0    OH       @ 
HsNOSOs   ^-^   SO4  +  HH 

In  an  interesting  reaction  reported  by  Appel  and  Btiehner29  however,  we  find  it 
hard  to  rule  out  direct  participation  of  HH  radicals.  The  reaction  of  H2NOS03H  in 
sodium  meth oxide -methanol  with  butadiene  gave  a  low  yield  of  ^-py^sline..  This 
result  is  suggestive  of  a  1,4  addition  of  HH: 

r      +     MH   — — ->    U       n-H 

Other  added  olefins  had  produced  no  detectable  amounts  of  aziridines.  It  is  more 
likely  that  1,2  addition  of  HH  actually  occurs  first  followed  by  rearrangement  of  the 
vibrationally  excited  aziridine  to  3  pyrroline.  This  sequence  is  not  without  analogy 
in  carbene  chemistry^  addition  of  CH2  to  butadiene  in  the  gas  phase  produces  vinyl- 
cyclopropane  as  the  major  product  and  smaller  amounts  of  cyclopentene.3  The  cyclo- 
pentene  arises  presumably  from  isomerization  of  excited  vinylcyclopropane^  at  least 
it  has  been  shown  that  vinylcyclopropane  readily  undergoes  this  isomerization 
thermally. 

Chloramine,  HH2C1,  has  also  been  utilized  recently  in  a  number  of  interesting 
"imination"  reactions.  ° 

HH  has  been  shown  to'  be  a  product  of  the  photolysis  of  solid  chloramine  at  low- 
temperatures,  and  of  the  thermal  decomposition  of  gaseous  chloramine.  However,  in 
the  solution  reactions  of  the  compound,  the  rate  of  product  formation  is  dependent 
on  the  nature  of  the  substrate  being  attacked,  which  would  not  be  the  case  if  HH 
formation  were  rate -determining „  Thus,  these  reactions  seem  to  proceed  by  direct 
nucleophilic  additions:31 

<^J^H-CeH11   +   HH2C1   — >   (_/C|     — »  W^H 

cf  \  H 

BIBLIOGRAPHY 

1.  R„  A.  Abramovitch  and  B.  A.  Davis,  Chem.  Rev.,  &*_,  lk$   (1964). 

2.  L„  Horner  and  A.  Christmann,  Angew.  Chem.  Intern.  Ed.  Engl.,  2,  599  (I963). 

3.  For  an  extensive  review  see  ¥.  Kirmse,  Carbene  Chemistry,  Academic  Press, 
Hew  York,  1964,  ch.  2.  For  a  critical  analysis  of  the  gas  phase  reactions 

of  methylene  see  J.  Bell  in  Prog,  in  Phys.  Org.  Chem.,  S.  Cohen,  A.  Streitwieser, 
Jr.,  R.  W.  Taft,  Ed.,  Interscience,  Hew  York,  1964,  Vol.  2,  p.  1. 

4.  J.  M.  Eder,  Monatsh. ,  12,  86  (1892). 

5.  B.  A.  Thrush,  Proc.  Roy.  Soc.  (London),  A£J£,  143  (1956). 

6.  D.  E.  Milligan  and  M.  E.  Jacox,  J.  Chem.  Phys.,  kl,   2838  (1964). 

7.  G.  Herzberg,  Spectra  of  Diatomic  Molecules,  2nd  ed.,  Van  Hostrand ,  New  York, 
1950. 

8.  D.  W.  Cornell,  R.  S.  Eerry,  and  W.  Lwowski,  J.  Am.  Chem.  Soc.,  88,  544  (1966). 

9.  w.  H.  Saunders  and  E,  A.  Caress,  J.  Am.  Chem.  Soc.  9   86,  86l  (19oTT). 


■*,04- 


10.  H.  Go  S.  Evans,  G.  R.  Freeman,  and  C.  A„  Winkler,  Can.  J,  Chem, ,  3>4,  1271 

(1956)o 
H.  J.  Dubrin,  R.  Wolfgang,  and  C.  McKay,  J.  Chem.  Phys.,  40,  2208  (1966). 

12.  Jo  Dubrin,  C.  McKay,  M.  L,  Pandow,  and  R.  Wolfgang,  J.  Inorg.  Nucl,  Chem., 
26,  2113  (1964). 

13.  E0  D.  Miller,  Ph.D.  Dissertation,  Catholic  University  of  America,  Catholic 
University  of  America  Press,  Washington,  D.Co,  I96I. 

14.  J.  Y.  P.  Mui  and  R.  A.  Back,  Can.  J.  Chem.,  kl,   826  (I963). 

15.  R.  A.  Back,  J.  Chem.  Phys.,  40,  3493  (1964). 

I60  J.  Lo  Brash  and  R.  A.  Back,  Can.  J.  Chem.,  k%,   1778  (I965). 

17.  J.  Billet,  Univ.  of  111,  Organic  Seminar  Abstracts,  1966,  p.  128, 

18.  M.  E,  Jacox  and  D.  E.  Milligan,  J.  Am.  Chem,  Soc,  85,  278  (I963). 

19.  G.  Herzberg,  Proc.  Roy.  Soc.  (London),  A262,  291  (19ol). 

20.  M.  E.  Jacox  and  D.  E.  Milligan,  J,  ChemTTEys.,  kO,   24^7  (1964). 

21.  D.  E.  Milligan,  M.  E.  Jacox,  S.  W.  Charles,  and  G.  C.  Pimental,  J.  Chem,  Phys,, 

2L,  2302  (1962). 

22.  Jo  Do  Baldeschwieler  and  G„  C.  Pimentel,  J,  Chem.  Phys.,  33,  1008  (i960). 

23.  I.  Burak  and  A.  Ireinin,  J.  Am.  Chem.  Soc.,  87,  4031  (1965)°. 

24.  R,  K.  Keller  and  P.  A.  S.  Smith,  J.  Am.  Chem.  Soc,  66,   1122  (1944). 

25.  D.  H.  R.  Barton  and  L.  R.  Morgan,  Jr.,  J.  Chem.  Soc,  622  (1962). 
2c  E.  Schmitz,  R.  Ohme,  and  D.  Murawski,  Angew.   Chem.,  73,  708  (I96I)  . 

27»  G.  Bargigia,  Atti  Accad.  Nazi.  Lincei,  Rend.  Classe  Sci.,  Fis.,  Mat,,  Nat., 
£,   587  (1964). 

28.  P.  A,  So  Smith,  H.  R.  Alul,  and  R.  L0  Baumgarten,  J.  Am..  Chem.  Soc,  86,  U39 
(1964)  o  ™ 

29.  R.  Appel  and  0.  Bftchner,  Angew.  Chem.  Intern.  Ed.  Engl.,  1,  332  (1962)0 

30.  E.  Schmitz  and  R.  Ohme,  Ber, ,  %£,  2X66   (1961) . 

31.  E.  Schmitz,  Angew.  Chem.  Intern.  Ed8  Engl.,  £,   333  (1964). 


-305- 

GEOMETRIC  ISOMERISM  IN  DIAZOKETONES 

Reported  by  Daniel  B.  Pendergrass  April  13,  1967 

Although  diazoketones  (i)  vere  known  as  early  as  2.894(1),  they  were  not 
widely  used  "by  the  synthetic  organic  chemist  until  L.  Wolff (2)  reported  their 
rearrangement  to  the  corresponding  ketenes.  This  rearrangement  is  the  charac- 
teristic step  In  the  Arndt-Eistert  synthesis  by  which  a  carboxylic  acid  may  be 
converted  to  its  next  highest  homolog  or  one  of  its  derivatives^). 

R-C02H  ■>  R-C0-C1  (a) 

R-C0-C1  +  2  CH2N2  ■>  R-C0=CKN2  +  CH3C1  +  N2  (b) 

R~C0~CHN2  +  H-R5^R=CH2-CO-R5   +  N2  (c) 

r«    =  -OH    ,    -OR"    ,    -NHR"    ,   or  -NH2 

The  thermolysis (4,5)  and  photolysis  (5.)  of  diazoketones,  as  well  as  their 
decomposition  in  the  presence  of  ae.ids(6,7>8) ,  bases(5,9).?  and  various  metals 
(3>5>10)  have  been  discussed.   The  literature  and  reactions  of  diazobxides  have 
also  been  reviewed(  5.»ll) » 

The  most  widely  known  synthesis  of  diazoketones  is  the  reaction  of  an 
acyl  halide  with  two  equivalents  of  diazomethane  or  one  of  its  substituted  de- 
rivatives.  There  exists  in  the  literature  a  variety  of  representations  of  the 
charge  distribution  and  bond  orders  in  these  compounds  as  illustrated  below. 

R-CO-X  +  R'-CHN2  ->  R-CG-CN2 

L     1 

$  9     0  -©        $ 


R-C-C=N=N  i-f  R-C-C-N5N  <— ►  R-C=C-N=N 

II  »,  I*  J,  f  I, 

0  R'         0  R!        ©0  R! 


These  compounds  have  also  been  prepared  by  the  oxidation  of  a  monohydra- 
sone  of  a  dik.etone(  12,13,14,15 ,16,17)  obtained  in  a  variety  of  ways.  They  are 
frequently  products  of  nitrous  acid  oxidation  of  amines  alpha  to  a  -CO-R  group 
(18,19),  solvolytic  attack  on  the  monotosylhydrazone  of  a  diketone( 20,21,22,23,24) 
in  the  presence  of  base,  or  the  action  of  chloramine  or  hydroxylamine-O- sulfonic 
acid  on  an  oximinoketone(25) .  G.  R.  Harvey(26)  has  reported  that  a  series  of 
compounds  with  structure  II  react  with  pj-toluenesulfonazide  in  methylene  chloride 
to  give  the  corresponding  diazoketones. 

^P^C-CD-R'  +  N3S02_/q\-  CH3  ->  R:-C0~C-N2  +  <t>3P=NTs 

R 


li 


II 

Until  recently,  there  were  few  references  in  the  literature  to  the  absorption 
spectra  of  diazocarbonyl  compound s( 27,28) .  In  the  ultraviolet  region,  there 
appeared  to  be  a  characteristic  band  at  245-250  m\i   for  diazoesters  and  diazo- 
ketones (29,30) .   In  the  infrared  spectra,  an  abnormally  low  carbonyl  frequency 
(1630-1660  cm"1)  has  been  attributed  to  the  contributions  from  the  species  with 
the  charge  distributions  shown  above.      A  strong  band  at  1535-1410  cm  1, 
not  observed  in  the  diazohydrocarbons,  must  be  the  symetric  stretching  mode  of  =CM 
which  has  been  shifted  to  higher  frequency  by  a  high  degree  of  conjugation(3l)  -> 

Earlier  observations  of  the  ultraviolet  spectra  of  diazoketones  were  con- 
firmed by  Miller  and  White(32)  when  they  investigated  the  spectra  of  a  series 
of  diazocarbonyl  compounds.   Solvents  were  found  to  have  only  a  slight  effect  in  most 
of  the  casBs  considered..  However  1  8-bisdiazo~2  ,7— octanedione  (III'  gave  two  bands „ 
one  at  247  mu  and  another  at  273  &M-°  The  247  mu  band  is  observed  in  nonhydroxylic 


-306- 
solvents.   In  hydroxylic  solvents,  the  "band  at  2^7  mu  is  weakened  and  a  second  band 
at  273  W-   appears.   See  Figure  1(32).   The  energy  difference  is  about  11  kcal/raole 


between  the  two  bands.   A  difference  of  25 
shift  normally  observed( 33) • 


mu  is  much  larger  than  the  5-10  mu  solvent 


o 

H 

X 


Figure  I 


10 

8 
6 


l.Dioxane   \ 

2.Cyclohexane 

3»CH3CN^CH2C12 

4.CHCI3 

5.C2H5OH 


1 


I  M  \ 


9 


10- 

6.n-C3H70H 

7.(CH20H)2 
8.n-C4H90H 
9.H20 
IO.CH3COOH 


2k 


22 


20 


18 

co 

16 

O 
H 

X 

Ik 

W 

1 


12 


10 


8 


6 


230  240  250  260  270  280  290    230  240  250  260  270  280  290 
Wavelength  in  Mu.  Wavelength  in  Mu. 

Ultraviolet  spectra  of  III 
N2CH-C0-(CH2)  4-C0-CHN2  N2C(  CH3)  -C0-C(CH3)  3 

III  IV 

In  a  series  of  mixtures  of  acetonitrile  and  water,  the  relative  intensities  of  the 
two  peaks  shift  smoothly  with  only  a  very  slight  change  in  the  frequency  of  either  band. 
The  same  curves  are  obtained  if  water  is  added  to  the  acetonitrile  solution  or  vice 
versa,  Fig.  II.   This  family  of  curves  passes  through  an  isosbestic  point  which  consti- 
tutes proof  of  an  equilibrium  between  the  two  molecular  species  responsible  for  the 
bands( 3^->35)  •   The  fact  that  the  curves  do  not  form  a  perfect  isosbestic  point  has  been 
attributed  to  the  superposition  of  changes  in  the  refractive  index  of  the  solvent  on  the 
shift  (32). 

Figure  II 


o 

1 

0> 

H 

o 


u 

CD 

■p 

■H 
H 


O 
H 

X 


100 


230  2^0  250  260  270  280  290 
Wavelength  in  Mu. 
Spectrum  of  III  in  mixtures  of  water  and  acetonitrile.   Numbers  are  the  volume  percent 

of  water. 


-307- 

Similar  plots  were  also  obtained  by  Fahr(36)  for  several  diazoketones  and 
diazoesters  in  a  series  of  dioxane-water  mixtures. 

The  possibility  of  a  keto-enol  tautoraerism  was  considered  first.  Miller 
and  White(32)  prepared  a  series  of  diazoketones  with  substituents  at  the  positions 
a  and  ocl    to  the  carbonyl.   The  compound  (IV)  chosen  to  represent  the  case  with 
no  hydrogens  available  for  enol  formation  was,  unfortunately,,  unstable.   In  each 
of  the  four  solvents  chosen,  it  gave  a  band  at  2.90-  ^Ok  mu  which  increased  in 
intensity  with  time  and  was  attributed  to  decomposition  products.   The  major 
absorption  at  247-2^9  m(i  was  always  present }   but  there  was  no  band  at  273-275 
mu.   Those  compounds  of  the  series  having  enolizable  hydrogens  gave  both  bands 
in  the  expected  manner.   Bromine  could  not  be  used  to  test  for  a  keto-enol  taut- 
omerism  because  diazoketones  decompose  in  its  presence. 

Examination  of  the  infrared  spectra  of  these  compounds  showed  no  hydroxyl 
band  at  3200-3^00  cm  x   in  nonhydroxylic  solvents.  In  hydroxylic  solvents }   where 
the  enol  should  be  most  prevalent,  the  solvent  masked  the  region  of  interest. 
The  carbonyl  absorption  of  III  is  found  at  16^-0  cm  x   in  methylene  chloride.  It 
shifts  to  1625  cm  a  in  n-butanol,  but  it  retains  the  same  intensity.   Since  this  shift 
may  be  due  to  the  difference  in  solvent  effects }   it  does  not  prove  that  the  enol 
does  not  exist ;  but  it  argues  against  this  possibility. 

Turning  to  the  diazo  band  at  2090  cm  x  ,  they  noted  that  the  integrated  in- 
tensity was  21$  less  in  n-propanol  than  in  methylene  chloride.   This  corresponds 
to  a  20$  reduction,  observed  in  the  ultraviolet  spectra.   On  this  basis ,  they 
postulated  the  existence  of  a  diazo-isodiazo  tautoraerism^  citing  the  observation 
of  similar  forms  in  diazohydrocarbons(37) •   The  structure  Va  could  then  be  assigned 
to  the  absorption  at  2V7  mu,  while  the  peak  at  275  niu  could  be  attributed  to 
Vb  or  Vc  either  of  which  should  show  a  =N-H  band  at  3300-3^00  cm"1.  As  in  the 
case  of  the  keto-enol  tautoraerism,  the  nature  of  the  solvents  hides  the  region 
of  the  spectra  which  would  provide  proof  of  the  presence  of  the  postulated  equilibrium. 

9  $  #9  £       9 

N=N=CH~C0-CH2-R        H-N=N=C-C0-CH2-R        H-N=N=CH-C0-CH-R 

Va  Vb  Vc 

In  1959?  Fahr  (36)  reported  the  spectra  of  a  number  of  diazocarbonyl  com- 
pounds j,  both  diazo  ketones  and  diazoesters ,   which  had  no  hydrogens  available 
for  either  of  the  tautomerisms  described  above,  but  which  still  exhibited  an 
isosbestic  point  in  dioxane-water  mixtures.  He  suggested  that  a  hydrogen  bonded 
complex  was  formed  between  the  solvent  hydroxyl  and  the  carbonyl  of  the  diazo- 
carbonyl compound*   This  is  not  entirely  satisfactory,  because  it  requires  that 
the  intensity  of  the  carbonyl  remain  constant  in  spite  of  the  hydrogen  bond  for- 
mation and  also  that  the  intensity  of  the  diazo  band  be  lowered  by  reduction 
of  the  double  bond  character  of  the  carbon-nitrogen  bond. 

Foffani  and  co-workers (38}  entered  the  discussion  in  1964  with  a  study  of 
the  infrared  spectra  of  diazoacetophenone  and  its  derivatives  m  a  variety  of 
solvents.   They  also  ruled  out  the  possibility  of  a  keto-enol  tautoraerism  on 
the  basis  of  the  invariance  of  the  carbonyl  intensity.  For  diazoacetophenone _, 
the  diazo  nitrogen-nitrogen  stretching  frequency  does  not  change  (2108-2112 
cm  -M  over  the  solvent  range  investigated.   See  Table  I.   The  integrated  inten- 
sity of  this  band  is  also  nearly  independent  of  solvent.  Changing  from  apolar 
to  polar  solvents  or  from  apolar  to  commonly  hydrogen  bonded  solvents  does  cause 
the  half  width  to  increase. 

From  this  data  and  the  fact  that  some  of  their  substituted  diazoacetophenones 
had  no  hydrogens  available  but  still  gave  an  isosbestic  pointy  they  also  ruled 
out  a  diazo-isodiazo  tautoraerism. 

To  test  the  hydrogen  bonded  complex  suggested  by  Fahr^,  they  examined  the 
behavior  of  the  phenolic  hydroxyl  group  absorption  in  solutions  containing  diaz- 
oacetophenone,,  With  a  diazoketone: phenol  ratio  of  20s 1^  they  found  only  normal 
hydrogen  bonding  to  the  carbonyl.   Similar  results  were  observed  with  ratios  of 


•=308- 

Table  I 


Solvent 


-J  m 


&)i/. 


max 


X  10 


Am  x  10"4 


hexane 

CCL4 

C2CI4 

CH2C12 

CHCI3 

n-butanol 

CH3.NO2 


2108 
2108 
2108 
2109 
2111 
2112 
2108 


11 

14 

13 

18 
18 
18 
22 


1.27 
1.12 
1.15 
0.93 
O087 

0.75 
0.91 


5*4 
5.8 
5°2 


A™  is  the  integrated  intensity  in  liter /mole- 


10il  and  1:1.  In  these  cases ,  as  well  as  those  in  which  the  phenol  concentration 
would  he  too  low  to  form  appreciable  adduct,  a  second  type  of  carbonyl  was  detected. 

Having  disproved  each  of  the  three  possibilities  presented  thus  far,  they 
suggested  a  rotational  isomerism.   The  rotational  isomers'  stability  could  be 
affected  by  the  intermolecular  interactions  with  the  carbonyl.  The  spectral  data 
presented  are  consistent  with  this  view.   In  addition;,  the  low  carbonyl  frequency 
suggests  a  form  such  as  VI,  as  does  the  broadening  of  the  nitrogen-nitrogen 
stretching  band.   They  note  that  diazobxides,  in  which  rotation  is  not  possible., 
show  only  one  band  in  the  ultraviolet,  that  at  245-250  mu. 


0 


R-C 


N2 

jl 

'C-R1 


VI 


In  a  preliminary  communication  and  later  in  a  paper(39^4o) ,  Kaplan  and  Meloy 
reported  the  temperature  dependence  of  the  n.m.r.  spectra  of  diazoketones. 
Previous  investigations (41,42,43)  had  been  made  of  the  cis  and  trans  isomers 
arising  from  restricted  free  rotation  about  the  central  CVN,  0-N,  or  N-N  bond 
of  amides,  nitrites,  and  nitrosoamines.  A  similar  situation  is  possible  for 
diazoketones.  Rotation  about  the  central  C-C  bond  may  be  restricted  by  inter- 
action of  the  lone  pair  of  the  a  carbon  with  the  n   system  of  the  carbonyl. 
This  should  give  rise  to  two  isomers  of  the  diazoketone  which  may  be  designated 
cis  or  trans  from  the  geometry  of  the  jt  system. 


CIS 


trans 


The  temperature  dependence  of  the  n.m.r.  spectra  confirmed  the  existence 
of  an  equilibrium.  At  30°,  the  spectrum  of  diazoacetaldehyde  consists  of  two 
broad  singlets.  Raising  the  temperature  to  71°  causes  both  the  methine  and 
aldehyde  protons  to  exhibit  time  average  doublets  with  a  coupling  constant  of 


-309- 

2.2  ops.  At  or  below  8°,  each  region  of  the  spectrum  contains  a  singlet  and 
a  doublet  (J  =  7*5  CPS)  in  a  7«3  ratio.  In  the  methine  region,  the  singlet  is 
at  lower  field  than  the  doublet.  For  the  aldehydic  proton,  this  is  reversed. 
Other  alky.1  diazoketones,  for  example ,  diazoacetone ,  exhibit  similar  behavior. 
At  30° ,  they  nave  a  single  methine  signal  that  broadens  as  the  temperature  is 
lowered  until,  at  some  temperature,  the  methine  peak  splits  into  a  low  field 
singlet  and  a  higher  field  doublet.  The  intensity  of  the  singlet  is  usually 
about  nine  times  that  of  the  doublet.  In  the  ten  compounds  reported,  the  low 
temperature  cis: trans  ratio  varied  from  9° 1  to  1:1.  A  shift  to  lower  field  at 
lower  temperatures  was  observed  and  attributed  to  hydrogen  bonding.  The  relatively 
unhindered  cis  compounds  showed  a  greater  shift  than  that  found  for  the  trans. 

Several  aryl  diazoketones  exhibited  only  a  slight  broadening  at  the  lowest 
temperature  at  which  they  were  examined.  In  addition,  l-diazo-3j3-<liHiethyl-2- 
butanone  showed  no  broadening  over  a  70°  range.  Diazoesters  have  the  same  type 
of  temperature  dependence  as  diazoketones,  but  they  do  not  show  two  methine 
peaks  until  lower  temperatures  are  reached(-30°  to  -50°). 

The  methine  region  was  chosen  for  study  because  it  was  free  from  other 
absorptions,  showed  the  greatest  chemical  shift  difference  between  the  cis  and 
trans  form  at  low  temperatures,  and  because  it  was  a  part  of  the  constant  struc- 
tural feature  of  these  compounds.   Both  the  keto-enol  and  the  diazo-isodiazo 
tautomerisms  may  be  eliminated  from  consideration  by  the  fact  that  the  13C=-1H 
coupling  constant  (J  =  199  cps)  for  the  methine  proton  in  the  time-averaged  species 
and  that  of  the  major  species  at  low  temperatures  are  the  same  for  alkyl  diazo- 
ketones. In  the  case  of  either  tautomerism,  the  same  coupling  constant  would 
not  be  observed  for  the  time-averaged  species  because  the  exchange  of  protons 
would  average  out.   This  feature  of  the  spectra  argues  that  the  C-H  bond  is  es- 
sentially unchanged  during  the  equilibrium.  Thus  the  case  for  rotational  iso- 
merization  is  confirmed. 

Because  the  spin-spin  coupling  constant  of  trans  protons  is  expected  to  be 
much  greater  than  that  for  cis  protons  in  such  a  system,  (33*^A5)  the  low  field 
methine  singlet  (J<(  0.3  cpsT~may  be  assigned  to  the  cis  form  while  the  doublet 
is  assigned  to  the  trans  species.   Calculations  based  on  the  resonance  lines 
of  the  methine  region  were  used  to  obtain  the  relative  populations  of  the  two 
isomers,  their  mean  lifetimes(^-6)  ,  and  the  activation  energies  for  conversion 
of  the  isomers. 

Table  11(1*0) 

The  Equilibrium  Constants,  Standard  Free  Energy  Difference  Between  the 
cis  and  trans  Forms,  the  Energy  of  Activation,  and  Temperature  of  Coal- 
escence, T  ,  for  Diazoketones:  RC0CKN2. 

R        K   ( cis-Hrans)  A  F  K         T  ,  °C 

eq.  — —  a         c 

kcal/molc  kcal/mole     kcal/mole 

CH3  0.082(-40°)  1.16  15.5  ±0.9  13.9 

C2H  0.063(-40°)  1.28  16.2  +0.6  6.5 

4>CH2  0.040(-40°)  1.49  18.2  +  0.6  1.0 

CH3O  o.859(-50°)  0.07  12.5  +  0.9  -25.0 

c2ho  o.8ifO(-50°)  0.08  9.0  +  0.8  -32.5 

The  existence  of  this  cis-trans  isomerism  must  be  taken  into  account  when 
describing  the  mechanisms  of  the  reactions  of  diazoketones  since  it  may  direct 
the  course  of  the  reaction.  If  the  rate  of  interconversion  of  the  two  isomers 
is  faster  than  the  rate  of  reaction,  this  will  not  be  a  consideration. 


-310- 

The  demonstrated  preference  of  diazoketones  for  a  cis  configuration  leads 
one  to  consider  the  possibility  that  the  Wolff  rearrangement  to  ketenes  may  pro- 
ceed through  a  smooth  concerted  process  in  which  the  migrating  group  is  trans 
to  the  leaving  group.   See  Figure  III.  If  the  rearrangement  takes  place  by  in- 
itial carbene  formation,  there  -would  be  no  preference  for  geometry.  Evidence 
may  be  presented  for  both  mechanisms.   The  appearance  of  hydroxy  ketones  in. 
the  decomposition  of  some  diazoketones  in  water  would  be  more  likely  to  occur 
by  hydrolysis  of  the  carbene  than  directly  from  the  diaz ©ketone (^7) . 

Figure  III 


+ 


W= 


On  the  other  hand,  decomposition  of  VII  under  thermal,  photolytic,  or  copper- 
catalysed  conditions(l5)  leads  to  VIII  in  Q0-927> yield  with  very  little  of  the 
rearrangement  product  IX.  Models  of  the  cis  form  of  the  diazoketone  (VII)  are 
very  hard  to  make  because  of  the  large  steric  interaction  of  the  t-butyl  groups . 
Since  the  compound  exists  almost  entirely  in  the  trans  form,  the  rearrangement 
process  postulated  above  could  not  readily  take  place. 


P 


I  I 

(CH3)3C-C~d-C(CH3)3 

VTI 
((CH3)3C)2C=C=0 

IX 


R-C-CH2-R1 
XII 


Q  CH2-R' 
R-C-C-N2 
X 


0  CH3 
il  I 

(CKj)  C-C-C=C(CH3)2 


VIII 

R-CO-CH-CH- 

-R! 

XI 

For  X,   XI, 

and  XII 

R 

HL 

<t>- 

H- 

£-N024> 

CH3- 

£-CH304>- 

£-N02<t>~ 
£-Cl*- 

CH3- 
CH3- 

CH3- 

In  a  similar  manner,  X  undergoes  photolytic  and  silver  oxide  catalyzed  de- 
composition to  XI  at  room  temperature,  but  gives  appreciable  amounts  of  XII  at 
elevated  temperatures (48) .   The  cis  form  may  be  more  prevalent  if  the  additional 
thermal  energy  is  sufficient  to  overcome  the  steric  interactions. 


The  possibilities  of  both  keto-enol  and  diazo-isodiazo  taut omer isms  In  diazo- 
ketones have  been  investigated.  They  were  found  to  be  unlikely.   To  explain 
the  persistence  of  spectral  data  for  an  equilibrium  in  solutions  with  hydroxylic 
solvents,  a  hydrogen  bonded  complex  was  postulated.   This  was  also  shown  not 
to  be  the  case.   The  existence  of  a  cis-trans  isomerism  has  been  demonstrated, 
and  found  to  be  consistent  with  spectral  data. 


-311- 

BIBLIOGRAPHY 

1.  A,  Angeli,  Gazz.  Chem.  Ital. ,   24,  370  (1895). 

2.  L.  Wolff/  Ann. ,  34^,  25  (19.12). 

3.  W.  E.  Bachmann  and  W.  S.  Struve ,  "Organic  Reactions,"  Vol.  1,  John  Wiley  and 
Sons,  Inc.,  New  York,  N.  Y. ,  1942,  p  38. 

4.  R.  Huisgen,  H.  Konig,  G.  Binch,  and  H.  J.  Sturm,  Angew.  Chem.,  7]5,  368  (1961). 

5.  J.  Go  Atkinson,  M.  I.  T.  Seminars,  Spring  1959,  p  4ll. 

6.  R.  Huisgen,  Angew.  Chem.,  67,  439  (1955). 

7.  Jo  Lane  and  R.  1.  Feller',  J. '  Am.  Chem.  Soc,  73,  4230  (I95I). 

80  B.  Eistert,  "Newer  Methods  of  Preparative  Organic  Chemistry,"  Interscience 

Publishers,  Inc.,  New  York,  N.  Y. ,  1948,  p  513. 
9.  P.  Yates  and  B.  L.  Shapiro,  J.  Am.  Chem.  Soc,  8l,  212  (1959). 

10.  F.  Haupter  and  A.  Pucek,  Chem.  Ber. ,  93,  249  (1956). 

11.  E.  Wo  Robb,  Ph.D.  Thesis,  Harvard,  1956. 

12.  CD.  Nenitzescu  and  E.  Solomonica,  "Organic  Synthesis,"  Collective  Vol. 
2,  496  (1943). 

13.  0.  Diels  and  K.  Pflaumer,  Chem.  Ber.,  48,  223  (1915). 

l4o  Mo  0.  Forster  and  A.  Zimmerli,  J.  Chem.  Soc,  £7,  2156  (1910). 

15.  M.  S.  Newman  and  A.  Arkell,  J.  Org.  Chem.,  24,  385  (1959). 

16.  S.  Hauptmann,  M.  Kluge,  K.  D.  Seidig,  and  H.  Wilde,  Angew.  Chem.,  JJ_,   678 
(1965). 

17.  P.  Yates,  H.  Morrison,  and  S.  Danishefsky,  J.  Org.  Chem.,  26,  2617  (1961). 
1.8.  Ro  Bo  Wagner  and  H.  D.  Zook,  "Synthetic  Organic  Chemistry,"  John  Wiley  and 

Sons,  Inc.,  New  York,  N.  Y.  ,  1953,  P  770. 

19.  Ao  Singh,  E.  R.  Thompson,  and  F.  W.  Westheimer,  J.  Biol.  Chem.,  2J>7,  3OO6P 
(1962). 

20.  W.  R.  Bamford  and  T.  S.  Stevens,  J.  Chem.  Soc,  4735  (1952). 

21.  J.  W.  Powell  and  M.  C.  Whiting ,  Tetrahedron,  7,  305  (1959). 

22.  M.  P.  Cava  and  R.  L.  Litle,  Chem*  and  Ind. ,  (London) ,  367  (1957). 

23.  M.  Regitz,  Chem.  Ber.,  38,  1210.(1965). 

24.  M.  Regitz,  Tetrahedron  Letters,  22,  1403  (1964). 

25.  Mo  0o  Forster,  J.  Chem.  Soc,  107/260  (I915). 

26.  G.  R.  Harvey,  J.  Org.  Chem.  ,' 31,  1587  (1966). 

27.  G.  Kortum,  Z.  Chem. ,  jjO,  361  OSkl) . 

28.  A.  Jo  Ultee  Jr.  and  J.  B.  J.  Soons,  Rec  Trav.  Chim. ,  71,  565  (1952). 

29.  S.  A.  Fusari,  et.  al. ,  J.  Am.  Chem.  Soc  ,  j[6,   2878  (1954)". 

30.  So  A.  Fusari,  T.  H.  Haskell,  R.  P.  Fronardt,  and  Q.  R.  Bartz ,  J.  Am.  Chem. 
Soc,  7iL  2Q®1   (1954). 

51.  A.  Foffani,  C.  Pecile,  and  S.  Ghersetti,  Tetrahedron,  11,  285  (i960). 

32.  F.  A.  Miller  and  W.  Bo  White,  J.  Am.  Chem.  Soc,  79,  5974  (1957). 

33«   J«  R«  Dyer,  "Applications  of  Absorption  Spectroscopy  of  Organic  Compounds," 

Prentice -Hall,  Inc ,  Englewood  Cliffs,  N.  J.,  1965,  p  l4,  (b)  p  115. 
34.  Ho  Jaffe  and  M.  Orchin,  "Theory  and  Applications  of  Ultraviolet  Spectroscopy," 

John  Wiley  and  Sons,  Inc.,  New  York,  N.  Y. ,  1962,  p  562. 
35«  W.  West,  "Techniques  of  Organic  Chemistry,"  A.  Weissberger,  Ed.,  Interscience 

Publishers,  Inc,  New  York,  N.  Y. ,  1949,  Vol.  1,  Part  II,  p  1299 . 

36.  E.  Fahr,  Chem.  Ber.,  %2,   398  (.1959). 

37.  E.  Muller  and  D.  Ludsteck,  Chem.  Ber.,  87,  1887  (1954). 

38.  Ao  Foffani,  C.  Pecile,  and  S.  Ghersetti,  Tetrahedron,  20,  823  (1964). 

39.  F.  Kaplan  and  Go  K.  Meloy,  Tetrahedron  Letters,  2427,  (1964). 

40.  F.  Kaplan  and  G.  K.  Meloy,  J.  Am.  Chem.  Soc,  88,  95O  (1966). 

41.  M.  T.  Rogers  and  J.  C.  Woodbry,  J.  Phys.  Chem.,  66,   540  (1962). 

42.  P.  Graves  and  1.  W.  Reeves,  J.  Chem.  Phys.,  ^2,  1B78  (i960). 

43.  G.  J.  Karabatsos  and  R.  A.  Taller,  J.  Am.  Chem.  Soc,  86,  4373  (1964). 

44.  A.  J.  R.  Bourn,  D.  G.  Gilles,  and  E.  W.  Randall,  Tetrahedron,  20,  l8ll  (1964). 

45.  L.  H.  Piette,  J.  D.  Ray,  and  R.  A.  Ogg,  J.  Mol.  Specry. ,  2,  66T1958) . 

46.  F.  A.  Bovey,  E.  W.  Anderson,  F.  P.  Hood,  and  R.  L.  Kornegay,  J.  Chem.  Phys. ,  40, 
3099  (1964). 


-312- 

tyj.     E.  S.  Gould,,  "Mechanism  and  Structure  in  Organic  Chemistry/'  Henry  Holt  and 

Company,  New  York,  N.  Y. ,   1959,  P  628. 
48.  V.  Franzen,  Ann.,  602,  199  (1957). 


J 


^1 3- 


E.S.R.  STUDIES  OF  ORGANIC  GROUND-STATE  TRIPLET1  MOLECULES 


Reported  by  Robert  J.  Basalay 
INTRODUCTION 


April  17,  1967 


The  following  discussion  of  triplet-state  molecules  will  be  confined  to  molecules 
which  exist  as  triplets  in  their  ground-states  or  are  in  equilibrium  with  ground -state 
singlet  species o  The  use  of  esr  spectroscopy  to  observe  triplet-states,  and  the 
results  of  its  use  to  observe  the  triplet -states  of  diradicals,  methylene  derivatives, 
and  jt -electron  triplets  will  be  discussed  .  A  diradical  is  properly  defined  as  a 
molecule  with  two  unpaired  electrons  whose  centers  of  gravity  do  not  coincide,  but 
here  the  term  is  used  to  indicate  molecules  with  two  unpaired  electrons  localized 
on  different  atoms  which  may  or  may  not  interact  to  give  a  triplet -state 0 

THE  TRIPLET  STATE1'2 


numbers j  m 


The  two  spin  states  of  an  electron  can  be  represented  by  the  spin  quantum 
_l/2„  If  more  than  one  electron  is  involved,  the  spin  states  are 
represented  by  the  total  spin  quantum  number,  S,  given  by  S- |2(ms) ^1  .  If  two 

electrons  have  their  spins  parallel  in  a  given  state,  S-j (ms) electron  1  +  (ms) 
electron  2Ul„  The  degeneracy  of  a  state  due  to  electron  spin  is  given  by  2S  +  1. 
The  S-l  state  is  triply  degenerate  and  a  triplet  state „   If  the  electrons  are  anti- 
parallel,  S-0  and  it  is  a  singlet  state. 

Two  factors  determine  whether  the  ground  state  of  a  molecule  is  a  singlet  or  a 
triplet ,  the  orbital  energies  and  the  strength  of  the  exchange  interactions  between 
electrons  of  parallel  spln0  If  we  have  two  degenerate  orbitals,  the  triplet  state 
is  always  lower  in  energy  because  of  the  exchange  interactions^  but  if  the  orbitals 
are  of  different  energy  the  two  electrons  may  be  found  paired  in  the  orbital  of 
lowest  energy,  to  form  a  singlet  state .  A  competition  between  the  energy  separation 
of  the  orbitals  and  the  energy  gained  by  having  parallel  spins  determines  whether  a 
molecule  exists  as  a  ground -state  triplet  or  singlet „  If  the  two  factors  are 
comparable,  a  thermal  equilibrium  will  exist  with  comparable  amounts  of  singlet  and 
triplet . 

The  degenerate  triplet  state  levels  can  be  split  by  an  external  magnetic  field 
(the  Zeeman  effect).  The  triplet  levels  have  the  spin  quantum  numbers  m  =1,0,-1 
associated  with  them.  Mien  the  levels  are  split,  there  are  two  possible  esr 
transitions,  Ama+1  and  Zma<+2o  However,  the  selection  rule  for  esr  transitions  is 
Ams+i0  Hence,  the  Zsns+2  transition  is  forbidden  and  only  the  Am=+1  transition  would 
be  observed o  The  resulting  spectra  would  be  similar  to  that  of  a  radical  with 
S<L/2.  However,  only  exchange  and  electrostatic  interactions  between  electrons 
have  been  considered. 

If  the  magnetic  dipole-dipole 
forces  between  the  two  unpaired 
electrons  are  considered,  the 
degeneracy  of  the  triplet  is 
removed  without  the  application  of 
an  external  magnetic  field  (zero- 
field)  o  Mien  the  dipole  interaction 
is  considered  and  zero  field  splitting 
of  the  triplet  levels  occurs,  the 
"forbidden"  z^n=+2  transitions  becomes 
allowed o  The  appearance  of  this 
transition  is  not  a  violation  of  the  usual  selection  rule,  £ms4-l,  because  it  does  not 
apply  at  the  low  magnetic  field  strength  (~1500  gauss,  hV^OO  Mc./s)  where  the  £m=+2 
transition  occurs.  Hence,  there  are  two  types  of  transitions  in  the  esr  spectra  of 
triplet -states,  !,Am=+l"  and  "&e  \Z 

A  spin  Hamiltonian  for  the  two  major  magnetic  interactions,  the  electronic 
Zeeman^and  the  dipolar,  of  two  unpaired  electrons  within  a  molecule  is  X  s  g£H«S  + 
g2^ f  Si'S^/r3  -  3(S1»r)(S2'r)/r5J  where  r  is  the  vector  joining  the  two  electrons, 
P  is  the  electronic  Bohr  magneton.  The  isotropic  electron  g  factor  should  be  an 


m. 


s 

1 

0 

-1 


** — ~ 

anisotropic  g  tensor,  but  for  most  organic  triplets  the  g  tensor  anisotropy  is 
small  and  will  be  neglected  here. 

The  second  term  of  X  is  the  dipolar  interaction  term  )i  p.  Expanding  Xj)  by  means 
of  the  various  vector  products  and  expressing  everything  in  terms  of  the  total  spin  S, 
we  obtain  XD  ■  l/2g^2  (s|(r2-3x2)/r5  +  S§( r2-3y2)  /r5  +  S2(  r2-3y2)  /r5  -  (SxSy  +  SySx)  ' 
3xy/r5  -  (SySz  +  SzSy)3yz/r5  -  (SXSZ  +  SZ5X) 3xz/r5J  „  This  result  can  be  expressed 
in  the  matrix  form:; 


'S\  /(r2-5x2)/rs  -3xy/r5  3xz/r5  \  ( SSS) 


l/2g  . 


3xy/r5 
=3xz/r" 


(r2-3y^r5  -3yz/r5 
3yz/r5  (fr2-3z^r5, 


S-D-S 


D is  a  symmetric  tensor  called  the  zero-field  splitting  tensor  which  may  be  diag= 
onalized  by  proper  choice  of  coordinate  systems.   In  terms  of  the  new  coordinate 
system  (x%y 


z») 


X=(r2-x^22/r5,  Y=(r2-y,2)/r5,  and  Z-(r2-z'2)/r5,  and  the  spin 


Hamiltonian  becomes  X  =  gpH-S-XSj-YS^-ZS2.  Since  the  tensor  is  traceless  (X+Y+Z-O), 
the  zero-field  splitting  can  be  expressed  in  terms  of  just  two  independent  constants, 
D=l/2(X+Y)-Z  and  E=-l/2(X-Y)  ,  giving  X  s  D(S2-l/3S2)  +  E(sf-S2)  +  gBH'St 

In  order  to  find  the  energy  level  splitting  in  the  triplet  state  due  to  zero- 
field  splitting j  the  eigenvalue  problem,  W<§  =<8>fc  ,  must  be  solved.  The  spin 
functions  will  be  chosen  to  diagonalize  the  zero-field  Hamiltonian  matrix  where 


(V»o 


ff^Z* 


f 


1 


field,  H, 


X-Ei 

igBHz 
-igpH, 


-igpHz 

Y~E2 

igpHx 


The  solution  of  the  eigenvalue  problem  is  given  by  the  matrix,  T  o 

If  there  is  no  external  field 
( Hx=Hy=Hz=0) ,  the  matrix  is  diagonalized 
n      and  E1=X>  E£=Y,  and  E3=Z,  the  energies 
of  the  levels  of  the  triplet  at  zero 
magnetic  field. 

If  we  impress  an  external  magnetic 
on  our  system  parallel  to  the  z  direction,  H-Hz  and  Hx=Hy=0,  and  the 
matrix  has  the  following  solutions;  Ej=l/2(X+Y)  +  l/2(X-Y)  tan  0+  g0H,  Eg=l/2(X+Y)- 
l/2(X-Y)  tan  0  -  gpH,  and  Ea=Z,  where  tan  2Q  =  (X-Y)/2gPH.  A  set  of  graphs 
simulating  the  zero-field  splitting  when  H=HZ,  Hy,  and  Hx  is  given  in  Figure  l. 

Figure  1  Zero-Field  Splitting  of  Triplet  Energy  Levels 


Magnetic  Field  Strength  (H)- 


-315- 
The  arrows  in  Figure  1  represent  electron  spin  resonance  transitions  at  constant  hv 
while  varying  H. 

The  esr  spectra  of  a  triplet  whose  molecules  were  all  oriented  the  same  way 
would  have  three  transitions  whose  position  would  vary  as  the  orientation  of  the 
triplet  with  respect  to  the  external  magnetic  field  were  changed .  An  example  of 
this  case  will  be  seen  when  the  esr  spectrum  of  diphenylmethylene  is  discussed. 
Mien  the  triplet  molecule  is  dissolved  in  a  liquid,  the  rapid  molecular  tumbling 
averages  out  the  effect  of  dipole-dipole  interaction  on  the  esr  spectra  of  randomly 
oriented  molecules.  Mien  the  triplet  molecule  is  frozen  in  a  glass,  the  effect  of 
the  dipolar  interaction  is  not  averaged  out  for  the  randomly  oriented  molecules.  The 
intensity  of  the  esr  signal  is  greatest  when  the  external  magnetic  field  is  parallel 
to  a  principal  magnetic  axis  of  the  triplet.  The  derivative  curve  of  an  esr 
absorption  spectrum  gives  special  prominence  to  the  magnetic  field  strengths  which 
cause  esr  transitions  to  occur  in  the  molecules  which  have  principle  magnetic  axes 
parallel  to  the  external  magnetic  field „  This  type  of  triplet  esr  spectrum  is 
simulated  in  Figure  2.  Approximate  zero-field  splitting  parameters  can  be  obtained 
from  studies  of  the  Am=±2  transitions  whose  anisotropy  is  relatively  small,  allowing 
them  to  be  easily  observed  for  randomly  oriented  triplet  molecules  in  frozen  matrices.4 

Figure  2.   Simulation  of  ESR  Spectra  of  Randomly  Oriented  Triplets3 


Absorbtion 
Spectrum 


E=0 


E^O 


ST-J  H 


A   A 


A. 


NV 


^H 


Derivative 
Spectrum 


DIRADICALS 

Certain  organic  compounds  exhibit  a  chemical  behavior  characteristic  of  free 
radicals,  but  contain  even  numbers  of  electrons.  One  of  these  compounds  is 
Chichibabin's  hydrocarbon,  p,p» -biphenylene-bis-(diphenylmethyl)  (I).  This  compound 


Triplet 
(paramagnetic) 


-  AS-O-O-fo 


Singlet 
(diamagnetic) 


\  // 


was  thought  to  exist  in  two  forms,  a  triplet  state  and  a  singlet  state.  The  fact 
that  an  esr  signal  is  observed  for  Chichibabin's  radical  in  solution  indicates  a 
paramagnetic  species  is  present.5 

Another  type  of  diradical  (II),  where  X  is  a  bridging  group  (i.e.  -0-,  (-CH2-)n,)> 

can  not  have  a  quinoid     singlet  state.   If  there  is 
X  yy  Ciflo     spin  interaction  between  the  two  halves  of  the  diradical, 
vZy  there  will  be  a  ground -state  triplet  and  an  excited - 

state  singlet.   If  there  is  no  spin  interaction,  the 
halves  of  the  molecule  will  be  independent  of  one  another  and  behave  like  two  mono- 
radicals.  Compounds  of  this  type  (II)  were  examined  by  Jarrett,  Sloan,  and  Vaughan.6'7 
The  sharpness  (esr  linewidth  <-~  12  gauss)  of  the  esr  spectra  indicated  a  comparatively 
small  interaction  between  the  unpaired  electrons.  The  larger  the  spin  interaction, 
the  broader  the  esr  spectrum  becomes  because  the  spin-spin  coupling  produces  strong 
relaxation  effects  which  decrease  the  relaxation  time  increasing  the  esr  line-width. 


-316- 

Ill  IV  V 

The  spin  exchange  ( spin  interaction  between  the  two  radical  centers)  in  several 
biradicals  (I,  III,  IV,  and  V)  which  were  labeled  with  C13  at  the  triphenylmethyl 
carbon  atoms  was  studied  by  means  of  the  resultant  nuclear  hyperfine  splitting 
observed  in  the  esr  spectra.  If  the  spin  interaction  is  large,  the  esr  spectra  will 
show  nuclear  hyperfine  splitting  due  to  two  equivalent  C13  (1=1/2)  atoms.  This  is  a 
triplet -state.  If  the  spin  interaction  is  small,  the  esr  spectra  will  show  nuclear 
hyperfine  splitting  due  to  one  C13  atom.  This  is  equivalent  to  a  pair  of  doublets 
or  a  triplet  with  very  weak  spin  interaction.  The  latter  case  was  observed. 

However,  Weissman8  has  shown  that  the  electron  interaction  between  the  aromatic 
rings  of  the  anions  of  paracyclopheaes  (VI),  provided  n  or  m  is  1  or  2,  is  large  (i.e. 
CqH4—  (CH2)m        "tne  nuclear  hyperfine  splitting  due  the  hydrogen  atoms  of 
X  .    X         both  aromatic  rings  are  observed  in  the  esr  spectra) .  If  n, 
(CHaJn-CsH*      m  >  3,  then  the  nuclear  hyperfine  splitting  due  to  the 

VI  hydrogen  atoms  in  one  aromatic  is  observed  in  the  esr  spectra 

indicating  small  electron  interaction  between  the  aromatic 
rings.  MeConnell9  has  performed  theoretical  calculations  for  the  ground -excited 
state  splittings  in  large  biradicals  of  this  type  (II)  and  according  to  the  calcu- 
lations a  larger  spin  interaction  is  indicated.  The  spin  interaction  observed  for 
a  -CHsCH^-  bridge  (i.e.  X  in  II)  is  equivalent  to  a  calculated  spin  interaction 
resulting  from  a  bridge  of  about  five  -CH2-  groups.  This  discrepancy  between  the 
spin  interaction  observed  and  that  one  would  expect  is  the  "biradical  paradox." 

A  possible  explanation  for  the  smaller  than  expected  spin  exchange  was  offered 
by  Bersohn. 10  One  or  two  -CH2-  groups  do  not  present  a  serious  barrier  to  the 
passage  of  a  single  electron,  but  it  is  a  formidable  barrier  to  the  spin  interaction 
of  two  electrons  of  parallel  spin. 


1  CI 


VII  VIII 

An  examination11  of  the  esr  spectra  of  compounds  I,  VII,  VIII,  and  DC  indicated 
that  the  paramagnetic  species  in  biradical  solutions  is  a  dimer  or  higher  polymer. 
As  a  result  the  spin  interaction  between  electrons  would  be  much  smaller  than  would 
be  expected  for  the  monomer.  It  was  found  that  the  room  temperature  intensity  of 
the  esr  signal  can  be  maintained  at  low  temperature  if  the  specimen  is  cooled  rapidly 
enough.  If  the  temperature  is  lowered  slowly  the  intensity  of  the  esr  signal  decreases, 
Also,  when  the  specimen  is  heated,  the  esr  signal  is  enhanced.  If  the  heated  solu- 
tion is  cooled  to  room  temperature,  the  esr  signal  diminishes  to  a  final  intensity  of 
slightly  less  than  the  starting  intensity.  If  the  specimen  is  cooled  to  a  temperature 
at  which  equilibrium  is  slow,  the  intensity  of  the  esr  signal  can  be  enhanced  by 
illumination.  The  enhancement  persists  until  the  solution  is  heated  to  a  point  where 
the  equilibrium  is  rapid.  Thereupon  the  intensity  diminishes  again.  These  effects 
indicate  that  biradical  monomers  associate  in  solution  to  form  dimers  or  polymers. 

In  two  communications,  Chandross  and  Kreilick12  reported  attempts  to  produce 
a  triplet  molecule  by  linking  two  phenoxyl  radicals  by  a  -CMe2-  bridge  (XII).  No 
triplet  spectra  were  observed,  but  chemical  and  spectral  data  indicated  the  reaction 
scheme  on  the  following  page.  The  infrared  spectrum  of  the  oxidation  product  (XIII) 
shoved  no  band  at  6,h   u,  characteristic  absorption  of  phenoxyl  radicals,  but  there 
was  a  doublet,  characteristic  of  this  type  (XIII)  of  cyclohexadienone ,  at  6.08  and 
6,16  u.  The  presence  of  the  biradical  (XII)  in  equilibrium  with  XIII  is  indicated 
by  a  rapid  reaction  of  the  oxidation  product  with  02,  The  dienone  (XHI)  would  not 
be  expected  to  react  rapidly  with  oxygen  (reaction  complete  in  minutes  at  room  temp.), 
but  the  biradical  (XII)  would. 


Fb02£ 
Ag20, 
MnOo 


To  make  the  internal  formation  more  difficult,  they  then  replaced  the  -CMs2- 
group  with  a  2,2' -biphenylene  group  (XIV).  When  XIV  was  oxidized  with  aqueous 
alkaline  f erricyanide ,  the  infrared  spectrum  of  the  oxidation  product  in  CCI4  had 
a  pair  of  strong  absorptions  at  6.0  and  6.2  u  assigned  to  XVI  and  a  strong  band  at 
6.4  [i   which  was  assigned  to  XV.  The  esr  spectrum  revealed  the  presence  of  triplet 
species.  There  were  two  sets  of  /^m=il  lines  as  well  as  a  Am=±2  transition.  The 
larger  splitting  was  attributed  to  XV  and  the  smaller  was  thought  to  be  due  to  a 
dimer  possibly  linked  by  a  peroxide  bond  (XVII) .  If  this  smaller  splitting  is  due 
to  the  dimer  (XVII) ,  it  would  indicate  that  the  unpaired  electrons  can  interact 
relatively  strongly  across  long  organic  molecules,  possibly  to  the  degree  which  could 
be  predicted  by  theory  without  creating  the  "biradical  paradox." 


-* 


XIV 


XVI 


0-0 


XVII 


A  pyridinyl  diradical  has  been  prepared  and  examined  by  Kosower  and  Ikegami 
Their  esr  studies  of  1,1"  ethylene -bis -(4-carbomethoxypyridinyl)  ( XVIII)  indicates 
radical-radical  association  producing  dimers  or  polymers.  The  ends  of  the  n-mers 
behave  like  monoradicals  as  in  the  case  of  Chichibabin's  hydrocarbon.  The  esr 
spectrum  of  XVIII  at, 77°K  at  low  concentration  is"  characteristic  of  a  triplet  with 


13 


CH302C^^n^-CH2-CH2-1<^>-C02CH3   ^ffi  On30^-^\.Q^^R2 


I 


-N^)-C02CH3 
XVIII 


D=0.0178  cm""1  and  E=0.0017  cm"1.  With  increasing  concentration,  the  triplet  spectrum 
disappears  as  a  strong  signal  due  to  the  monoradical-like  ends  of  the  polymers  appears, 
Again,  it  would  be  interesting  to  see  whether  the  spin  interaction  here  could  be 
predicted  without  creating  a  "biradical  paradox. " 

Young  and  Castro14  prepared  a  stable  biradical  (XXI)  which  Kreilick15  studied  by 
esr.  The  esr  spectrum  at  room  temperature  of  the  oxidation  product  of  XIX  initially 
gave  an  esr  signal  with  hyperfine  splitting  due  to  four  equivalent  protons,  from  the 
formation  of  XX.  On  further  reaction,  an  esr  signal  appeared  with  splitting  due  to 
six  equivalent  hydrogen  atoms,  from  the  formation  of  XXVII.  Apparently,  the  unpaired 
electrons  can  delocalize  into  the  three  aromatic  rings  to  interact  with  all  six  meta 
protons.  The  biradical  (XXVII)  esr  spectrum  consists  of  two  pairs  of  lines  disposed 
about  g=2.  The  separation  of  the  inner  more  intense  lines  is  38. k   gauss. 


-318- 

Since  D«:  l/r3  where  r  is  the  average  distance  between  the  two  unpaired  electrons 
and  the  proportionality  constant  is  known  from  theory,1  a  separation  of  9  A  between 
the  two  unpaired  electrons  is  calculated.  The  degree  of  spin  interaction  here  is 
less  than  observed  for  XV . 

OH  0'  0* 


PbO: 


MTHF 


&W 


XXX 
METHYLENE  DERIVATIVES 


XX 


XXI 


The  first  methylene  compound  to  be  studied  by  esr  was  diphenylmethylene „  Murray, 
Trozzolo,  Wasserman,  and  Yager16*17  irradiated  a  solution  of  diphenyldiazomethane  in 
"Flororolube "  (polyehlorotrifluoroethylene)  at  77°K  with  an  Hg  arc.  The  esr  spectrum 
consisted  of  derivative  signals  at  1227,  1619,  ^588,  5272,  and  7522  gauss.  The 
intensity  of  the  spectrum  did  not  decrease  as  long  as  the  temperature  was  kept  below 
123°K,  indicating  the  paramagnetic  species  was  the  ground -state  or  was  at  least  in 
thermal  equilibrium  with  the  singlet-state.  The  five  line  spectrum  corresponded  to 
a  triplet  esr  spectrum  of  randomly  oriented  molecules  having  a  spin  Hamiltonian  with 
D  =  0.401  cm"1  and  E  =  0.018  cm"1. 

Brandon,  et  al.18  subsequently  presented  the  results  of  a  detailed  study  of  the 
esr  spectra  of  ground -state  triplet  diphenylene  molecules  oriented  in  benzophenone 
single  crystals.  A  preliminary  examination  of  the  esr  spectra  revealed  that  there 
were  four  differently  oriented  sets  of  principle  axes  indicating  four  equivalent  but 
differently  oriented  positions  in  the  benzophenone  unit  cell  could  be  occupied  by 
the  triplet  species.  The  major  magnetic  axes  of  the  triplets  in  one  of  these 
equivalent  sites  were  determined.  If  the  magnetic  field  were  made  parallel  to  one 
of  these  axes  and  the  crystal  was  rotated  about  this  axis,  the  esr  line  positions  at 
several  fixed  angles  of  rotation  were  measured.  The  variation  of  the  magnetic  field 


Figure  3.  Variation  of  Triplet  Line 
Position  with  Rotation  about  a  Magnetic 
Axis.   (The  points  are  determined  from 
separate  esr  spectra  and  curves  are 
drawn  to  fit  them.) 


H( magnetic  fie! 


strengths  necessary  for  resonance  with  respect  to  the  angle  of  rotation  is  simulated 
in  Fig.  3"  The  angles  of  rotation  where  maximum  separations  occur  indicate  the 
positions  of  the  other  two  axes.  The  principle  magnetic  axes  were  determined  relative 
to  the  crystallographic  axes  of  the  benzophenone  single  crystal.  The  measurements 
of  the  stationary  magnetic  field  strengths  necessary  for  resonance  versus  the  angle 
of  the  external  magnetic  field  in  the  plane  of  the  principle  magnetic  axes  were  fitted 
by  the  spin  Hamiltonian  X  -  I?«g»(0)S  +  D(Sf  -J./3S2)  +  E(S|  -_S|)  „  Those  values  of 
D  and  E  which  gave  the  best  fit  are  0.4050  cm"1  and  0.0192  cm"1,  respectively. 

The  zero-field  splitting  parameter  (D)  is  very  large  compared  to  diradicals 
(D^O.Ol  cm"1)  and  jt -electron  triplets  (D"O.20  cm"1).   Since  Dgel/r3,  the  large  value 
of  D  indicates  the  electrons  in  the  diphenylmethylene  are  on  the  same  carbon  atom  a 
large  fraction  of  the  time.  Higuchi19  has  calculated  the  value  of  D  for  CH2# 
methylene,  where  the  two  unpaired  electrons  are  each  in  one  of  two  orthogonal  p  orbitals, 


-319- 

to  be  0,90,55  cm"1.  The  D  value  of  diphenylene  is  smaller  than  Dqjj  indicating  the 
unpaired  electrons  can  delocalize  into  the  phenyl  rings,   Skell2°  has  proposed  a 
linear  structure  for  diphenylmethylene  with  D^  symmetry,  with  each  orthogonal  p 
orbital  conjugated  with  a  phenyl  ring.  If  the  triplet  were  linear  the  zero-field 
splitting  parameter  E  would  be  zero 3,  but  this  is  not  observed.  Hence,  the  diphenyl- 
methylene molecule  is  bent.  The  structure  of  diphenylmethylene  probably  has  one 
electron  generally  localized  at  the  divalent  carbon  and  the  other  conjugated  with  the 
phenyl  rings.  This  bent  structure  for  methylene  compounds  has  been  confirmed  by  the 
observation  of  geometric  isomers  of  1-  and  2-naphthylmethylenes  by  esr.21  The  esr 
spectra  indicated  the  presence  of  two  sets  of  triplet  species  whose  zero-field 
splitting  parameters  are  similar  indicating  similar  species  and  the  difference 
between  the  zero-field  splitting  parameters  of  each  isomer  does  not  vary  as  the  host 
(the  frozen  matrix)  is  changed, 


o 


XXIV 


XXV 


XXVI 


The  zero  field  splitting  parameters  for  many  methylene  compounds  have  been 
determined  .17i'18''21"26  The  D  (O.5I8  cm"1)  value  of  phenylmethylene  is  larger  than 
the  D  of  diphenylmethylene  as  expected  since  two  phenyl  groups  would  allow  greater 
derealization  of  electrons,  lowering  D.  The  compounds  XXII -XXVII  have  D  values 
(0.37-0.42  cm"1)  similar  to  that  of  diphenylmethylene  indicating  a  similar  electronic 
structure  (i0e.  one  electron  localized  at  the  divalent  carbon  atom  and  one  delocalized) 
Propargylene  derivatives26  have  relatively  large  values  of  D  (0. 55-0. 63  cm"1,  but  one 
would  expect  that  the  unpaired  electrons  could  be  delocalized  into  the  triple  bond. 
These  compounds  also  have  E  —  0  indicating  a  linear  geometry  at  the  methylene  group. 
If  the  unpaired  electrons  delocalize^,  they  would  go  into  two  perpendicular  it   systems 
(see  Fig.  V» .  Then,  each  carbon  atom  of  the  conjugated  chain  would  have  some  fraction 
of  triplet  methylene  character  according  to  the  spin  density  of  the  jt-system  at  the 
carbon  atom.  Thus,  the  effect  of  derealization  of  the  unpaired  electrons  to  lower 
D  is  nulified.  Theoretical  calculation  indicate  that  negative  spin  densities  must 
be  used  to  predict  the  experimental  value  of  D. 


H — 


1/  TPfy 

c— c — c— 

1<tnlll  ,  / 1 


-R 


delpcaliz at  i  on 


Liwii|yi.i»|y 
H — C  —  C  —  C— 


•R 


Figure  ka     De localization  of  Unpaired  Electrons  in  Propargylene  (H-C-CeCH)  Derivatives 
jt-ELECTRON  TRIPLETS 

The  possibility  that  certain  derivatives  of  cyclic  k-n   jt-electron  systems  could 
have  triplet  ground  states  is  suggested  by  simple  molecular  orbital  theory.  This  is 
based  on  the  degeneracy  of  the  highest  occupied  molecular  orbitals.  These  orbitals 
are  degenerate  only  if  the  system  is  symmetrical.  If  the  symmetrical  system  is 
distorted  ( Jahn-Teller  effect) ,  the  degeneracy  of  the  orbitals  is  lost.  If  two 
electron  are  placed  in  the  degenerate  orbitals,  a  ground -state  triplet  results,  but 
if  the  orbitals  are  not  degenerate  both  electrons  may  occupy  the  lowest  orbital 
giving  a  singlet. 

Cyclic  aromatic  compounds  with  three  fold  or  greater  axes  of  rotation,  have  two 
degenerate  unfilled  lowest  anti-bonding  it-electronic  levels.   If  dinegative  anions 
were  made  from  the  addition  of  two  electrons  to  anti-bonding  orbitals,  they  could  be 
triplet  states.  Jahn-Teller  effects  could  remove  the  degeneracy  to  allow  the  two 
electrons  to  occupy  one  orbital  and  form  a  singlet  ground-state.  The  simplest  case, 
benzene  dinegative  ion,  has  not  been  reported.  However,  the  esr  spectra  of  the 
dinegative  ions  of  triphenylbenzene27  (XXIX)  ,  aecacylene^l'XXVTII)  ,  and  coronene28 
(XXX)  have  been  reported.  The  coronene  dinegative  ion  did  not  give  a  very  intense 
esr  spectrum  in  solution,  but  if  the  di-potassium  coronene  salt  is  crystallized  from 
tetrahydrofuran,  the  solid  obtained  gives  a  triplet  esr  spectrum  whose  signal 


■320- 


XXVIII 


Y 

XXIX 


0 


XXX 


intensity  varies  intensively  with  temperature.  At  -l60°C  no  triplet  signal  was 
observed.   Increase  in  temperature  resulted  in  a  continuous  increase  of  esr  triplet 
signal  intensity  up  to  +60°C  where  the  sample  decomposed.  This  temperature  dependence 
indicates  a  thermally  excited  triplet -state.  The  zero-field  splitting  parameters  (D) 
for  (XXVIII),  (XXIX),  and  (XXX)  were  respectively  0.021  cm"1,  0.0^2  cm"1,  and  0.^2. 
0rO6O  cm"1.   Since  D*cl/r  ,  the  magnitude  of  D  should  vary  inversely  with  the  size 
of  the  jt-electronic  level  the  unpaired  electrons  occupy,  and  it  does. 

Cyclopentadienyl  cations,  which  may  have  triplet  ground-states  according  to 
simple  molecular  orbital  theory,  have  been  studied  by  esr.29  Penta-P-naphthyl 


Energy^-- 


1L 


ii 


Figure  5.  Energy  Level 
Diagrams  for  Symmetrical  and 
Distorted  Cyclopentadienyl 
Cation 


symmetrical 


distorted 


cyclopentadienyl  cation  did  not  show  a  triplet  esr  spectrum  at  any  temperature.  The 
pentaphenyl  and  pentachloro  derivatives  gave  very  distinct  triplet  esr  spectra.   If 
the  molecule  has  a  triplet  ground -state  and  the  excited  singlet  state  is  not  appre- 
ciably populated,  or  the  singlet  and  triplet  states  are  equal  in  energy,  the 
intensity  of  the  esr  spectrum  should  follow  Curies  law  (i.e.  intensity  is  inversely 
proportional  to  temperature).   If  the  triplet  is  not  the  ground -state,  the  intensity 
of  the  esr  signal  due  to  the  triplet  still  obeys  Curie's  law,  but  the  concentration 
of  the  triplet  varies  with  temperature  being  a  thermally  excited  state,  and  thus 
Curie's  law  is  not  obeyed.  The  esr  spectrum  of  the  pentaphenylcyclopentadienyl 
cation  did  not  obey  Curie's  law  and  the  triplet  is  a  thermally  excited  state 9   but 
the  esr  spectrum  of  the  pentachlorocyclopentadienyl  cation  did  obey  Curie's  law  and 
the  triplet  is  the  ground -state  or  very  close  to  it.  The  symmetry  of  the  cyclo- 
pentadienyl ring  would  be  expected  to  be  distorted  more  the  larger  its  substituents 
become.  The  more  distorted  the  ring  is  the  more  likely  the  ground -state  will  be  a 
singlet,  and  it  is  observed  as  the  substituents  (P-naphthyl>  phenyl>  chloro)  on  the 
cyclopentadienyl  ring  become  more  bulky  the  triplet  is  less  stable.  The  zero-field 
splitting  parameters  for  the  pentachloro  and  pentaphenyl  derivatives  were  D  =  0.1^95 
cm"1  and  D  -  O.IO5O  cm"1,  respectively,  as  would  be  expected  since  the  pentaphenyl 
derivative  would  allow  the  two  unpaired  electrons  to  have  a  greater  average  distance 
between  them  than  in  the  pentachloro  derivative. 

CONCLUSION 

Many  examples  of  triplet-state  species  have  been  presented  which  can  be 
observed  by  esr  spectroscopy.  The  zero-field  splitting  parameters  can  be  used  to 
infer  information  about  the  electronic  structure  of  the  triplet  species.  By  use  of 
esr  to  study  reaction  mixtures,  it  should  be  possible  in  some  case  to  detect  and 
characterize  triplet  reaction  intermediates. 

BIBLIOGRAPHY 

1.  A.  Carrington  and  A.  D.  McLachlan,  "Introduction  to  Magnetic  Resonance," 
Harper  and  Row,  New  York,  N.Y. ,  1967. 


-321- 

2.  K.  G.  Harbison,  M.I.T.  Seminar  in  Organic  Chemistry,  Fall,  1963. 

3.  E0  Wasserman,  L.  C»  Snyder,  and  W.  A.  Yager,  J.  Chem.  Phys.,  4l,  1763  (1964) . 

4.  M.  S.  deGroot  and  J.  H.  van  der  Waals,  Mol.  Phys. ,  2,  333  (1959)  »   Mol.  Phys., 
3,  190  (1960)o 

5„  Co  A.  Hutchison,  Jr.,  A.  Kowalsky,  R.  C0  Pastor,  and  G„  W.  Wheland,  J.  Chem. 
Phys.,  20,  148.5  (1952). 

6.  H„  S.  Jarrett,  G.  J.  Sloan,  and  W.  R.  Vaughan,  Jc  Chem.  Phys.,  25_,  697  (1956)  . 

7.  G.  Jo  Sloan  and  W.  R„  Vaughan,  J0  Chem.  Phys.,  22,  750  (1957). 

8.  S.  I,  Weissman,  J.  Am„  Chem.  Soc,  80,  6462  (195B). 

9.  H.  M.  McConnell,  J.  Chem.  Phys.,  27,  115,  968  (1957). 

10.  R.  Bersohn,  Ann.  Rev.  Phys.  Chem.,  11,  382  (i960)  . 

11.  R.  K.  Waring,  Jr.  and  G.  J.  Sloan,  J.  Chem.  Phys.,  40,  772  (1964). 

12.  E.  A.  Chandross  and  R.  Kreilick,  J.  Am.  Chem.  Soc.,~B£,  2530  (1963) j  J.  Am. 
Chem.  Soc,  86,  117  (1964). 

13.  E.  M.  Kosower  and  Y.  Ikegami,  J.  Am.  Chem.  Soc,  89,  46l  (1967). 

14.  N.  C.  Yang  and  A.  J.  Castro,  J.  Am.  Chem.  Soc,  82,  6208  (I960). 

15.  R.  Kreilick,  J.  Chem.  Phys.,  4^,  308  (1965). 

16.  R.  W.  Murray,  A.  M.  Trozzolo,  E.  Wasserman,  and  W„  A.  Yager,  J.  Am.  Chem.  Soc. 
84,  3213  (I962). 

17.  A.  M.  Trozzolo,  E.  Wasserman,  and  R„  W.  Murray,  J.  Am.  Chem.  Soc,  84,  4990 
(1962). 

18.  R.  W.  Brandon,  G.  L.  Closs,  C.  E„  Davoust,  C.  A.  Hutchison,  Jr.,  B.  E.  Kohler, 
and  R.  Silbey,  J.  Chem.  Phys.,  4^,  2006  (1965) . 

19.  J.  Higuchi,  J.  Chem.  Phys.,  ^8,  1237  (I963). 

20.  R.  M.  Etter,  H„  S.  Skovronek,  and  P.  S.  Skell,  J.  Am.  Chem.  Soc,  &L,  1008 

(1959). 

21.  A.  M.  Trozzolo,  E.  Wasserman,  and  W.  A.  Yager,  J.  Am.  Chem.  Soc,  87,  129 

(1965). 

22.  E.  Wasserman,  L.  Barash,  A.  M.  Trozzolo,  R.  W.  Murray,  and  W.  A.  Yager,  J.  Am. 
Chem.  Soc,  86,  2304  (1964). 

23.  I.  Moritani,  S.  I.  Murahashi,  M.  Nishino,  Y.  Yamamoto,  K.  Itah,  and  N.  Mataga, 
J.  Am.  Chem.  Soc,  89^  1259  (1967). 

24.  E.  Wasserman,  J.  Chem.  Phys.,  42,  3739  (1965). 

25.  E.  Wasserman,  L.  Barash,  and  W„  A.  Yager,  J.  Am.  Chem.  Soc,  87,  4974  (1965). 

26.  R„  A.  Bernheim,  R.  J.  Kempf,  J.  V.  Gramas,  and  P.  S.  Skell,  J.  Chem.  Phys., 
i£,  196  (1965). 

27.  R„  E.  Jesse,  P.  Biioen^,  R.  Prins,  J.  D.  W.  van  Voorst,  and  G.  J.  Hoijtink, 
Mol.  Phys.,  6,  633  (I963). 

28.  M.  Glasbeek,  J.  D.  W.  van  Voorst,  and  G.  J.  Hoijtink,  J.  Chem.  Phys.,  45,  I852 
(1966). 

29.  R.  Breslow,  H.  W.  Chang,  R.  Hill,  and  E.  Wasserman,  J.  Am.  Chem.  Soc,  89, 
1112  (I967). 


-322= 

RECENT  STUDIES  CONCERNING  THE  MECHANISM 
OF  THE  FAVORSKII  REARRANGEMENT 

Reported  by  Peter  A.  Gebauer  April  20,  1967 

I.   Introduction 

The  Favorskii  rearrangement  of  a-haloketones  to  derivatives  of  carboxylic 
acids  by  the  action  of  bases  has  been  reviewed. 1>2*3'4  A  general  review  of  reactions 
of  a-haloketones  with  bases  has  also  appeared.5  The  purpose  of  this  seminar 
will  be  to  present  the  recent  work  related  to  the  mechanism  of  the  Favorskii 
rearrangement,  especially  that  concerning  the  question  of  whether  a  cyclopro- 
panone  intermediate,  If  formed }   is  generated  via  an  intermediate  zwitterion  or 
delocalized  species,,  or  whether  it  is  formed  in  a  concerted  manner. 

II o   Historical 

The  labeling  experiments  by  Loftfield  which  indicated  that  the  a  and  a' 
carbon  atoms  of  2-chlorocyclohexanone  become  equivalent  during  the  course  of  the 
Favorskii  rearrangement  limit  the  possibilities  for  the  mechanism  of  the  Favorskii 
rearrangement.   Although  one  case  will  be  discussed  in  which  the  a  and  a'  carbons 
do  not  become  equivalent }   it  appears  that  generally  the  reaction  must  proceed 
through  either  a  cyclopropanone  (i),  or  through  a  planar  species  (II)7-*8  which  is 
probably  best  described  as  a  diradicalj,  but  in  keeping  with  the  usage  of  most 
of  the  workers  in  this  field,  this  mechanism  will  be  referred  to  as  the  zwitter- 
ionic  mechanism.   Usually  the  intermediate  (II)  is  considered  to  form  a  cyclo- 
propanone in  the  course  of  the  rearrangement.8 


0  0 

II  II 

_c/CxcV~~*-c/C 

"1        I '  I 

I  II 

Burr  and  Dewar8  argued  against  the  concerted  formation  of  the  cyclopropa- 
none on  the  grounds  that  the  geometry  of  the  p  orbital  on  the  carbanion  is  not 
correct  for  nucleophilic  displacement  of  the  halogen. 

III.  Reactions  of  Cyclopropanone s 

Breslow  and  co-workers  have  reported  the  interception  of  a  cyclopropanone 
under  conditions  similar  to  those  of  the  Favorskii  rearrangement.   Cycloprope nones 
such  as  IV  were  obtained  by  treatment  of  the  dibromides  III,  Vs  and  VI  with  tri- 
ethylamine  in  methylene  chloride  or  chloroform. 

RCHBr-CO-CHBrR  -Et3N  j 
III  R=<!> 
V   R=n-C4H9 


Turro  and  Hammond10  found  that  cyclopropanones  will  react  to  form  products 
analogous  to  those  of  the  Favorskii  rearrangement.  Treatment  of  2,4-dimet-hyl- 
2-bromo=3-pentanone  ( VTl)  with  sodium  methoxide  in  ether  produced  only  12$  of 
the  rearrangement  product  VIII  while  tetramethylcyclopropanone,  IX,  the  expected 
intermediate  in  the  Favorskii  rearrangement  of  VTl  and  the  hemiketal  of  tet- 
ramethylcyclopropanone (X)  produced  97$  and  )>98$  of  VIII  respectively.   The  hemi- 
ketal (X)  in  refluxing  methanol  also  produced  24$  of  VIII.  In  all  cases,  XI 
was  the  only  other  product  reported. 


-323- 


0 


IX 


CH3 

0  OMe 
(CH3)2CH-C-C(CH3)2 

(CH3)2CH-C-C02Me 
1 

CH3   VIII 

XI 

^a0Me  >              973$ 
Me  OH  or  DME             yu 

3$ 

Me<X  /OH 


>9&$  <1# 


X 


MeOH(refl.) 
> 


2k% 


(CH3)2CH-CO-CBr(CH3)2   ^gg  )  12$ 

VII 

Similar  results  were  obtained  with  2,2-dimethyIcyclopropanone.11 
Turro  and  Hammond10  concluded  that  VII  does  not  form  either  the  cyclopro- 
panone  or  the  hemiketal  as  the  major  product  when  treated  with  "base.   They 
also  conclude  that  the  cyclopropanone  is  in  equilibrium  with  an  acyclic  species 
such  as  II.   This  is  not  necessarily  required  since  it  is  possible  that  methanol 
might  add  directly  to  the  cyclopropanone  as  suggested  by  House.12 

It  would  seem  from  a  comparison  of  the  results  of  the  reactions  of  the 
bromide  in  base  and  the  hemiketal  in  refluxing  methanol  that  cyclopropanones 
might  form  ketonic  solvolysis  products  such  as  XI  if  the  concentration  of  base 
were  low  enough.  However,  other  means  of  formation  of  these  products  must,  in 
general,  be  operative  also. 

IV.   Tb.e  Question  of  a  Delocalized  Intermediate 

Fort13  suggested  that  the  intermediate  (II)  proposed  by  Aston  and  Newkirk7 
and  Burr  and  Dewar8  might  not  close  to  a  cyclopropanone  but  react  directly  with 
solvent  to  form  ketonic  solvolysis  products  or  with  base  to  form  rearrangement 
products.  He  suggested  that  XII  might  be  the  most  efficient  way  of  describing 
the  intermediate. 


^  £K  -*- 

L  I  ~  I  I  I  I 

XII 


As  indicated  in  XII,  Fort  meant  to  imply  that  there  was  some  overlap  between 
the  radial,  carbons  and  oxygen  but  he  conceded14  that  this  would  be  slight  due 
to  the  large  distances  between  the  radial  atoms. 

Fort13  found  that  a-chlorodibenzyl  ketone  (XIII)  reacted  with  lutidene  in 
methanol  to  produce  only  a-methoxydibenzyl  ketone  (XEV).   Desyl  chloride  (XV)  and 
a-chloroacetone  (XVI)  did  not  react  under  these  conditions. 


-324- 


-CH- 


WCH3        ?Me 


45CH2-CO-CHCl<t> 

Me 

OH  * 

$CH2-C0-CH4> 

XIII 

XIV 

<t>-C0-CHCl4> 

CH3-CO-CH2Cl 

XV 

XVI 

It  was  assumed  that  the  delocalized  intermediate  vas  formed  in  the  first  case 
hut  not  in  the  last  hecause  the  base  is  not  strong  enough  to  form  it  in  the 
absence  of  stabilization  by  phenyl  groups. 

Fort14  also  treated  compound  XVII  with  varying  concentrations  of  base. 


NaOMe 


MeOH 


C02CH3 


C02CH3 


+ 


+ 


XVII 


XVIII 


XIX 


He  found  that  as  the  concentration  of  base  decreased,  the  yields  of  rearrangement 
products  also  decreased.,  and  concluded  that  a  delocalized  intermediate }   which 
can  react  with  base  to  form  rearranged  products  (XIX)  or  with  solvent  to  form 
ketonic  solvolysis  products  ( XVIII.)  was  being  formed. 

Fort's13'14  conclusions  seem  to  rest  on  the  tacit  assumption  that  a  cyc- 
lopropanone would  not  give  ketonic  solvolysis  products  in  the  absence  of  strong 
base.  However,  as  noted  earlier,  the  work  of  Turro  and  Hammond11'12  seems  to 
indicate  that  all  the  reactions  of  the  delocalized  intermediate  could  as  easily 
be  explained  by  a  eye lopropanone. 

Fort,15  Richey  and  co-workers,16  and  Cookson  and  co-workers17  have  trapped 
adducts  which  they  all  assumed  were  formed  by  the  addition  of  an  acyclic  species , 
such  as  XEI,  to  furan.  Hammond  and  Turro11  have  also  obtained  a  furan  adduct  of 
2 ,2-dimethylcyc lopropanone.   Although  these  workers18  suggest  that  the  cyclopro- 
panone  may  be  in  equilibrium  with  some  acyclic  species,  it  might  be  possible  for 
the  cyclopropanone  itself  to  form  this  adduct. 

V.   Solvent  Effects  and  Stereospecificity 

The  conflict  between  Loftfield's6  concerted  mechanism  and  the  stepwise 
zwitterion  mechanism7'8  cannot  be  resolved  by  kinetic  determinations,  but  the 
stereochemistry  of  the  products  is  different  for  the  two  mechanisms.1  Where 
two  stereoisomeric  products  are  possible ,  the  concerted  mechanism  predicts  in- 
version of  configuration  at  the  a   carbon  bearing  the  halogen  while  the  zwitterion 
mechanism  should  proceed  with  racemization. 

Stork  and  Borowitz19  found  that  the  cis  and  trans  isomers  of  1-chloro-l- 
acetyl-2-methylcyclohexane  (XX)  underwent  Favorskil  rearrangement  with  nearly 
complete  inversion  and  concluded  that  the  cyclopropanone  formation  was  concerted. 


House  and  Gilmore20  found,  however,  that  the  stereospecificity  of  the  reaction 
seemed  to  depend  on  the  polarity  of  the  solvent.   The  cis  isomer  of  XX  produced 
essentially  racemized  product  when  the  rearrangement  was  carried  out  in  methanol, 


-325- 

and  mainly  inverted  product  in  dimethoxyethane  (DME).   It  was  suggested  that  this 
solvent  effect  was  due  to  polar  solvents  facilitating  the  loss  of  chloride  to 
form  the  planar  intermediate  (II)   while  the  concerted  mechanism  predominates 
in  nonpolar  solvents. 

Similar  results  were  obtained21  using  piperitone  oxide  XXI  and  isophorone 
oxide  XXII  as  models  for  a-haloketones. 

0 


XXI  XXII 

Tchoubar  and  co-workers22'23  found  a  similar  solvent  effect  with  compound  XX, 
shown  in  Table  I. 

T,ule  I 

E-Vect  of  Solvent  on  Stereospecificity  of  Favorskii  Rearrangement 

Base         Solvent  fo   Inversion  jo  Retention 

NaOMe        DMSO  75-5  2^.5 

(")*        (DME)  (95)*  (5) 

t~BuOH  83.5  1.6.5 

KOH  H20-Pyridine  37  63 

^Values  in  parenthesis  are  from  House20  for  purpose  of  comparison. 

Tchoubar22  argued  that  solvation  of  the  anion  was  the  factor  which  deter- 
mined the  solvent  dependence.   They  assumed  that  in  a  nonpolar  solvent  the  negative 
charge  of  the  anion  (XXIIl)  resides  on  the  a!  carbon  which,  they  say,  would  be 
tetrahedral.   In  a  polar,  hydrogen  bonding  solvent,  the  solvent  would  be  pre- 
sumed to  hydrogen  bond  with  the  oxygen  and  localize  the  negative  charge  on  the 
oxygen  atom.   In  this  case  they  expected  the  a1  carbon  to  be  hybridized  sp2  and 
planar. 

0 
II 

VC\  ci 


-fa' 


ay\ 


XXIII 

They  therefore  concluded  that  the  critical  difference  between  the  suggestions 

of  Loftfield6  and  Burr  and  Dewar8  is  whether  the  a'  carbon  is  tetrahedral  or  planar. 

It  is  unlikely  that  a  tetrahedral  carbanion  would  have  a  very  long  life- 
time when  the  possibility  of  derealization  of  the  charge  exists.  However,  the 
carbanion  might  react  rapidly  before  rehybridization.   This  would  require  that 
the  anion  displace  the  chloride  either  very  soon  after,  or  simultaneously  with,  loss 
of  the  proton.   This  possibility  can  be  checked  by  labeling  studies.   If  the 
reprotonation  is  faster  than  or  as  fast  as  the  loss  of  chloride,  the  anion  prob- 
ably has  a  lifetime  long  enough  to  form  the  more  stable  planar  species. 

House  and  Thompson24  found  that  essentially  no  deuterium  was  incorporated 
into  unreacted  9-chloro-trans  -I-decalone  when  treated  with  sodium  methoxi.de  in 
deuterated  methanol. 

Olson25  found  that  no  deuterium  was  lost  from  the  C-k   position  in  unreacted 
2-a-bromo-  anS  2-a-iodocholestan-3-one-2,4-d3  (XXIV)  ,  but  that  the  corresponding 
chloro  compound  lost  20/6  of  the  label  after  one  half  life  of  the  Favorskii  rearrangement 


-326- 
when  treated  with  0.2  M  sodium  ethoxide  in  absolute  ethanol. 


XXIV 


X=C1,  Br,  I 


Olson  concluded  that  in  the  bromo-  and  iodo-  compounds  loss  of  a  proton 
is  rate  determining  while  in  the  chloro  compound,  loss  of  chloride  occurs  at 
approximately  the  same  rate  as  reprotonation.  This  conclusion  was  verified  by 
the  observation  that  the  trideuterated  compounds  above  lost  halide  at  one-fifth 
the  rate  of  their  hydrogen  analogues. 

Mayer26  and  Tchoubar  and  co-workers27  studied  the  Favorskii  rearrangement 
of  1-acetyl-l-chlorocyclohexane  (XXV),  using  sodium  phenoxide  as  the  base  in 
the  presence  of  deuterated  phenol. 


:och3 


XXV 


They  found  that  about  half  a  mole  of  deuterium  was  incorporated  per  mole  of 
starting  material  which  was  recovered  and  concluded  that  the  first  step,  at 
least  in  this  case,  of  the  Favorskii  rearrangement  is  a  pre-equilibrium  between 
the  anion  and  starting  material.   This  conclusion  is  not  necessarily  valid  since 
it  is  possible  that  the  reaction  which  is  responsible  for  the  deuterium  in- 
corporation may  not  be  related  in  any  way  to  the  Favorskii  rearrangement. 

House  and  Gilmore21  also  found  deuterium  incorporation  in  unreacted 
isophorone  oxide  and  piperitone  oxide. 

It  appears  that  the  rate  of  reprotonation  of  the  anion  is  quite  close  to 
that  of  loss  of  halide.   The  anion  discussed  earlier  therefore  probably  has  a 
fairly  long  lifetime  and  is  planar  with  the  negative  charge  delocalized  on  to 
both  the  carbon  and  oxygen  atoms. 

House  and  Thompson24  suggested  that  an  equatorial  halogen  would  favor  the 
concerted  formation  of  a  cyclopropanone  while  an  axial  halogen  should  favor 
formation  of  the  zwitterion  or  delocalized  intermediate.  However,  work  by 
House  and  Frank12 ■,24  using  both  the  cis  and  trans  isomers  of  9-chloro-l-decaIone 
(XXVl)  led  to  the  conclusion  that  the  ketonic  solvolysis  products  did  not  arise 
from  a  planar  intermediate  in  either  case  since  the  cis  and  trans  isomers  pro- 
duced different  product  ratios.   Since  the  same  planar  intermediate  should  arise 
from  both  isomers,  the  products  also  should  have  been  the  same.  Ketonic  sol- 
volysis products  were  obtained  from  both  isomers  in  methanol  and  rearrangement 
products  were  obtained  in  DME.   The  possibility  of  direct  nucleophilic  displace- 
ment was  ruled  out  since  the  9-methoxy  compounds  were  formed  mainly  with  retention 
rather  than  inversion. 

The  trans  isomer  produces  mainly  the  products  (XXVII)  which  would  be  formed 
by  the  formation  of  the  most  stable  carbanion,  while  the  cis  isomer  produces 
mainly  the  unexpected  products  (XXVIIl). 


-327- 


C02Me 


C02Me 


H   C02Me 


XXVII 


H 
b 


H 
a 


XXVIII 


COpMe 


The  stereochemistry  at  the  9-Posi'tion  in  "the  9-carbomethoxy  derivatives 
(XXVTl)  was  consistent  with  a  concerted  formation  ox   the  cyclopropanone.   That 
is,  the  cis  isomer  formed  XXVIIa  and  the  trans  isomer  formed  XXVIIb.  The  ke- 
tonic  solvolysis  products  were  suggested  to  be  formed  via  either  an  alkylidine 
epoxide  (XXIX)  or  the  hemiketal  of  the  cyclopropanone  (XXX). 


CH 


3-- 


XXIX 


XXX 


JH 


House  concluded  that  the  conformation  of  the  halogen  did  not  affect  the 
mechanism  of  the  Fa-vorskii  rearrangement. 

Smissman  and  co-workers28  also  have  studied  this  effect.   They  subjected 
>a-hromo-trans-2-decaIone  ( XXXI)  ^  >e-bromo-tranS"2-decalone  (XXXIIa)  and 
2-e-bromo-9-methyl-trans-3-decalone  (XXXIlb)  to  Pavorskii  rearrangement  con- 
ditions in  polar  (EtpH)  and  nonpolar  (DME)  solvents.  The  axial  coumpouni  ( XXXI) 


R 

H 

XXXLI 

a  R=H 

b  P.=CH3 

in  both  solvents  produced  only  products  which  were  presumed  to  arise  from  the 
epoxide  (.XXXIII).   No  rearrangement  products  were  obtained. 


■-CEt 


XXXIII 


The  non-methylated  equatorial  isomer  ( XXXIIa)  formed  both  the  products  found 
in  the  axial  case  and  some  rearrangement  product  in  both  solvents.   The  methyl- 
ated equatorial  isomer  ( XXXIlb)  produced  38  to  k-hfo  rearrangement  product  (XXXIV) 
a  diacetic  acid  (XXXV)  and  9-methyl-trans~decaiin-2 , 3-dione  (XXXVl) .   This 
last  compound  was  presumed  to  be  formed  by  oxidation  and  to  be  the  precursor  of 
the  diacetic  acid. 


-328- 


COpMe 


C02H 
C02H 


CH3 
XXXV 

Rowland29  suggests  that  products  such  as  these  arise  from  air  oxidation  of  in- 
itially formed  products,  hut  it  is  unlikely  that  this  is  occurring  in  this  case. 

Smissman  concluded  that,  in  the  case  of  the  compounds  which  were  studied, 
the  cyclopropanone  mechanism  is  operative,  and  the  conformation  of  the  halogen 
rather  than  the  polarity  of  the  solvent  determined  whether  rearrangement  will 
occur  or  not.   He  explains  the  solvent  effects  observed  hy  House  and  Gilmore21 
by  saying  that  the  a,P-epoxyketones  cannot  reach  a  true  equatorial  position  and 
so  this  probably  effects  the  route  of  the  reaction.   He  harks  back  to  Wendler's30 
suggestion  that  halogen  migration  might  occur  before  rearrangement  in  order 
to  rationalize  the  solvent  dependence  in  the  case  of  l-acetyl-l-chloro-2-methyl- 
cyclohexane.20  Kende1  claims  that  halomethyl  compounds  such  as  XXXVTII  react 
slower  than  those  having  the  halogen  on  a  carbon  in  the  ring  as  in  XXXVII. 


CHpX 


XXXVII 


xxxvrii 


If  this  is  true,  then  the  chlorine  cannot  be  migrating  before  reaction.   On  the 
other  hand  House20  says  that  XXXVIII  qualitatively  reacts  faster  thai  XXXVII, 
in  which  case  no  conclusion  may  be  drawn. 
VI.   Rearrangement  of  2-Bromocyclobutanone 

The  Favorskii  rearrangement  of  2-bror.iocyclobutanone  (XXXIX)  is  known 
to  occur  under  conditions,  such  as  carbonate  as  the  base  or  simply  in  water 
solution,31  which  are  much  milder  than  the  usual  Favorskii  conditions.   This 
fact,  coupled  with  the  fact  that  cyclopentanones  generally  do  not  undergo  Favorskii 
rearrangement,  seems  to  suggest  that  the  mechanism  which  is  operative  in  this 
case  is  not  the  general  one  for  the  Favorskii  rearrangement. 


</ 


TBr 


CO- 


H20 


(>-co£ 


Two  possible  mechanisms  were  considered  for  this  rearrangement: 

A.  Semibenzilic 

B.  Cyclopropanone 

The  predicted  labeling  results  for  these  two  mechanisms  when  run  in  D20  are 
shown  in  Fig.  1. 


C02D 


COoD 


B 


The  labeling  studies32  indicate  that  the  semibenzilic  mechanism  is  operative 
since  no  deuterium  is  incorporated  into  the  ring.   It  has  previously  been  shown1'6 
that  this  mechanism  is  not  the  general  one  for  the  Favorskii  rearrangement 
since  this  mechanism  is  unsymmetrical.   However,  the  rearrangement  of  XXXIX 
does  indicate  that,  although  the  cyclopropanone  mechanism  is  generally  the  pre- 
dominating one,  other  mechanisms  are  operative  in  cases  where  the  formation 
of  a  cyclopropanone  is  unfavorable  or  where  the  other  mechanisms  are  particu- 
larly favored o 

Conclusion 

The  Favorskii  rearrangement  appears  usually  to  go  through  a  cyclopropanone 
intermediate,  although  other  intermediates  cannot  be  strictly  ruled  out.   Other 
points  concerning  this  mechanism  such  as  solvent  effects  and  the  relationship 
of  the  conformation  of  the  halogen  to  the  course  of  the  reaction,  appear  to 
depend  to  a  great  extent  on  the  system  under  consideration. 


Bibliography 

1.  A.  S.  Kende,  Organic  Reactions,  11,  261(1960). 

2.  J.  G.  Shell,  M.  I.  T.  Seminars' in  Organic  Chemistry,  Dec,  14,  I960. 

3.  R.  Jacquier,  Bull.  soc.  chim,  France,  17,  D35  (1950). 

4.  D.  W.  Lamson,  Univ.  of  Illinois  Seminars  in  Organic  Chemistry,  May  21, 
196U. 

5»   B.  Tchoubar,  Bull.  soc.  chim.  France,  1363  (1955). 

6.  R.  B.  Loftfield,  J.  Am.  Chem.  Soc,  73,  4707  (1951). 

7.  J.  Aston  and  J.  Newkirk,  J.  Am.  Chem.  Soc,  73,  39OO  (1951). 

8.  J.  G.  Burr  and  M.  J.  S.  Dewar,  J.  Chem.  Soc,  1201  (1954). 

9.  Ro  Breslow,  J.  Posner,  and  A.  Krebs,  J.  Am.  Chem.  Soc,  85,  234  (1963). 

10.  N.  J.  Turro  and  W.  B.  Hammond,  J.  Am.  Chem.  Soc,  87,  3258  (1965). 

11.  W.  B.  Hammond  and  N.  J.  Turro,  J.  Am.  Chem.  Soc,  88,  2880  (1966). 

12.  H.  0.  House  and  G.  A.  Frank,  J.  Org.  Chem.,  30,  29E8  ( 1965) . 

13.  A.  W.  Fort,  J,  Am.  Chem,  Soc,  84,  2620  (1962)". 

14.  A.  W.  Fort,  J.  Am.  Chem.  Soc,  "cW,  2625  (I962). 

15.  A,  V.   Fort,  J,  Am.  Chem.  Soc,  BE,  1+979  (1962). 

16.  H.  G.  Richey,  Jr.,  J.  M.  Richey,  and  D.  C.  Clagett,  J.  Am.  Chem.  Soc, 

86,  3906  ( 1964) . 

17.  Ro  C.  Cookson,  M.  J.  Nye,  and  G.  Subrahmanyan,  Proc  Chem.  Soc,  144  (1964) 

18.  N.  J.  Turro,  W,  B.  Hammond,  and  P.  A.  Leermakers,  J.  Am.  Chem.  Soc, 

87,  277^  (1965). 

19.  G.  Stork  and  I.  J.  Borowitz,  J.  Am.  Chem.  Soc   82,  4307  (1962). 
20. 


H,  0.  House  and  W.  F.  Gilmore,  J.  Am.  Chem.  Soc, 


.,  3980  (1961) 


-330- 

21.  H.  0.  House  and  W.  F.   Gilmore,  J.  Am.   Chem.  Soc,  83,  3972  (1961). 

22.  B.  Tchoubar,  Bull.  soc.  chim.  France,  1533  (1963). 

23.  A.  Gaudemer,  J.  Parcello,  A.  Skrobek,  and  B.  Tchoubar,  Bull.  soc.  chim. 
France,  2^05  (1963) . 

2k.      H.  0.  House  and  H.  W.  Thompson,  J.  Org.  Chem. ,  2_8,  l6k   (1963). 
23.   B.  A.  P.  Olson,  Dissertation  Abstracts,  2jj,  MkL3  (1965). 

26.  M.  Mayer,  Compt.  Rend.,  233,  ^88  (1961). 

27.  M.  Charpentier-Morize,  M.  Mayer,  and  B.  Tchoubar,  Bull.  soc.  chim.  France, 
529  (1963). 

28.  E.  E.  Smissman,  J,  L.  Lemke,  and  0.  Kristiansen,  J.  Am.  Chem.  Soc,  88, 
335  (1966), 

29.  A.  J.  Rowland,  J.  Org.  Chem.,  27,  1135  (1962). 

30.  N.  L.  Wendler,  R.  P.  Graber,  and  G.  G.  Hazen,  Tetrahedron,  £,   ikk   (1958). 

31.  Jo-M  Conia  and  J. -M.  Ripoll,  Bull.  soc.  chim.  France,  755  (19&3) • 

32.  J. -Mo  Conia  and  J.  Salaun,  Tetr.  Letters,  1175  (1963). 


PROSTAGLANDIN  SYNTHESES 


Reported  by  Edward  Bertram 


April  2k,   I967 


Prostaglandins  are  C20  carboxylic  acids  with  hormone -like  qualities.  They 
show  varying  pharmacological,  effects1  such  as  smooth  muscle  activation,  lowering  or 
raising  of  blood  pressure,  mobilization  of  lipids  and  affecting  the  reproductive 
tract o  They  are  found  in  almost  all  parts  of  the  body,  but  mainly  in  the  vesicular 
glands  of  man  and  animals .  This  seminar  will  deal  with  their  recent  laboratory 
syntheses  and  the  proposed  mechanisms  for  biosynthesis . 

The  initial  structure  determination  was  carried  out  by  Samuelsson  and  Berg- 
strom.2'3  By  the  utilization  of  mass  spectra,  infrared,  ultraviolet,  optical 
rotation  and  chemical  degradation,  they  proposed  the  mirror  image  of  I  as  the 
absolute  stereochemistry  of  prostaglandin  Ex  (PGEx) „  This  structure  was  basically 
confirmed  by  Abrahamsson4  by  an  X  ray  analysis  of  the  tris-p-bromobenzoate  methyl 
ester  of  prostaglandin  F^a  (PGFxp)  Ho 

0  H,  rm  0 


( -) -ll-a-15(s) -dihydroxy-9-°xo-13 

trans  prostenoic  acid 


Since  then  Samuelsson,  Van  Dorp  and  co-workers5  have  re -evaluated  the  evidence 
and  the  determination  of  the  optical  rotations  of  some  by-products  and  degradation 
products,  and  compared  them  to  other  alcohols  of  known  absolute  stereochemistry  and 
decided  that  the  absolute  stereochemistry  for  PGEx  is  as  shown  in  I0 

Other  prostaglandins  are  PGE2  which  contains  an  additional  cjLs  double  bond  at 

C5  and  PGE3  which  contains  two  additional  cis  double  bonds  at  C5  and  Cx?.  The  PGFq, 

and  the  FGFg  ser5.es  are  the  prostaglandins  obtained  by  reduction  of  the  C9  oxo  group, 

3  has  the  hydroxy!  trans  to  the  other  ring  hydroxy!  and  a   contains  the  two  hydroxyl 

groups  ciso   Other  series  are  the  PGE-217  III  and  PGE-278  IV.  These  are  named  after 

their  ultraviolet  spectra  and  are  obtained  as  natural  products  or  by  dehydration  of 

the  Cn  hydroxy!  group.  The  C19  hydroxyl.  compounds  of  some  of  these  a ,6 -unsaturated 

derivatives  have  also  been  isolated.6 
OH 


The  prostaglandins  are  derived  from  essential  fatty  acids,'  PGEx  from  all  cis 


8,  11,  lk   eicosatrienoic  acid  V 


8, 


11,  lk   eicosatetraenoic 


,  PGE2  from  all  c_is 
8,  11,  lkf   17  eiccsapentaenoic  acid  VII 
determined  by  ^  or  14C  labeling  studies.8 


acid  VI  and  PGE3  from  all  cis 


This  was 


0 

VI 


OH 


VII 


770 
-JJ1-' 


It  was  also  shown  that  the  hydrogens  at  0Qf   Cu  and  C12  were  not  lost  in  the 
formation  of  PGEX  using  vesicular  sheep  enzyme.  Kleriberg  and  Samuelsson9  proved 
this  by  specifically  labeling  the  three  positions  with  ^  using  a  14C  labeled  8, 
11,  Ik   eicosatrienoic  acido 

Samuelsson10'5'11  and  later  VanDorp12  showed  conclusively  that  the  oxygen  on  the 
ring  and  the  C15  hydroxy!  group  came  from  molecular  oxygen.  They  also  proved  that 
both  the  ring  oxygens  came  from  the  same  molecule  of  oxygen,  Samuelsson  incubated 
all  cis  8,  11,  lk   eicosatrienoic  acid  in  an  oxygen  atmosphere  of  180~180  5670, 
160-T50  1%   and  160-160  hjS   and  then  reduced  the  VGEX   to  the  PGFxp  compound  with 
NaBH^  in  EtOH,  thus  preventing  the  exchange  of  the  keto  oxygen  on  workup.  He  then 
methylated  the  hydroxyl  groups  and  cleaved  the  double  bond  using  KMh04  and  periodate 
to  obtain  the  diacid  which  he  then  ethylated.  The  mass  spectrum,  of  this  compound 
was  run  and  the  P  +  2  and  the  P  +  k   peaks  were  compared  for  ions  which  still  con- 
tained both  oxygens.   It  was  determined  that  if  both  oxygens  in  the  ring  resulted 
from  the  same  oxygen  molecule  then  the  ratio  P  +  2/P  +  k   would  be  0.02,  but  if  two 
oxygen  molecules  were  involved  then  the  ratio  would  be  I.5«  He  obtained  a  ratio  of 
0.06  and  so  concluded  that  both  ring  oxygens  must  be  from  the  same  molecule.  Van- 
Dorp12  used  an  oxygen  mixture  of  180-18Q  kj$>,   160-180  6i   and  160-160  h 7$  and  a  much 
cleaner  enzyme  preparation  (hardly  any  endogenous  ¥GEX   present)  and  treated  the 
sample  in  three  different  ways,  (A)  kept  all  original  oxygens  (B)  exchanged  the  C9 
oxo  group  for  all  160  and  (C)  removed  the  oxygen  at  C1S  by  oxidation.  He  obtained 
the  following  table  of  values. 


jo   of  molecules  with18© 


no 


18/ 


(A) 
(B) 

(C) 


k-6% 


one  180 
25. 5# 

Wo 

6i 


two 


18 


0 


three  180 


25.5% 


hQi 


These  experiments  show  that  all  of  the  oxygen  functions  are  obtained  from 
molecular  oxygen  and  that  both  ring  oxygens  are  from  the  same  molecule.  Van  Dorp 
also  measured  the  up -take  of  oxygen  and  estimated  that  two  moles  of  oxygen  were 
consumed  per  mole  of  PGEx  formed.  He  also  determined  that  Glutathione  was  a  co- 
factor  (almost  exclusively)  and  that  there  was  a  need  for  an  anti -oxidant  (propyl 
gallate  or  £-hydroquinone)  to  obtain  maximum  yields  of  BGrElo 

The  Van  Dorp  group  also  identified  several  by-products  obtained  by  varying  the 
reagents  of  incubation.  In  a  normal  reaction  using  Glutathione  and  an  anti-oxidant, 
they  obtained  PGEx  I  and  compound  VIII  ( ll-hydroxy-8°trans -10°trans -l^-cis - 
eicosatrienoic  acid).  If  the  antioxidant  was  left  out,  I,  VIII,  IX  (il-a-hydroxy- 
9,15-clioxo~13prosteneic  acid)  and  X  (15=hydroperoxy-ll^»hydroxy-9-oxo-13-prostenoic 
acid)  resulted.  If  Glutathione  was  not  added,  there  was  a  significant  drop  in  PGEX 
formation  and  compounds  XI  ( 12~hyaroxy~8-trans -IQ-trans-heptadecadienoic  acid),  XII 
(9-=Q!"15"dihydroxy-ll-oxo-I3"prostenoic  acid)  and  XIII  PGF^  were  isolated. 

From  the  labeling  experiments  Samuelsson  proposed  two  possible  mechanisms,  one 
(Scheme  l)  was  the  direct  molecular  oxygen  addition  across  the  Cu,  Ce  position  and 
a  conrotatory  Cs,  C1S  bond  formation.  This  would  be  done  by  a  dioxygenase,  a  sub- 
sequent monooxygenase  for  the  addition  of  a  second  oxygen  at  C15  followed  by 
isomerization  of  the  double  bond  to  the  C13  trans  configuration  and,  then  opening  of 
the  peroxide. 

9 


dicxygena.se 


COgH 


monooxygenase 


^^GO^H 


-7   "2-7 


A   second  mechanism  (Scheme  2)  he  proposed  was  the  formation  of  a  hydroperoxide 
at  Cu>  isomerization  of  the  double  bond  from  A11  to  A12,  then  attack  on  the  Cg  by 
the  hydroperoxide  to  form  a  cyclic  peroxide  followed  by  a  concerted  ring  closure  to 


form  the  C&$  0^2  bond  with  isomerization  of  the  A   bond  to  trans  A  ,  then  the 
addition  of  a  hydroxyl  group  at  C15  and  opening  of  the  peroxide  to  give  PGE1(,  With 
only  the  use  of  the  labeling  experiment 9   Samuelsson  could  not  determine  if  initial 
oxygenation  was  at  Cll9   Cg,  or  C15o 


^^^CO^L 


-i     I 


Scheme  2 


Van  Dorp's  proposed  mechanism  (Scheme  3)  is  similar  to  Scheme  2  except  that  he 
uses  a  peroxide  radical  and  supports  the  C^x   attack  with  the  fact  that  compound  VIII 
is  formed,  but  no  compounds  have  been  isolated  with  only  the  C15  hydroxyl  group  or 
Cg  hydroxyl  group,,  He  supports  the  radical  mechanism  because  of  an  esr  band  which 
forms  when  the  enzyme  and  substrate  are  added  and  slowly  disappears  on  incubation 0 
He  also  shows  that  compound  XI  is  formed  from  an  intermediate  in  the  PGEX  synthesis 
since  he  has  isolated  a  mole  to  mole  ratio  of  XI  and  malonic  aldehyde,  the  latter 
being  trapped  by  a  thiobarbituric  reaction.  He  suggests  that  since  no  intermediates 
are  obtained  when  the  normal  reaction  is  run,  the  process  is  of  a  concerted  nature 
and  all  of  it  takes  place  while  the  substrate  is  attached  to  the  enzyme.  Also,  none 
of  the  products  isolated  could  be  transformed  into  PGEx  on  further  incubation. 


OH 
VIII 

R  -  (CHs)6eoc/°j 

R'*e  CsHxx 


Scheme  3 

The   esr  data  favor  the  peroxide  radical,  but  they  do  not  rule  out  the  hydro- 
peroxide since  the  esr  band  may  be  due  to  other  sources.  The   initial  attack  at  Cn 
seems  to  be  favored  by  the  isolation  of  VIII,  in  slight  contrast  uo  uue  fact  uhau  tne 


-z.-z.lt 


w>6  double  bond  compounds   (double  bond  6  carbons  from  the  CH3  end  of  the  chain)    are 


13 


the  most  specific  for  the  enzyme.1-3  A  free  carboxyl  group  is  also  required   and  a 
C20  fatty  acid  is  favored  over  C18>  C19j)  C21  or  C22  fatty  acids.15*16  A  systematic 
substrate  specificity  search  has  not  yet  been  done.  There  has  been  extensive  work 
done  on  inhibitors  and  co-f actors  by  Van  Dorp  and  co-workers „12 

A  simple  synthesis  of  the  prostaglandin  structure  was  used  in  the  structural 
determination  by  Samuelsson  and  Stallberg.17  They  succeeded  in  synthesizing  XIV  by 
two  methods  using  Grignard  and  condensation  type  reactions.  No  report  on  the 
biological  activity  was  given. 


XIV 


Since  then  there  have  been  several  syntheses  reported.  Bagli18  and  co-workers 
have  succeeded  in  making  XV,  a  derivative  of  PGEl0  They  started  with  the  potassium 
salt  of  ethyl  2»cyclopentanone  earboxylate  and  alkylated  it  using  7-bromoethyl- 
heptanoate.   It  was  then  monobrominated  and  after  refluxing  the  monobromo  derivative 
in  20$  HsS04  XVI  resulted.   On  reaction  of  XVI  with  acetone  cyanhydrin  in  NaC03  and 
H2Q/MeOH,  a  nitrile  was  obtained  which  yielded  the  trans  carhoxylic  acid  XVII  on 
hydrolysis,  A  monoesterification  to  XVIII  was  then  accomplished  by  refluxing  with 
a  0.5  mole  excess  of  p_-toluenesulfonie  acid  in  MeOH  for  55  minutes.  This  ester  was 
transformed  to  the  acid  chloride  XIX  and  the  acid  chloride  was  reacted  with  acetylene 


in  the  presence  of  A1C13  in  CCI4  to  form  a  vinyl  chloride  XX  which  yielded  the  methyl 
ketal  XXI  on  basic  hydrolysis  in  MeOH. 

H     *QH 

~0oRi 


XVI 


XVIII 
XIX 

jLa, 


XXI 


H' 


H' 


,c-c: 


„c-c> 

I    ! 

H  H 


,H 

*C1 

,0Me 

v0Me 


Me 


Me 


The  reaction  of  XXI  with  NaBRg.  yielded  a  mixture  from  which  XXII  was  obtained 
after  refluxing  in  2N  H2S04(>  This  compound  was  then  reacted  with  n-pentyl  magnesium 
bromide  to  yield  compound  XV. 

Since  then  the  Bagli  group  have  modified  their  synthesis.19  They  reacted  1- 
heptyne  with  XIX $   the  acid  chloride,  to  form,  the  vinyl  chloride  XXIII  which  was 
transformed  by  NaOH  in  MeOH  to  a  vinyl  methoxy  compound..  A  NaBEg.  reduction  followed 
by  hydrolysis  yielded  a  ketone  XXIV  which  on  further  NaBHi  reduction  resulted  in 
the  formation  of  XV. 

The  stereochemistry  of  the  product  was  shown  by  the  reduction  of  XVIII  with 
NaBEj,  to  yield  a  mixture,  85$  of  which  could  be  transformed  into  a  lactone  XXV.  This 
same  lactone  could  be  formed  from  the  product  of  oxidative  cleavage  of  XV  thus 
placing  the  hydroxyl  and  carboxyl  group  cis  to  one  another. 


The  trans  stereochemistry  of  the  two  chains  was  suggested  by  analogy  to  the  reaction 
in  which  only  the  trans  compound  is  obtained  when  cis  2,3  dialkyl  cyclopentanone  is 
re fluxed  in  base ,  as  was  XVII  when  it  was  formed  from  the  nitrile „ 


AJvJLJL 


XXIII 


Hs  i£H 


CO^Me 


&   A0. 


COpMe 


XXIV 


XXV 


The  compound  that  was  synthesized  (XV)  has  not  yet  been  identified  as  a  natural 
product,  but  it  still  possesses  a  vaso  depressor  effect s   which  is  amazing  in  view  of 
the  lack  of  oxygen  at  the  C1±   poistion.  A  method  for  adding  this  Cn  oxygen  function 
has  been  reported  by  Bagli,20  but  no  details  axe  available. 

The  first  total  synthesis  of  a  naturally  occurring  prostaglandin  has  been 
accomplished  by  Beal  and  co-workers22  who  succeeded  in  synthesizing  the  dihydro 
TGE±   XXXII  which  is  a  metabolite  present  after  the  incubation  of  TGE1   in  pig  lung 


enzyme  preparation 


21 


Beginning  with  a  formylation  of  3=ethoxy-2-cyclopentenone , 


they  obtained  almost  a  quantitative  yield  of  the  sodium  salt  XXVI.  This  compound 
then  underwent  an  in-situ  Wittig  reaction  to  give  compound  XXVII  which  on  catalytic 
reduction  and  a  second  formylation  gave  the  sodium  salt  XXVIII. 


EtO 


COsEt 


CO^Et 


XXVI 


XXVII 


XXVIII 


This  sodium  salt  was  then  put  through  a  second  modified  Wittig  reaction  using 
n-hexanoylmethylenetriphenylphosphonium  chloride  and  gave  XXIX  which  on  catalytic 
reduction  resulted  in  85$  XXX.  Acid  catalyzed  solvolysis  of  XXX  in  benzyl  alcohol 
resulted  in  the  corresponding  benzyl  enol  ether.  The  ketone  was  reduced  to  the  15 
hydroxy  compound  using  tri-t-butyl  aluminum  hydride  and  this  compound  was  then  con- 
verted to  the  diketone  XXXI  by  hydrogenolysis.  XXXI  was  rigorously  reduced  using 
30fo   rhodium  on  carbon  to  give  a  mixture,  11$  of  which  corresponded  by  tic  to  XXXII< 

OEt  0 


OsEt 


JOsEt 


XXIX 


AVvA 


.^6- 

^-^— 


HCL  ,H 


OsEt 


CO^Et 


AJLa,J= 


xxxii 


Compound  XXXII  was  compared  with  the  natural  dihydro  PGEX  using  ultraviolet, 
infrared,  mass  spectra,  nmr  and  tic.  It  was  also  shown  by  a  radio  isotope  dilution 
experiment  using  the  NaB3H4  reduced  product  and  crystalline  optically  active  natural 
(biosynthetic)  dihydro  PGFxp  methyl  ester  to  contain  at  least  22$  of  a  compound  with 
all  asymmetric  centers  identical  to  the  natural  compound.  If  the  reaction  had  been 
random,  it  would  have  contained  only  6$  of  the  desired  product. 

A  third  method  of  synthesis  by  Just  and  Simonovitch23*24  has  been  reported  in 
Chemistry  in  Canada.  They  oxygenated  the  hydroboration  product  from  cyclopentadiene 
and  protected  the  resulting  cyciopenta-3-en°l  as  the  tetrahydropyranol  ether  XXXIII. 
A  reaction  of  XXXIII  with  diazoacetic  ester  over  copper  powder  resulted  in  a  mixture 
°£   exo-endo,  syn-anti  compounds  which  on  refluxing  in  methanolic  sodium  methoxide 
resulted  in  an  exo,  syn-anti  mixture.  The  exo  compound  was  reduced  with  MAIH4  and 
the  alcohol  obtained  was  oxidized  to  the  aldehyde  XXXIV.  The  aldehyde  underwent  a 
Wittig  reaction  with  hexyltriphenyl  phosphonium  bromide  to  yield  four  isomers  which 
could  be  separated  by  tic.  Hydrolysis  in  refluxing  0.5$  oxalic  acid  in  methanol 
resulted  in  the  alcohol  which  was  oxidized  using  Jones  reagent  to  the  ketone  XXXV 
as  a  mixture  of  cis -trans  isomers.  These  isomers  were  separated  by  tic,  the  trans 
having  a  957  cm"x  band  in  the  infrared. 


XXXIII 


XXXIV 


XXXV 


The  alkylation  of  XXXV  was  accomplished  using  seven  equivalents  of  methyl -7- 
iodoheptanoate  in  dimethoxye thane  and  potassium  t-but oxide  as  the  base.  The  enamine 
alkylation  was  also  attempted  but  gave  a  lower  yield.  The  alkylation  was  made 
difficult  by  the  problems  of  separation  of  XXXV  from  the  product  XXXVI,  and  the 
instability  of  compound  XXXV.  An  effective  separation  was  accomplished  by  reduction 
of  the  ketone  with  NaBH*  to  yield  XXXVII. 


CO-Me 


C02Me 


XXXVI 


XXXVII 


The  acid  obtained  by  NaOH  hydrolysis  in  aqueous  methanol  of  XXXVII  was 
oxidatively  solvolyzed  using  one  equivalent  of  H^Og  in  a  NagCOs  buffered  solution 
of  formic  acid,  followed  by  shaking  in  10$  aqueous  Na^Qa  and  resulted  in  dl  PGF^ 
XXXVIII.  PGEX  was  also  obtained  by  a  similar  oxidative  solvolysis  of  XXXVI.  The 
PGFjq,  was  identical  bv  mass  spectra  and  tic  to  natural  PGF^. 


Several  biological  tests  were  run  and  the  dl  PGFxq,  was  found  to  have  one -half  the 
smooth  muscle  activity  of  natural  PGF^  while  the  activity  of  the  synthesized  PGEX 
compound  was  only  of  the  order  of  10-25$  of  natural  PGElo 


1. 

2. 

3. 

4. 

5. 

6. 

7. 
8. 


9. 
10, 

11. 

13. 
14. 

15. 
16. 

17. 
18. 

19. 
20. 

21. 
22. 
23. 

24. 


BIBLIOGRAPHY 

W.   Horton,  Experientia,  21,  113  (19^5)  <• 

Samuelsson,  Angew.  Chem.  Internat.  Edit.,  4,  klO   (I965). 

Bergstrom  and  B.  Samuels son,  An.  Rev.  Biochem. ,  34,  101  (1965) 
S.  Abrahams  son,  Acta  Crystallogr . ,  l6,  k0$)   (1963). 
D.  H.  Nugteren,  D.  A.  Van  Dorp  and  S.  Bergstrom,  M 


E. 
B. 
S. 


Nature,  212,  38  (1966). 


Hamberg  and  B.  Samuels son, 
257  (1966). 


M.  Hamberg  and  B.  Samuels  son,  J.  Biol.  Chem. ,  24l, 

H.  Vonkeman,  Chem.  Weekblad.,  62,  361  (I966). 

(a)  S.  Bergstrom,  H.  Danielsson  and  B.  Samuelsson,  Biochim.  Biophysic.  Act., 

90,  207  (1964) 1  (b)  S.  Bergstrom,  H.  Danielsson,  D.  Klenberg  and  B.  Samuelsson, 

J.  Biol.  Chem'.,  239,  PC  k006   (1964);  (c)  D.  A.  Van  Dorp,  R.  K.  Beerthius, 

D.  H.  Nugteren  and  H.  Vonkeman,  Nature,  20j5,  839  (1964);  (d)  D.  A.  Van  Dorp, 

R.  K.  Beerthius,  D„  H.  Nugteren  and  H.  Vonkeman,  Biochim.  Biophysic.  Acta., 

90,  204  (1964), 

D.  Klenberg  and  B.  Samuelsson,  Acta.  Chem.  Scand.,  19  (I965). 

R.  Ryhage  and  B.  Samuelsson,  Biochim.  Biophysic.  Res.  Coram.,  192,  79  (19^5). 

B.  Samuelsson,  J.  Am.  Chem.  Soc,  87,  3 Oil  (1965). 

D.  H.  Nugteren,  R.  K.  Beerthius  and  D.  A.  Van  Dorp 

405  (I966). 


Rec.  Trav.  Chim. ,  82, 


Struizk,  R.  K.  Beerthius,  H.  J.  J.  Pabon  and  D„  A.  Van  Dorp,  Rec.  Trav. 


Dorp,  Rec.  Trav.  Chim.,  85, 
Samuelsson,  J.  Biol.  Chem., 

Vonkeman,  Nature,  203, 


C.  B 

Chim.,  85,  1233  (1966). 
H.  J.  J.  Pabon,  L.  Van  der  Wolf  and  D.  A 
1251  (1966). 

S„  Bergstrom,  H.  Danielsson,  D.  Klenberg  and  B 
259,  PC   4006  (1964). 

D.  A.  Van  Dorp,  R„  K.  Beerthius,  D.  H.  Nugteren  and  H 
839  (1964). 
B.  Samuelsson  and  G„  Stallberg,  Acta.  Chem.  Scand.,  17,  810  (1963). 

J.  F.  Bagli,  T.  Bogri  and  R„  Deghenghi  and  K.  Wiesner,  Tetrahedron  Letters, 

465  ( 1966) . 

J.  F.  Bagli  and  T.  Bogri,  Tetrahedron  Letters,  5  (1967). 

J.  F.  Bagli,  10th  Annual  Medicinal  Chemistry  Symposium, 

June  I966.   (C.  and  E.  News,  44,  No.  27,  32  (196  " 

E.  Anggard  and  B„  Samuelsson,  J.  Biol  Chem. ,  239 
P.  F.  Beal,  J.  C.  Babcock  and  F.  H.  Lincoln,  J 
G.  Just  and  C.  Simonovitch.  Chem.  in  Can. ,  19. 


4097 


Bloomington,  Indiana, 
1964) . 


G.  Just,  private  communication,  March  20,  1967< 


Am.  Chem.  Soc., 
Jo.  1,  41  (I967). 


88,  3131  (1966) 


-338- 

POSSIBLE  VINYL  CATION  INTERMEDIATES 


Reported  by  David  A.    Simpson 


.May  1,  I.967 


Within  the  past  few  years  vinyl  cations  have  been  postulated  as  inter- 
mediates in  a  variety  of  reactions.  The  textbook  rule  regarding  the  reputed 
instability  of  these  ions  is  based  primarily  on  the  unreactivity  of  vinyl  halides 
toward  alcoholic,  silver  nitrate.  1>2     Thus  vinyl  chloride  on  long  heating  with 
solutions  of  silver  nitrate  in  ethanol  gives  no  silver  chloride.  However, 
recent  studies  have  questioned  the  general  instability  of  vinyl  carbonium  ions. 
It  is  the  purpose  of  this  seminar  to  review  those  reactions  which  have  been 
claimed  to  proceed  through  such  intermediates. 

The  first  postulation  of  a  vinyl  cation  intermediate  was  made  in  1951 
by  Newman  and  associates  as  a  result  of  studying  the  alkaline  decomposition 
of  3--nitroso-2-oxazolidones  (l)°   The  following  mechanism  was  proposed;3 


aldehyde s_9  ketones,  acetylenes" 
(vinyl  ethers) * 


Ri    P-2 

II 

C# 
I 

H 


A 


(h) 


R3— R4— H 


Pi 


OH 


HoO 


R4 


)COPH 


'NH-NO 


R2" 
R3" 


(2) 


R4 


OC02H 


N=N-OH 


B 


(3) 


R.3,R4/H 


aldehydes,  allylic  alcohols,1 
glycols,  cyclic  carbonates 
(acyclic  alky!  carbonates., 
acyclic  dialkyl  carbonate 
ethers)  * 


\ 


R2 
R3" 


81 


OC02H 


(5) 


*  reaction  carried  out  in  anhydrous  alcohol,  using 
the  corresponding  sodium  alkoxide 

The  first  step  was  thought  to  involve  ring  opening  of  (l)  to  a  nitrosamine  (2) 
followed  by  tautomerism  to  an  hydroxyazo  intermediate  ( 3) .   Two  paths  of  reaction 
are  now  open  to  (3)  :  path.  A  involves  the  initial  formation  of  an  unsaturated 
diazonium  hydroxide  intermediate  (by  the  base  catalyzed  elimination  of  carbonic 
acid)  followed  by  loss  of  nitrogen  to  yield  the  vinyl  carbonium  ion  (k) ,  and 
then  formation  of  products;  or  (3)  could  lose  nitrogen  first  yielding  the  sat- 
urated carbonium  ion  (s)  which  then  yields  products.  All  products  obtained  from 
the  4,4-disubstituted  oxazolidones  could  be  explained  by  path  B,  but  it  was 
necessary  to  invoke  path  A  to  account  for  the  products  resulting  from  the 


-339- 

5,5-disubstituted  derivatives.5  Although  path  A  seems  to  explain  all  the  exper- 
imental observations  adequately ,  a  mechanism  involving  an  intermediate  methylene  (6) 
cannot  be  rigorously  excluded*6 

RK   /*        OH"    R<      ••     -lb    \ 

C=<3         — - — >       ^C=C=N=<N.   — iH-4   ^C=Cj  ^   PRODUCTS 

R2    NW=U~0H  R/  R2 

(6) 

At  any  rate ,  there  does  not  seem  to  he  enough  evidence  to  support  any  of  the  pos- 
sible mechanisms  very  strongly  relative  to  the  others. 

Acetylene  derivatives  have  served  as  the  source  for  a  number  of  proposed 
vinyl  cation  intermediates.  The  earliest  example  was  reported  by  Jacobs  and 
Searles  who  claimed  that  the  hydration  of  acetylenic  ethers  appeared  to  be  sim- 
ilar to  that  of  vinyl  ethers.   That  is,  the  hydration  involved  a  rate-determining 
formation  of  a  carbonium  ion.7  They  measured  the  rates  of  hydration  of  several 
acetylenic  ethers  in  alcohol-water  solutions  by  a  dilatometric  method.   The 
rates  were  found  to  be  first  order  with  respect  to  acetylenic  ether  and  to 
hydronium  ion.  A  mechanism  consistent  with  the  observed  kinetics  was  formulated 
as  follows.8 

I.   R-Q~CrCH  +  H30+ +    [R-0~OCH2]+  +  R20 

II.   [R-0-C=CH2]+  +  H20  —^   [R-Q-G<SH2j+ 

6h2 

iii.    [r-c~c=€h2]+  +  h20  >  r-0-c=ch2  +  h30+ 

6h2  6h 

iv.    r-0-  — >  r-o-c-ch3 

OH  8 

Tne  authors  believed  that  the  first  step  was  rate-determining  and  that  the  reaction 
was  specific  acid-catalyzed. 

Sixteen  years  later  Drenth  and  co-workers  undertook  a  study  of  the  hydra- 
tion of  acetylenic  ethers  and  thioethers.  Their  rate  studies  largely  confirmed 
the  findings  of  Jacobs  and  Searles,  however s   the  reactions  were  found  to  be  general 
acid-catalyzed. 9j>lc  The  observation  of  general  acid  catalysis  indicated  that 
the  first  step  of  Jacobs  and  Searles9  mechanism  was  rate-determining.  Most  of 
the  work  of  Drenth  was  aimed  at  confirming  the  rate-determining  first  step  and 
educing  the  timing  of  the  addition  of  the  water  molecule.  Only  the  results  of 
their  investigations  of  the  acetylenic  thioethers  will  be  summarized  in  this  sem- 
inar. However,  many  of  the  same  experiments  and  results  were  also  shown  to  ap- 
ply in  studies  of  the  acetylenic  ethers. 

.Evidence  presented  for  a  rate-determining  proton  transfer  in  the  general 
acid-catalyzed  hydration  is  condensed  as  follows ;  (l)  the  reaction  is  faster  in 
H20  than  in  D2C>for  ethylthioethyne  the  ratio  of  k-.  .JK,   0  amounts  to  O.475 

(2)  substitution  of  a.  tertiary  butyl  group  for  the  ethyl  group  in  ethylthioethyne 
results  in  a  rate  enhancement.10  If  step  II  is  rate-determining,,  it  would  be 
expected  that  the  t-butyl  compound  would  hydrate  slower-  than  the  ethyl  derivative, 
since  the  addition  of  a  water  molecule  in  11  would  be  more  sterically  hindered 
in  the  case  of  the  t-butyl  derivative.  The  fact  that  the  latter  compound  is  hyd- 
rated  even  faster  than  the  ethyl  derivative  indicates  that  the  proton  transfer 
must  be  the  kineticaily  important  step  (application  of  the  Taft  equation  to  a 
series  of  thioethers  further  shows  that  steric  effects  are  not  essential).11 
And  ( 3) ,  infrared  analysis  of  recovered  ethylthioethyne  from  a  reaction  in  heavy 
water  shows  that  a  pre -equilibrium  is  not  important  in  this  reaction.12 

Since  the  addition  of  the  water  molecule  could  occur  simultaneously  with  the 
proton  transfer  or  in  a  subsequent  step,  Drenth !s  group  set  out  to  determine  the 


-3^0- 

timing  of  addition.  Abbreviating  their  findings:  the  activity  postulate  of 
Grunwald  for  reactions  in  alcohol-water  mixtures;  the  relation  of  Zucker  and 
Hammetty  and  entropy  considerations  all  seem  to  indicate  that  a  water  molecule 
is  not  covalently  involved  in  the  rate -determining  step  of  the  reaction.  There- 
fore, the  authors  concluded  that  addition  of  water  takes  place  in  a  step  subse- 
quent to  protonation. 12  The   evidence  presented  render  both  a  it -complex  mech- 
anism and  a  cyclic  transition  state  involving  the  ether,  hydronium  ion,  and 
water  improbable.   Thus,  the  rate -determining  formation  of  a  vinyl  cation  inter- 
mediate is  strongly  indicated. 

Peterson  has  reported  evidence  for  a  rate-determining  formation  of  a  vinyl 
cation  in  the  acid-catalyzed  hydration  of  phenylpropiolic  acids  (7)  and  phenyl- 
acetylenes.  Pseudc-first-order  rate  constants  for  the  acids  (7a-d)  were  measured 


(7) 


:5C-C02H 


a,  X=0CH3    c,  X=E 
b,   X=CH3     d,  X=C1 


+ 


spectrophotometrically  and  the  rate  data  were  found  to  correlate  with  cr,  p  =  -4.79+ 
0.02*  The  authors  claim  that  the  large  negative  value  of  p  indicates  a  high 
degree  of  positive  charge  on  the  benzylic  carbon  in  the  transition  state  and  there- 
fore that  the  C-OH2  bond  formation  occurs  after  proton  transfer.  In  addition 
a  solvent  kinetic  isotope  effect  implies  a  rate -determining  proton  transfer.13 
Further  studies  of  ortho,  meta,  and  para  substituted  phenylacetylenes  by  Bott, 
Eaborn,  and  Walton  again  suggest  a  rate-determining  proton  transfer.  This  is 
followed  by  a  rapid  reaction  of  the  formed  vinyl  carbonium  ion  with  solvent  to 
form  the  ketone.14 

Recently,  vinyl  cations  have  been  implicated  in  the  reactions  of  trifluoro- 
acetic  acid  with  alkynes.  In  a  preliminary  report  Peterson  and  Duddey  regarded 
the  vinylic  trifluoroacetates  formed  as  arising  from  intermediate  substituted 
vinyl  cations.15  This  communication  was  followed  by  a  more  detailed  study  of  the 
reactions  of  unsubstituted  hexynes  and  5-substituted-l-pentynes.  Evidence  for 
the  cationic  nature  of  the  transition  states  was  obtained  by  comparing  the  reactions 
of  the  substituted  alkynes  to  the  previously  studied  reactions  of  identically 
substituted  alkenes<>x6  A  Hammett-Taft  plot  (Taft's  a     values  for  trifluoroacetic 

acid  derived  from  fluorine  nmr  frequencies) 17  showed  the  rate  constants  for  the 
alkynes  were  only  slightly  smaller  than  those  of  the  alkenes,  and  demonstrated 
nearly  the  same  pattern  of  substitueni  effects  for  both,  indicating  similar 
cationic  like  transition  states.  The  study  of  the  reaction  of  3-hexyne  is  es- 
pecially interesting. 


OCOCF 


A 


CF3C02H 


OCOCF3 


DIMERS 
POLYMERS 


The  vinyl  trifluoracetates  (9)  and  (10)  were  formed  in  nearly  identical  amounts 

and  were  shown  to  be  stable  under  the  reaction  conditions.  Thus,  they  were  reported 


to  have  arisen  from  nonstereospecific  addition.,   The  yields  of  (9),  (10)  and 
(12)  were  found  to  vary  with  the  hexyne  concentration  as  shown  in  Table  I,18 

Tablg_I», 

Molarity  of  (9)  and  (10)  (12) 

3-hexyne   ^ mole  <p  mole  $ 

0.107  98  2 

lo03  2k  21 

2o00  18  30 

Besides  a  sweeping  generalization  that  vinyl  cations  are  readily  accessible  by 
addition  of  protons  to  alkynes,  the  authors  had  little  to  say  concerning  the 
above  findings.  However ,  it  is  tempting  to  interpret  the  observation  of  non- 
stereospecific addition  as  rendering  mechanisms  involving  a  jt -complex  or  con- 
cerned, type  transition  state  improbable  for  this  system.   The  product  depen- 
dence on  the  concentration  of  5"hexyne  could  be  rationalized  by  attack  of  the 
vinyl  cation  derived  from  3-hexyne  on  a  second  molecule  of  substrate. 

Peterson  and  Kamat  have  reported  the  formation  of  a  transition  state  re- 
sembling a  vinyl  cation,  in  the  related  trifluoroacetolysis  of  6~heptyn-2-yl 
p-toluene  sulfonate ,  -1 9 

Fahey  and  Lee  have  recently  studied  the  reaction  1-phenylpropyne  with  hyd- 
rogen chloride  in  acetic  acid.  The  observed  kinetics  (reaction  found  to  follow 
a  third-order  rate  law,  first-order  in  acetylene  and  secord -order  in  hydrogen 
chloride)  and  the  stereochemistry  of  the  products  lead  the  authors  to  propose 
a  mechanism  involving  intimate  and  solvent  separated  vinyl  carbonium-hydrogen 
dichloride  ion  pairs  as  intermediates.20 

Another  reaction  involving  protonation  of  an  unsaturated  system,  which 
may  lead  to  vinyl  cations,  is  the  addition  of  compounds  of  the  form  HX  to  allenes. 
The  chemistry  of  allene  has  recently  been  reviewed.21  The  addition  of  HX  to  al- 
lene has  been  observed  to  follow  Markovnikov ' s  rule.  For  a  long  time  it  was 
thought  that  2-substituted  propenes  and/or  2 ,2-disubstituted  propanes  were  the 
only  products  of  such  additions.22  However,,  Griesbaum  and  associates  have  recently 

HX 
HgC^C^CHg  — — — ^  H3C-CX—CH2  +  H3C-CX2-CH3 

(13)       (lh) 

found  such  reactions  to  be  more  complex.  It  was  reported  that  in  the  electro- 
philic  addition  of  hydrogen  bromide  to  allene  considerable  amounts  of  1,3-dibromo- 
l,3~dimethylcyclobutane  (two  isomers.)  were  formed,  as  well  as  the  conventional 
products  (13.)  arid  (l4)»23  Structural  proof  for  the  cyclic  compounds  consisted 
of  mass  spectral,  nuclear  magnetic  resonance „  and  infrared  data.  Additional 
evidence  was  obtained  by  reducing  the  dibromo  compound  with  tri-butyltin  hydride 
to  yield  a  mixture  of  the  isomeric  1,3-dimethylcyclobutanes. 

Formation  of  the  cyclic  products  by  a  thermal  reaction  was  considered  im- 
probable since  thermal  dimerization  of  allene  leads  to  a  1,2-1,2  adduct,  while 
hydrogen  bromide  addition  causes  a  formal  1,2-2,1  dimerization.  Reaction  of  2- 
bromopropene  (13,  X=Br)  with  hydrogen  bromide  under  the  conditions  of  the  allene 
addition  led  exclusively  to  2,2-dibromopropane.   Ultraviolet  irradiation  of 
2-bromopropene  gave  essentially  starting  material  and  showed  no  traces  of  the 
cyclic  products.  Thus,  the  authors  concluded  that  a  simple  dimerization  of 
2-bromopropene  was  not  Involved.  Furthermore,  since  the  free  radical  addition 
of  hydrogen  bromide  to  allene  yields  no  cyclic  products,24  the  cyclodimerization 
was  thought  to  be  an  ionic  reaction.   Since  the  reaction  of  methylacetylene  with 
hydrogen  bromide  also  led  to  cyclo-dimerization,25  and  because  the  rearrangement 
of  allene  to  methylacetylene  in  the  presence  of  hydrogen  bromide  could  not  be 
ruled  out,  the  following  path  was  proposed.26 


-3^2- 


.+ 


CH2=C=CH2 


CH2  -=0  =CH2 


H 


+ 


v 


II 


CH3-C— CH2 
(15) 


X 


CH3-CX=CH2 
X 


II 


.+ 


1.  X 


CH- 


2.  HX 


CH- 


X 


CH3-C=CH 


CH3C^CH 


1.  X 


2.  HX 


While  the  intermediate  steps  in  the  above  scheme  remain  unclear,  the  authors 
felt  that  the  vinylic  cation  (15)  was  involved  in  the  cyclizations.   Some  data 
for  the  addition  of  hydrogen  bromide  to  allene  and  methylacetylene  is  reported 
in  Table  II. 

Table  II.26 


Relative  amounts  of  components  in  adduct  mixture,  wt.  70 
Substrate     BrCH=CHCH3     CH3CBr=CH2     CH3CBr2CH3 


Br 


Br 


Br 


Run#1    ch£Sh2 


30 


Rnn  Jio*  CH3C=CH 
Run  #2*  CH2=C=CH2 


17 

2k 

2k 

3k 

23 
33 

6 

9 

.56 
60 

25 
27 

9 
10 

3 
.3 

*  run  7r2  contained  hydroquinone  as  inhibitor 


Evidence  that  the  l=bromopropene  is  formed  by  a  concurrent  free  radical  addition 
is  shown  by  the  effect  of  the  inhibitor  on  its  yield.  If  one  ignores  the  free 
radical  addition  product,  the  methylacetylene  and  allene  additions  show  the  same 
selectivity  for  the  remaining  products.  It  should  be  mentioned  that  the  deter- 
mination of  product  yields  In  Table  II  is  not  completely  clear.  The  authors 
claim  that  the  components  comprised  75-98$  of  the  total  adduct  mixtures  (depend- 
ing on  the  purity  of  the  starting  materials) .  However,  it  was  not  indicated  how 
this  estimate  was  made  (no  mention  of  internal  standard  use).  Accordingly  these 
values  may  not  be  too  reliable «,  Furthermore,  a  mechanism  Involving  concerted 
addition  cannot  be  rigidly  excluded. 

A  possible  route  to  a  vinyl  cation  could  be  deamination  of  vinyl  amines,, 
Curtin's  group  has  examined  several  such  deaminations .  The  reaction  of  2,2- 
diphenylvinylamine  (16)  with  nitrosyi  chloride  is  especially  relevant  to  this 


UlbL'USSl'J-l. 


The  formation  of  (17);  (18)  and  (19)  was  explained  in  terms  of  a 


-3^3- 


<t> 


/ 


t 


.CI 


$         m2 
c=c 

*(l6)* 


4>C=C<t>  +        v 

(17)      H   (18)    CI 
13#*        # 


+ 


h'(i9)no 

6$ 


CHgCxg 
2.5° 


CI 

(*)2C 

30^ 


■H  +  NO 


+ 


4>-C-<t> 


15^ 


*  amounts  of  products  obtained  are  reported  as  mole  per  cent  based  on  amine 
employed 

carbonium  ion  rearrangement  from  the  ion  (20) ,  formed  by  loss  of  nitrogen  from 
the  diazonium  ion,  to  the  ion  (21).  The  ion  (2l)  could  lose  a  proton  to  form  the 
acetylene  (17)  or  suffer  attack  by  chloride  ion  to  give  the  cis-  (18)  and  trans- 
(19)  chlorides. 


C=C 

(20) 


4>C=C 


s 


H 


0x  ^ra  c: 

/c=cv 

V  H 


(21) 


(22) 


The  nearly  random  formation  of  these  three  products  is  of  particular  interest. 
The  authors  interpret  the  results  as  being  consistent  with  three  differently 
oriented  vinyl  carbonium-chloride  ion  pair  intermediates.   The  possibility  of 
(20)  fragmenting  into  a  phenyl  cation  and  phenyl  acetylene  followed  by  readdition 
was  deemed  improbable  since  gas  phase  chromatography  of  the  product  mixture  showed 
no  traces  of  phenylacetylene  or  chlorobenzene.   The  product  of  replacement  without 
rearrangement ,  2 ,  2-diphenylvinylchloride^was  shown  to  be  present  to  a  maximum 


However,  an  intermediate  bridged  ion  could  not  be  rigorously  excluded 


28" 


Of  2$ 

by  the  authors. 

All  attempts  to  intercept  the  diazonium  intermediate  (22)  with  the  sodium 
salt  of  P-naphthol  failed.  It  was  concluded  that  the  diphenylvinyl  system  did 
not  provide  the  necessary  stabilization  of  (22)  and  de stabilization  of  (20) 
and  (21)  which  is  required  for  trapping.  The  instability  of  the  diazonium  ion 
was  attributed  to  possible  participation  of  the  aromatic  ring  in  the  transition 
state  for  loss  of  nitrogen  and/or  freedom  of  the  vinylcarbonium  ion  to  assume 
sp  hybridization.29  A  study  of  3~atf1ino-2-phenylindenone  (23)  was  initiated  in 
hopes  that  it  would  be  free  of  these  difficulties.  Again  attempts  to  intercept 
the  diazonium  intermediate  failed.  However,  during  the  reaction  of  (23)  with 
nitrosyl  chloride  at  -10°  an  infrared  absorption  appeared  at  2090  cm"1.   Observation 
of  this  peak  over  a  period  of  time  showed  that  it  disappeared  according  to  a  first- 
order  rate  law.   The  authors  stated  that  no  definite  assignment  could  be  made  to 
this  absorption,  but  it  might  be  ascribed  to  a  diazonium  or  diazoaikane  intermediate30'31 

Studies  by  Grob  and  Cseb  on  the  solvolysis  of  a-bromostyrenes  {2k)   have 
questioned  the  "well-known"  general  unreactivity  of  vinyl  halides.   The  first-order 


(24) 


R=  -H 
-NH2 
-OCH3 


rate  constants  for  0.01  M  solutions  of  these  compounds  in 
triethylamine  were  determined  and  appear  in  Table  III.32 


-NHCOCH3 
-W02 

&  ethanol  with  0.0.1  M 


~NH2 

0.00 

-OCH3 

100.10 

-WHCOCH3 

115.20 

-H 

170. Q0 

-344- 

Table  III. 
(24)       temperature       k^ec"*1)  k  , 

-Ms  0.00       9.57(9.119)*  X  10"5        5.5  X  108 

3.60(3.60)  x  10"5      8.5  x  103 

3.80(3.92)  X  10~5        2.2  X  103 
6.00(6.80)  X  10"6  1 

*values  in  parenthesis  are  in  the  presence 
of  0.05  M  triethylamine 

The  following  additional  observations  were  made  concerning  these  reactions:  (l) 
preparative  solvolyses  in  the  presence  of  triethylamine  or  calcium  carbonate  gave 
only  the  corresponding  acetophenones  from  the  substituted  a-bromostyrene s  (24s 
r=  -NH2,  -OCH3  and  -NHCOCH3) ,  a-bromostyrene  (24;  R=  -H)  yielded  in  addition 
22/®  phenylacetylene;  (2)  the  reaction  of  the  p-nitro  derivative  was  exclusively 
second-order  and  gave  only  p-nitrophenylacetylene;  (3)  each  of  the  substituted 
phenyiacetylenes  was  shown  to  be  stable  under  the  reaction  conditions;  (4)  after 
a  short  warming  with  silver-nitrate  in  80$  ethanol  the  methoxy  derivative  (24: 
R=  -OCH3)  gave  a  precipitate  of  silver  bromide ,  the  nitro  compound  (24:  R=  -N02) 
after  several  hours  at  100°  gave  no  precipitate;  (5)  and  a-bromostyrene  (24: 
R=  -H)  in  50$  ethanol  reacts  ten  times  faster  than  in  the  less  ionizing  80$ 
ethanol  solution.32 

Based  on  these  observations ,  the  authors  propose  that  the  a-bromostyrene 
derivatives  (24;  R=  -H,-MI2,  -0CH3,-NHC0CH3)  react  via  an  S^l  -  El  type  mechanism 

involving  the  intermediate  formation  of  a  vinyl  cation,  and  that  the  nitro  compound 
(24:  R=  -N02)  reacts  by  a  bimolecular  elimination  (E2)  process.  However,  the 
following  facts  must  be  considered.  First,  products  were  never  isolated  from  the 
kinetic  runs  themselves,  and  the  preparative  solvolyses  were  not  run  under  identical 
conditions  with  the  former.   Secondly,  no  data  or  plots  of  the  kinetic  studies 
were  given,  only  the  resulting  rate  constants  were  reported.  And  thirdly,  the 
authors  claim  that  the  phenylacetylene  formed  from  a-bromostyrene  (24:  R=  -H) 
could  not  arise  from  a  competitive  E2  process  due  to  the  small  deviations  observed 
in  the  first-order  rate  constants  on  addition  of  five  molarequivalents  of  tri- 
ethylamine. However,  the  question  remai.ns  whether  a  competing  E2  process  could 
have  been  detected  by  the  workers.  Crude  calculations  indicate  that  a  competitive 
E2  process  here  appears  not  unlikely.  Further  indication  that  the  reactions  may 
be  more  complex  than  the  authors  claim  appears  in  the  magnitudes  of  the  relative 
rate  constants  (Table  III).  The  reported  accelerations  do  not  parallel  the  known 
electrical  properties  of  the  substltuents.  In  spite  of  these  criticisms  the  relatively 
facile  reaction  of  alcoholic  silver-nitrate  with  the  methoxy  derivative  (24: 
R~  -OCH3)  indicates  that  further  studies  of  "stacked"  vinyl  halide  systems  may  prove 
interesting. 

In  a  related  study  Grob  and  co-workers  have  examined  the  solvolytic  decarboxy- 
lation of  the  potassium  salts  of  cis  and  trans  a,£-unsaturated-J3-halo  acids. 
Salts  of  the  cis  series  when  heated  in  aqueous  solution  yielded  the  corresponding 
acetylenes,  while  those  of  the  trams  series  gave  ketones  in  addition  to  acetylene 
derivatives.  The  reaction  of  the  cis  salts  was  explained  by  a  concerted  mechanism, 
on  the  other  hand  a  rate-determining  ionization  to  an  intramolecularly  solvated 
vinyl  carbonium  ion  was  proposed  for  the  solvolysis  of  the  trans  salts.33 

Stable  cations  having  a  contributing  vinyl  cation  resonance  form  have  reportedly 
been  observed  by  nmr.  Richey's  group  has  examined  several  systems,  but  only  the 
propynyl  and  ethynyl-di-p-methoxyphenylcarbinol  (25)  case  will  be  discussed. 
Extraction  of  the  propynyl  compound  (25:  R-  -CH3)  from  carbon  tetrachloride  into 

CH3O-4  >-  C-CEC-R  R-  H  or  CH3 

(85)    0 

OCHc, 


-3^5- 

concentrated  sulfuric  acid  gave  solutions  whose  nrar  was  assumed  to  be  that  of  the 
corresponding  alkynyl  cation.   Strong  evidence  for  this  was  obtained  by  neutralizing 
the  sulfuric  acid  solutions  which  gave  approximately  yof/o  of  the  starting  alcohol. 
The  absorption  of  the  propynyi  methyl  group  (t  7°  ^-0)  appeared  considerably  downfield 
from  the  absorption  of  the  same  group  in  a  carbon  tetrachloride  solution  of  the 
substrate  (x  8.13).34  The  absorption  of  the  corresponding  ethynyl  derivative 
(25;  R=  -H)  snowed  a  much  larger  downfield  shift,   (from  T  7° 35  for  the  alcohol  to 
t  4.30  for  the  ion}  The  authors  suggest  that  this  indicates  that  the  carbon  to 
which  the  ethynyl  hydrogen  is  attached  is  significantly  involved  in  charge  dereal- 
ization as  shown  by  the  resonance  structure  (26).35 

$/  $    / 

-c=c-c  <e- — -»  -c=c=c 

(26)  ^ 

Footnotes  36-38   are  additional  references  to  reactions  in  which  vinyl  car- 
bonium  ions  have  been  postulated.  However,  due  to  their  similarity  to  systems 
already  discussed  and  the  limited  amount  of  information  available  they  will  not 
be  examined  in  this  abstract. 

In  conclusion,  evidence  presented  seems  to  indicate  that  in  certain  systems 
the  intermediacy  of  vinyl  cations  best  explains  the  experimental  results.   Such 
findings  cast  doubt  on  the  inferred  general  instability  of  such  intermediates. 
In  the  past  these  intermediates  have  been  largely  rejected  because  of  the  unreactivity 
of  vinyl  halides  toward  alcoholic  silver-nitrate.  However,  as  several  workers  have 
pointed  out,39-'40  the  inertness  of  vinyl  halides  is  based  on  ethylene  derivatives 
and  may  be  less  applicable  to  more  highly  alkylated  systems.  In  any  case  a  definitive 
study  of  the  reactivity  of  various  vinyl  halides  toward  alcoholic  silver-nitrate 
appears  to  be  missing  from  the  literature.   As  a  result  the  chemistry  of  vinyl 
carbonium  ions  is  somewhat  ill  defined. 

BIBLIOGRAPHY 

1.  R.  L.  Shriner,  R,  C,  Fuson,  and  D.  Y.  Curt-in,  "The  Systematic  Identification 
of  Organic  Compounds" ,  John  Wiley  and  Sons,  Inc.,  New  York,  1956,  P«  1^1  • 

2.  J.  D.  Roberts  and  M.  C.  Caserio,  "Basic  Principles  of  Organic  Chemistry", 
W.  A.  Benjamin,  Inc.,  New  York,  1964,  p.  321. 

3.  M,  S,  Newman  and  A.  F.  Weinberg,  J.  Am.  Chem.  Soc,  JJ3,  4654  (I956). 

4.  M.  S.  Newman  and  A.  Kutner,  ibid.  ,  73,  ^199  (1951)  •' 

5.  M,  S.  Newman  W.  M.  Edwards,  ibid.,  36,   1840  (1954). 

6.  J.  Hine,  "Divalent  Carbon",  'The  Ronald  Press  Company,  New  York,  1964,  p.  89. 

7.  A.  J.  Kresge  and  Y.  Chiang ,  J.  Chem.  Soc,  53  (1967),  and  references  cited  therein, 

8.  T.  L,  Jacobs  and  S.  Sear.Ies,  Jr.,  J.  Am.  Chem,  Soc,  66 ,  686  (1944). 

9.  E.  J,  Stamhuis  and  W.  Drenth,  Rec  Trav.  Chim, ,  80,  797  (I96.I). 

10.  W,  Drenth  and  H.  Hogeveen,  ibid. ,  7£,  1002  (1960)0 

11.  H.  Hogeveen  and  W.  Drenth,  ibid. ,  82,  375  (1963). 

12.  H,  Hogeveen  and  W.  Drenth,  'ibid. ,  B|,  410  (1963). 

13.  D.  S.  Noyce 9  M.  A.  Matesich,  M.  D.  Schiavel.li,  and  P.  E.  Peterson,  J,  Am, 
Chem.  Soc,  87,  2295  (19^5)  • 

14.  R.  W.  Bott,  C.  Eaborn,  and  D.  R.  M.  Walton,  J.  Chem,  Soc,  384  (1965). 

15.  P.  E.  Peterson  and  J.  E.  Duddey,  J,  Am.  Chem.  Soc,  85,  2865  (1963). 

16.  P.  E.  Peterson  and.  G.  Allen,  J.  Org.  Chem, ,  27.,  2290  (1961) . 

17.  R.  W.  Taft,  J.  Am.  Chem.  Soc,  8£,  709  (I963T7 

18.  P.  E.  Peterson  and  J.  E.  .Duddey,  ibid.  ,  88,  4490  (1966). 

19.  P.  E.  Peterson  and  R.  J.  Kamat,  ibid. ,  88,  3152  (I966). 

20.  R.  C.  Fahey  and  Do  Jae  Lee,  ibid. ,  88,  5555  (1966). 

21.  K.  Griesbaum,  Angew.  Chem.  Intern.  Ed.  Engl.,  5,  933  (1966). 

22.  T.  L.  Jacobs' and  R.  N.  Johnson,  J.  Am.  Chem.  Soc. ,  82,  6397  (i960). 

23.  K.  Griesbaum,  ibid.  .  86,  2501  (1964). 

24.  K.  Griesbaum,  A.  A.  Oswald,  and  D.  N.  Hall,  J.  Org.  Chem.,  29,  2404  (1964). 

25.  K,  Griesbaum,  Angew.  Chem.  Intern.  Ed.  Engl.,  3,  697  (1964). 


-346- 

26o  Ko  Griesbaum,  W.  Naegele,  and  G.  G.  Wanless,  J.  Am.  Chem.  Soc. ,  87,  3151 

( 1965)  • 
2T»  Do  Yo  Curtin,  J»  A.  Kampmeier,  and  R.  0! Connor ,  ibM0  >  ^L>   ^63  (1965)  • 
280  Do  Yo  Curtin,  private  discussion,  April  4,  1967* 

29«  J«  Ao  Kampmeier,  Ph. Do  Thesis,  University  of  Illinois,  i960,  p.  24. 
30.  D.  Yo  Curtin,  J.  A.  Kampmeier,  and  M.  L.  Farmer,  J.  Am.  Chem.  Soc,  87, 

87^  (l965)o 
3L  Mo  Lo  Farmer,  Ph.D.  Thesis,  University  of  Illinois,  1964,  pp0  29=30o 

32.  Co  A.  Grob  and  Go  Cseh,  Heiv.  Chim.  Acta,  4j,  194  (1964). 

33.  Co  A.  Grob,  J.  Csapilla,  and  G.  Cseh,  ibid.,  47,  1590  (1964), 

34.  H.  G.  Rlchey,  Jr.,  J.  C.  Philips,  and  L.  E.  Rennick,  J.  Am.  Chem,  Soc,  87, 
I38I  (I965). 

35*  H.  G.  Richey,  Jr. ,  L.  E.  Rennick,  A.  S.  Kuchner,  J.  M.  Richey,  and  J.  C. 

Philips,  ibid. ,  87,  4017  (1965). 
360  Ho  Vieregge ,  H.  M.  Schmidt,  J.  Renema,  H.  J.  T.  Bos,  and  J.  F.  Arens,  Rec 

Trav.  Chim.,  8£,  929  (1966). 

37.  Ho  Wo  Whitlock,  Jr.  and  P.  E.  Sandvick,  J.  Am.  Chem.  Soc,  88,  4525  (1966). 

38.  R.  Wo  Bott,  C.  Eaborn,  and  D.  Ro  M.  Walton,  Organometal  Chem.,  1,  420  (1964). 

39.  P.  E.  Peterson  and  J.  E.  Duddey,  J.  Am.  Chem.  Soc,  85,  2865  (1963). 

40o  No  Co  Deno,  "Progress  in  Physical  Organic  Chemistry",  Volo  II.,  Interscience 

Publishers,  New  York,  1964,  p.  181. 
4l.  Wo  M.  Jones  and  F.  W.  Miller,  J»  Am.  Chema  Soc,  8£,  i960  (1967). 


-2*7- 

SIGMATROPIC  REACTIONS 


Reported  by  R.  N.  Watson 
INTRODUCTION 


May  11,  1967 


The  concept  of  the  sigmatropic  reaction  was  derived  by  Woodward  and  Hoffmann,1 
on  the  basis  of  their  molecular  orbital  calculations,  to  correlate  formally  a  large 
number  of  separate  reactions „  The  definition  and  general  characteristics  of  sigma - 
tropic  reactions,  as  well  as  reactions  which  illustrate  the  various  types  of  sigma - 
tropic  changes,  are  discussed.  Emphasis  has  been  placed  on  the  geometry  of  the 
transition  state  involved  in  these  reactions,  and  the  correspondence  with  molecular 
orbital  predictions.  Although  the  majority  of  the  mechanistic  work  has  been  done  on 
thermal  sigmatropic  reactions,  the  available  evidence  for  photochemical  reactions  is 
also  considered,, 

GENERAL  CHARACTERISTICS 

Woodward  and  Hoffmann1  define  a  sigmatropic  reaction  of  order  (i,j)  as  an 
uncatalyzed,  intramolecular  reaction  in  which  a  sigma  bond,  flanked  by  one  or  more 
pi -electron  systems,  migrates  to  a  new  position  whose  termini  are  i-1  and  j-1  atoms 
removed  from  the  original  bonded  loci.  This  terminology  conforms  with  that  pre- 
sently in  the  literature  (thus  a  (1^5)  hydrogen  shift  is  also  a  (1,5)  sigmatropic 
shift).2  Although  not  specifically  mentioned  by  Woodward  and  Hoffmann  as  criteria 
for  sigmatropic  reactions,  it  is  found  that  many  of  the  reactions  are  thought  to  be 
concerted,  and  many  have  been  found  to  be  reversible  also.   Some  generalized  illus- 
trations of  sigmatropic  changes  are  given  below. 


v* 


P 


^K^ 


A- 


(1.5)  ; 


(  5  »3)   y 


hT 


2   1 


(5*5)  > 


It  is  believed  that  orbital  symmetry  relationships  are  the  main  factors  which 
determine  the  course  of  these  reactions.3  For  example,  the  (l,j)  sigmatropic 
migration  of  hydrogen  within  the  all-cis  polyolefinic  framework  (I  •>  II)  may  take 
place  by  two  paths,  suprafacial  or  antarafacial  (the  transition  state  for  this 
change  is  thought  of  as  consisting  of  a  hydrogen  atom  and  a  radical  containing 
2k  +  3  pi -electrons) .  In  the  suprafacial  process,  the  hydrogen  appears  at  all  times 


1 


rH  r 


R' 


CH 


on  the  same  face  of  the  pi-system,  with  the  transition  state  having  a  plane  of 
symmetry,  G.   In  the  antarafacial  process  the  migrating  hydrogen  passes  from  the 
top  face  of  one  carbon  to  the  bottom  face  of  another,  with  the  transition  state 
having  a  two-fold  axis  of  symmetry,  C2. 1     It  is  found  that,  in  order  to  maintain 


-348- 
positive  overlap  between  the  highest  occupied  orbital  of  the  olefin  system  and  the 
hydrogen  orbital,  the  isomerization  I  ■»  II  must  occur  thermally  (ground  state 
orbital  symmetry)  by  the  suprafacial  path  when  k  is  odd,  and  antarafacially  when  k 
is  0  or  even.  These  results  are  reversed  for  first-excited -state  transformations, 
and  are  supported  by  extended  Huckel  M. 0.  calculations.3  However,  if  the  migrating 
group  possesses  an  available  low-lying  pi-orbital,  alternative  transition  state 
processes  may  occur.  The  symmetry-allowed  (l,j)  sigmatropic  transformations 
(assuming  a-orbital  interaction  of  the  migrating  group  with  the  n-system)  for  j  ^7 
are  given  below  (Table  I). 

Table  I1 


Symmetry  Allowed  Transformations 


(1.J) 

(1,3) 

(1,5) 
(1,7) 


Thermal 

Antarafacial 

Suprafacial 

Antarafacial 


Excited  State 

Suprafacial 
Antarafacial 


For  sigmatropic  reactions  in 
which  i  and  j  are  greater  than  1, 
proceeding  through  transition 
states  with  a  plane  of  symmetry, 
thermal  changes  are  symmetry 
allowed  when  i  +  j  -  4n  +  2, 
whereas  excited -state  trans - 


Suprafacial 

formations  are  symmetry  allowed  when  i  +  j  =  kn.1     Apparent  sigmatropic  reactions 
which  violate  these  rules  may  be  taking  place  through  multi-step  processes,  perhaps 
involving  diradical  intermediates,  but  these  are  expected  to  require  vigorous 
conditions. 

(1,3)  SIGMATROPIC  REACTIONS 

No  established  examples  of  thermal,  uncatalyzed  (1,3)  hydrogen  shifts  could  be 
found  (antarafacial).  Woodward  and  Hoffmann  believe  this  is  because  the  carbon 
framework  must  not  become  so  distorted  during  reaction  as  to  cause  impairment  of 
coupling  within  the  pi-system.1  Thus,  the  antarafacial  process  would  be  difficult 
or  impossible  for  j  a  3. 

Photochemical  (1,3)  reactions  can  be  postulated  for  at  least  two  cases,  based 
on  product  analysis.  When  4,10-dimethyl-A3*5-hexalin  (III)  was  irradiated  in  pentane, 
a  GOf/o   yield  of  the  isomerized  non-conjugated  diene  (IV)  was  isolated,  along  with  other 

hydrocarbons.4  That  this  reaction  is 
intramolecular  and  uncatalyzed  was  not 
demonstrated,  but  when  a  similar  system, 
7,5,82 9, H-e;rgostadiene,  is  irradiated,  it 
is  found  that  the  reaction  is  independent 

TTT  ■**   TV 

AA±  xv      of  protic  and  aprotic  solvents  and  gives 

an  isomeric  non -conjugated  diene  with  the  original  ergostane  skeleton.  No  other 
mechanistic  work  was  done,  but  a.  (1,3)  hydrogen  shift  can  rationalize  these  products. 
A  (1,3)  shift  involving  a  carbon-carbon  sigma  bond  was  postulated5  for  the  photo- 
chemical conversion  of  verbenone  (V)  to  chrysanthenone  (VI) . 


hv 


H. 


H 


,H. 


VII 


vn: 


IX 


(1,5)  SIGMATROPIC  REACTIONS 

There  have  been  a  large  number  of  reactions6"20*42"45  which  involve  the  (1,5) 
sigmatropic  migration  of  a  hydrogen  atom.  Frey  and  Ellis6  showed  that  cis-2-methyl- 
l,3~pentadiene  (VII)  undergoes  a  reversible,  first  order  isomerization  in  the  range 
!97-237°C  to  4~methyl"l,3-pentadiene  (VIII) .  The  entropy  of  activation  of  this 
reaction  was  approximately  -8  e.u.,  the  negative  value  being  attributed  to  the  loss 
of  two  internal  free  rotations  in  going  from  the  reactant  to  a  6-membered  cyclic 
transition  state  (IX).  Wollnsky  and  co-workers7  observed  this  (1,5)  migration  in  a 
number  of  other  1,3-dienes  with  a  vinyl  and  alkyl  group  in  a  cis  arrangement,  while 
trans -1,3-dienes  are  stable  in  the  same  conditions.  Thus  the  reversible  rearrange- 
ment of  X  gives  15$  XI  at  260°C,  and  100$  XI  at  360°. 

Since  a  higher  temperature  is  required  for  the  equilibration  of  non -planar  1,3- 
dienes  (XII  does  not  reach  equilibrium  with  XIII  at  450°C) ,  and  since  the  trans -1,3- 


dienes  are  stable  ( trans -2-methyl-l, 3 -pentad iene  gave  less  than  1.0$  of  rearrange- 
ment products  at  k^0u)  ,  they  suggest  that  the  transition  state  has  5  carbon  atoms 
in  a  plane  (or  nearly  planar)  with  the  migrating  hydrogen  above  or  below  the  plane. 
This  is  consistent  with  Woodward  and  Hoffmann's  prediction  of  a.  suprafacial  route. 


X 


XII 


XIII 


r 


The  rearrangement  of  cyclopentadienes  also  proceeds  by  (1,5)  hydrogen  shifts. 8'9 
Mclean  and  Haynes8  determined  the  entropy  of  activation  for  the  conversion  of  5- 
methylcyclopentadiene  to  1-methylcyclopentadiene  to  be  -10  e.u.  On  the  basis  of  an 
n.m.r.  study  of  deuterated  cyclopentadiene  in  the  range  k^^6^°C9   Roth9  concluded 
that  the  first  order  rearrangement  which  resulted  in  a  statistical  distribution  of 
deuterium  was  an  intramolecular  succession  of  (1,5)  shifts.  Roth  also  investigated 
the  rearrangement  of  1-deutero-indene  (XIV)  which  was  thought  might  possibly  undergo 
a  (1,3)  shift  (from  indene  to  indene)  in  preference  to  a  (1,5)  shift  (from  indene  to 
the  non-aromatic  isoindene  (XV),  then  back  to  indene).  By  observing  the  relative 
intensities  of  the  three  non-aromatic  hydrogen  positions  of  1-deutero-indene  (3. 18, 
3.51>  an^  6.72f)in  the  n.m.r.  until  the  deuterium  is  statistically  distributed,  he 
concludes  that  a  (1,3)  hydrogen  shift,  does  not  occur  at  all,  and  estimates  that  the 
energy  difference  between  a  (1,3)  said  a  (1,5)  shift  in  cyclopentadiene  is  at  least 
11.5  kcal./mole. 


(1.5). 


XIV 


(l*5'K 


The  thermal  rearrangements  of  cycloheptatrienes  (recently  reviewed  by  Shermer10) 
involving  the  migration  of  hydrogen  around  the  ring  have  been  shown  to  involve  the 
suprafacial,  intramolecular  (l,5)sigmatropic  shift  of  hydrogen  to  the  apparent- 
exclusion  of  all  other  mechanisms.10 

Similarly,  (1^5)  migrations  have  been  observed  in  8-membered  ring  compounds.11"18 
Careful  integration  of  the  n.m.r.  spectrum  at  various  intervals  of  the  equilibration 
of  neat  l,3=cyclo8ctadiene  (XVTa°d)  at  150°  for  2k   hours  gave  kinetic  results  which, 
according  to  Glass,  Boikess,  and  Winstein,11  are  consistent  only  with  a  series  of 
successive  intramolecular  (1,5)  hydrogen  shifts  (AS'  s  -10  e.u.).  It  can  be  noticed 
that  (1,5)  hydrogen  shifts  have  comparable  rates  and  activation  energies  in  six 
seven,10  eight,13  and  nine13  member ed  rings  and  in  open-chain  systems. 


12 


6*14 


V 


/y   \^D 


XVI 


D 


v_/a 


Rearrangements  within  the  cycloSctatriene  system  are  also  thought  to  proceed  by 
a  facile  (1,5)  suprafacial  sigmatropic  shift,  mechanism. 13*17-*18  Roth15  demonstrated 
that  1,3,6-eyeioo'ctatriene  and  1,3 ,5 -cycloSctatriene  are  in  equilibrium  at  225°  by 
means  of  a  (1,5)  hydrogen  shift.  The  rate  of  this  reversible  reaction  is  first  order 
and  is  not  influenced  by  solvent  polarity.  Although  bicyclo[l|-.2o0]octa-2,4~diene  is 
present  in  the  equilibrium  mixture,  Roth  believes  that  this  is  formed  from  the  1,3,5- 
triene  only  and  is  not  an  intermediate  in  the  isomerization  between  the  two.  Roth 
rules  out  any  large  contribution  due  to  a  (1,3)  shift  in  this  isomerization  by 
observing  the  rearrangement  of  7j>8-ciideutero«.l,3,5=cyclodctatriene.  If  a  (1,3)  shift 
is  present  the  deuterium  atoms  would  be  statistically  distributed  through  all  positions 
on  the  ring,  whereas  for  a  (1,5)  shift  the  deuteriums  are  limited  to  the  J>9k   and  7*8 
positionsa  The  n.m.r.  of  rearranged  material  did  not  show  a  statistical  distribution 
through  all  positions.  Roth  also  computed  the  least  value  for  the  energy  difference 


-350- 

between  the  (1,5)  shift  and  the  hypothetical  (1,3)  shift  to  be  7.5  kcal./mole. 

The  rearrangement  of  5  j  8 -bis -(a-cyanoisopropyl)  -l,3,6~cyclo£Jctatriene  to  3,Q- 
bis -(a-cyanoisopropyl) -bicyclo  [^„2.0]octa- 2,4-diene  is  believed  to  proceed  through 
the  1,3,5  isomer.16  Kice  and  Cantrell16  think  the  1,3*6  isomer  would  exist 
preferentially  with  the  bulky  a-cyanoisopropyl  groups  in  a  quasi -equatorial  con- 
figuration (XVII,  R  -  a-cyanoisopropyl) ,  in  this  configuration  the  migrating  hydrogen 

would  be  located  directly  over  one  of  the  double  bonds  and 

well-situated  for  a  suprafacial  process. 

Photochemical  (1,5)  sigmatropic  reactions  ( antaraf acial) 

have  been  postulated  to  occur  in  several  open  chain  systems,18'19 

but  only  one  (8-membered)  cyclic  system.20  Thus,  the  irradia- 


tCf 

XVII 
the  allene  (XIX). 


XVIII 


tion  of  allo-dcimene  (XVIII)  was  reported  by  Crowley18  to  give 
That  antarafacial  (1,5)  processes  have  not  been  observed  for  small 

ring  compounds  would  seem  to  substantiate 
Woodward  and  Hoffmann's  selection  rules. 


hv 


R-CH3 


;i,T)    SIGMATROPIC  REACTIONS 


XJDC 


•24 


The  (1,7)  sigmatropic  shift  should 
occur  by  the  antarafacial  route  in  ground 
state  reactions  and  by  the  suprafacial 
route  in  excited  state  reactions.1  A 
thermal  (1,7)  hydrogen  migration  has  been  proposed  for  the  interconversion  of^ 
precalciferol  and  calciferol  and  for  conversions  in  analogous  triene  systems. s 
Schlatmann21  converted  cis°l-oyclohexylidene-2-(  5 8 -hydroxy-2 ' -methyleneeyclo- 
hexylidene)  ethane  ( XX)  into  cis~l~(  cyclohex-1 '  -ene)  -2-(  5  ,,-hydroxy-2,t-methylcyclohex- 
l"-ene)ethene  (XXI)  by  heating  at  70-90°  for  a  few  hours.  Schlatmann  determined 
that  the  reaction  is  found  only  with  conjugated  triene  systems  with  a  cis -configuration 

at  the  central  double  bondj 
that  the  reaction  rate 
(see  Table  II)  is  independ- 
ent of  solvent,  acids, 


A 


the  entropy  of  activation  is  negative  ( ~ 


bases,  arid 

inhibitors 


free  radical 
,  that  there  is 


no  exchange  with  CH3QD 
during  reaction 5  and  that 


le  II.  Rate  Constants  of  the  Isomerization  (XXf=^XXI) 
Under  Different  Conditions  at  60.8°C.22 


Medium 
decalin 


0   ethanol 
ethanol  + 
Q.8xlO"3MHCl 


k  in  see  • 
If.lxlO"5 

4.6xl0~5 
if.5*10~5 


Medium 


%   ethanol  + 
2.3xlO~3M  (CaHs)3W 


70   ethanol  + 
1.5xlO~4M  hydroquinone 


k  in  sec"1 
3.7xl0°"5 

4.7xl0"5 


Therefore  he  concludes  that  the  reaction  is  intramolecular  and  occurs  via  a  rigid 
cyclic  transition  state  (XXII) .  The  configuration  of  these  systems  will  easily 

permit  an  antarafacial  process. 

The  photochemical  (1,7)  hydrogen  shifts  of  cycloheptatriene 
systems10^25"27  in  contrast  with  the  (1,5)  thermal  behavior, 
gives  significant  confirmation  of  Woodward  and  Hoffmann's 
selection  rules.  Both  ter  Borg25  and  Roth26  observed  a  series 


.O  J.i. 


of  (1,7)  shifts  in  7-deutero~i,3,5-cycloheptatriene  (XXIIa)  by 


integrating  the  n.m.r.  spectrum  of  irradiated  samples.  Both 
report  the  initial  formation  of  the  1-deutero-isomer  (XXIIb) ,  and  believe  the 
equilibrium  is  completely  consistent  with  a  set  of  consecutive  (1,7)  shifts  ( XXIIa -d). 
In  a  similar  n.m.r.  study,  Murray  and  Kaplan27  found  that,  although  l,4-bis(7- 
cycloheptatrienyl) benzene  undergoes  a  series  of  thermal  (1,5)  hydrogen  shifts,  the 
compound  undergoes  (1,7)  shifts  on  irradiation.  Their  analysis  of  the  spin-spin 


■351- 


XXII 


coupling  of  this  compound  leads  them  to  believe  that  it  exists  in  the  preferred 
conformation  XXIII0  Thus,  positive  overlap  could  easily  be  maintained  between  the 
migrating  hydrogen  orbital  and  a  framework  orbital  at  the  terminus  of  the  migration, 
thereby  providing  easy  access  to  a  suprafacial  shift  in  each  case. 

Although  the  (1,3)  suprafacial  shift  is 
allowed  photochemically,  it  has  not  been  con- 
clusively demonstrated  as  occurring  in  these 
systems  concurrently  with  the  (1,7)  shift.  However 
Murray  and  Kaplan57"  believe  that  the  formation  of 
2~phenyleycloheptatriene  instead  of  1 -phenyl - 
cyclopehtatriene  from  7-phenylcycloheptatriene  is 
possibly  indicative  of  a  preferential  (1,3)  shift 
over  the  (1,7)  shift.25  To  explain  the  observed 


XXIII 
Rs4-(  7-cycloheptatrlenyl) 


phenyl 
dominance  of  the  (1,7)  shift,  Woodward  and  Hoffmann1  believe  that  the  higher  values 
of  j  are  preferred  in  order  to  achieve  a  maximum  degree  of  linear  conjugation  in 
the  transition  state. 

(3,3)  SIGMATROPIC  REACTIONS 

The  most  well-known  examples  of  (3,3) sigmatropie  reactions  are  the  Cope28"32 
and  Claisen33"38  rearrangements.  Both  reaction  types  are  thermal  (thermal  changes 
symmetry -allowed  when  i  +  jskn  +  2-61)  ,  intramolecular,  relatively  insensitive  to 
catalysis,  and  show  negative  entropies  of  activation,  thus  satisfying  the  general 
characteristics  of  sigmatropic  reactions.28"30  The  primary  concern  here  will  be  the 
influence  of  orbital  symmetry  in  determining  the  orientation  of  the  transition  state 
in  these  reactions. 

If  the  transition  state  of  the  Cope  rearrangement  is  depicted  as  a  complex  of 
two  allyl  radicals  situated  in  roughly  parallel  planes ,  then  the  question  of  the 
geometry  of  the  transition  state  is  whether  the  two  allyl  radicals  are  bound  between 
all  three  pairs  of  atoms  ( six-center  boat  form,  XXIV)  or  only  bound  through  the 


four  terminal  atoms  ( four -center  chair  form,  XXV) 


28  "30 


1   2 


or 


//  \\ 


XXIV 


XXV 


the  meso-isomer  would  give  a  mixture  of  cis,  cis  and  trans 


If  the  rearrangement  proceeds 
by  the  four-center  path, 
then  rac-3 A -dimethyl -1,5- 
hexadiene  will  produce 
only  cis ,  cis  and  trans , 
trans -octadiene ,  however 
the  me  s  o - c  propound  will 
produce  only  cis,  trans - 
octadiene  by  this  path.28""31 
By  the  six-center  path, 
trans -2  96 -octadiene , 


and  the  racemic  isomer  must  rearrange  to  cis,  trans -2,6-octadiene.  Doering  and 


trans=2,6-oetadiene  from  meso= 


Roth31  obtained  almost  exclusively  (99.77°)  cijL 

3, ^-dimethyl =1,5 -hexadiene  at  225°C,  and  found  that  the  rac -isomer  gave  IQffi  cis 
cis-  and  90fo  trans ,  trans  -octadiene .  Therefore  the  rearrangement  proceeds  by  the 
four-center  transition  state  (XXV)  s   the  free  energy  difference  between  the  ^-center 
and  6-center  arrangements  was  calculated  as  at  least  5°7  kcal./mole. 

Similarly,  Marvell  and  co-workers36  determined  by  conformational  analysis  that 
the  transition  state  of  the  Claisen  rearrangement  of  cis-  and  trails -a,7 -dimethyl 
allyl  phenyl  ethers  is  best  represented  by  a  cyclic  four-center  chair  form.  From 
the  rates  of  rearrangement  of  cis  and  trans -y -substituted  allyl  aryl  ethers  and 
those  of  p -alkylallylaryl  ethers,  White  and  Norcross33*34  also  conclude  that  the 
chair  configuration  is  preferred. 

A  qualitative  explanation  for  this  preference  was  offered  by  Doering  and  Roth.30 
In  an  allyl  radical  the  energetically  lowest  molecular  orbital  contains  two  electrons 


distributed  fairly  uniformly  throughout  the  system,,  whereas  the  next  higher  orbital 
has  only  one  electron  with  a  very  small  electron  density  at  the  central  carbon. 
Therefore  two  aliyl  radicals  with  this  combination  of  orbitals  will,  repel  each 
other  at  all  points  and  can  bond  only  at  the  terminal  carbons,  which  is  best 
represented  as  the  chair  form. 

Fukui  and  Fujimoto32  applied  a  simple  Huckel  perturbation  calculation  to  the 
double  allyl  system  and  determined  the  energy  increase  due  to  weak  conjugation 


between  atomic  it   orbitals  at  carbons  2 
op  j  J 


2s  (of  XXIV)  to  be  AE  a  2P, 


22: .   7  9 


where 


^22'  -  2  2  C^  Cs'  is  a  measure  of  the  "overlap  stabilization"  (C2  and  C2'  denote  the 

J 
coefficients  of  2pjt  atomic  orbitals  of  the  j  th  M.  0. ,  the  summation  being  carried  out 
over  all  occupied  orbitals),  and  7  is  the  resonance  energy  of  2,2s  conjugation. 
Their  calculations  give  AE~5  °*°  6  kcal./mole  which  is  consistent  with  Doering's31 
experimental  value  of  5.7  kcal./mole. 

Woodward  and  Hoffmann ,39  in  a  somewhat  different  manner,  also  suggest  that 
orbital  symmetry  relationships  play  a  predominant  role  in  determining  the  preference 
for  the  chair -like  transition  state.  A  correlation  diagram  was  drawn  for  the 
hypothetical  process  of  two  allyl  radicals  approaching  each  other  from  infinity,  in 
parallel  planes.  These  motions  involve  two  symmetry  elements ;  0lt   the  plane 
passing  through  carbons  2  and  5%,   and  02,   a  plane  parallel  to  the  radical  planes  in 
the  boat  form,  or  a  C2  axis  perpendicular  to  O^  in  the  chair  form.  The  end  products 


AS  4V- 
SS  -4+ 

ss  4f- 


0X02 

or 
tfiC2 


SA 
AA 


SS 


-Hr-SA 


41-  AS 

4f-ss 


in  this  hypothetical  motion  are  a  bicyclohexane  in  the  boat  case  and  cyclohexyl 
biradical  in,  the  chair  case.39  The  essential  difference  in  the  two  approaches  is  in 
the  behavior  of  the  occupied  SA  level  in  which  the  boat  approach  correlates  to  an 
antibonding  a  orbital,  while  in  the  chair  form  it  goes  over  to  a  non-bonding  radical 
orbital.  Thus  at  any  point,  the  chair-like  transition  state  is  at  a  lower  energy 
as  a  result  of  the  difference  in  the  correlation  properties  of  the  SA  orbital.  By 
comparing  this  diagram  with  the  actual  correlation  diagram  for  the  (3,3)  reaction, 
Woodward  and  Hoffmann  predict  that  the  chair  form  of  the  transition  state  is  of 
lower  energy. 

It  appears  reasonably  certain  that  the  course  of  the  (3,3) sigmatropic  reactions, 
such  as  the  Claisen  and  Cope  rearrangements  is  determined  primarily  by  the  orbital 
symmetry  requirements.  This  supports  Woodward  and  Hoffmann's  basic  assumption  that 
all  sigmatropic  reactions  are  directed  by  the  orbital,  symmetry. 


-353- 


(3,5)  SIGMATROPIC  REACTIONS 


The  existence  of  the  (3,5)'  process  has  not  been  effectively  demonstrated,  but 
it  could  be  a  partial  explanation  for  the  observation  that  7~14C-allyl-(2,6- 
dimethyl  phenyl) -ether  (XXVI)  rearranged  under  irradiation  for  2  l/2  days  at  25°  to 
give  l.kio   of  the  ^ -allyl  phenol  ( XXVII)  with  label  distributed  fairly  evenly  in  the 
a  and  y   positions  (at  20-30°  the  rate  of  the  thermal  rearrangement  is  practically 
zero,  and  would  give  100$  retention  of  label  in  the  y   position) „40  Since  2,6- 
dimethyl  phenol  was  formed  in  9$  yield  in  these  same  conditions,  this  could  be  a 


0~CH2~CHs€H2 
'  &   .  P  7 


XXVI 


hv 


AJvV  XX 


Position  Jo   Label 


J 


2.35 


CH2-CH-CH2 
Of       p     7 


97. 


Position     Jo  Label 


t] 


hi 
53 


case  of  dissociation  to  free  radicals  and  recombination.  A  more  definitive  examination 

of  this  reaction  was  not  done, 

A  thermal  (3,5)  type  process  has  been  considered  by  Fahrni  and  Schmid29*41  to 

explain  the  isotopic  distribution  obtained  by  heating  allyl-7-14C-mesityl  ether 

(XXVIII-is  not  capable  of  phenol  formation)  for  96  hours  at  170°  in  diethylaniline . 

The  heated  material  had  radioactivity  almost  evenly  distributed  between  the  a   and 

7  atoms,  whereas  the  starting  material  had  label  exclusively  in  the  7  atom  of  the 

allyl  group.  This  result  is  not  possible  by  the  accepted  Claisen  mechanism  of  a 

sequence  of  (3,3)  steps. ^  A  possible  pathway  which  could  account  for  this  result 

and  still  preserve  the  observed  intramolecular  character  of  the  rearrangement  is  a 

(3,5)  sigmatropic  ortho-ortho1  rearrangement.  This  type  of  rearrangement  may  also 

explain  small  discrepancies  in  the  isotopic  distribution  found  in  the  equilibrium 

of  other  allyl  ethers.29  Although  reasonable,  this  mechanism  has  yet  to  be  fully 

established . 

* 

.s^S*  0  *      0  0^ 


3^L 


xr/111 


(3,5X 


I5i3) 


(5,5)  SIGMATROPIC  REACTIONS 

Woodward  and  Hoffmann1  as  well  as  Fukui  and  Fujimoto32  predict  from  molecular 
orbital  considerations  that  a  chair-like  transition  state  would  be  preferred  in  the 
(5,5)  sigmatropic  shift  of  cis 9   cis-decatetraene  (XXIX). 


XXIX 


SIGMATROPIC  REACTIONS  INVOLVING  CYCLOPROPANE  RINGS  AND  HEIEROATOMS 

It  has  been  noted  that  a  cyclopropane  ring  may  replace  a  it -bond  in  the  frame- 
work system  for  sigmatropic  changes.1  This  observation  has  been  very  veil  documented 
in  the  literature. n*i3>i4*3o>42-45  It  has  been  f0Und  that  cis-l-methyl-2-vinyl- 
cyclopropane  rearranges  by  a  (1,5)  shift-  at  temperatures  above  l60°  to  cis°hexa-1.4- 
diene,42  (+)  -trans ^3-hydroxymethyl°A4 -carene  (XXX)  rearranges  stereospecifically 
( suprafacial)  at  200°  to  ('  =)  «6~hydroxymethyl~A2'8-p-menthadiene  (XXXI),14  1,4-cyclo- 
Sctadiene  and  bicyclo[5.1o0. ]  oct~2-ene  are  in  thermal  equilibrium  above  l80°((l,5) 
shift),43  and  (1,5)  sigmatropic  shifts  are  observed  in  the  rearrangements  of 

bicyclononadienes  to  cyclononatrienes.11,13,45 
iL^CH.3       Cyclopropyl  ring  intermediates  are  proposed  for 
v^\^CH2CH    the  Abnormal  Claisen  rearrangement46*47  and 
similar  systems.48"50 

The  incorporation  of  an  oxygen  in  the 
^  ^/Sr         framework  system  has  already  been  noted  in 

^   XXX  -z?\~      XXXI    the  Claisen  rearrangement.  There  is  evidence 

that  nitrogen  may  also  participate  in  sigma- 
tropic reactions.  Staab  and  co-workers51"54  believe  a  Cope  type  (3.°3)  sigmatropic 
rearrangement  (XXXII  •*•  XXXIII)  occurs  in  the  thermal  isomerization  of  double  Schlff 
bases  of  1,2-diaminocyclopropane  (XXXII)  to  2,3~disubstituted-lH,l,4-diazepines 
(XXXIV).  An  extreme  example  of  the  variations  possible  in  the  (1,5)  sigmatropic 
reaction  is  given  by  the  rearrangement  of  l-(p-nitrobenzoyl) ~2,2~dimethyl  aziridine 
(XXXV)  to  N-O-methallyl)  -l»p-nitrobenzamide  (XXXVI)  in  which  a  3-membered  ring, 
nitrogen,  and  oxygen  atoms  participate  in  the  framework  system.55"56 


2000 


■7 


N=CH-R 


(3,3) 


N N 

XXXIII 


CH3-C-CHa 

V 


1. 

2. 

■z 

J' 

4. 

5. 
6. 


o=c=/^Wo2 


11^ 


CH«^ 
I 
CH^^C "CHa 

N 


:-^Ws 


xxxv 


H0~C=<'    ">-AM^£ 


BIBLIOGRAPHY 


XXXIV 

CH3   H  0  /—x 
CH2=6  -CHg-N-C  =T_N >N02 


XXXVI 


R.  B.  Woodward  and  R.  Hoffmann,  J.  Am.  Chem.  Soc,  87,  2511  (1965). 

H.  M.  R.  Hoffmann,  Annual  Reports,  62,  246  (1965). 

R„  E.  Cunningham,  Jr.,  Univ.  of  111.  Org.  Seminars,  I  Semester,  I965-I966, 

p.  103. 

W.  G.  Dauben  and  W.  T.  Wipke,  Pure  Appl.  Chem.,  9,  539  (1964). 

J.  J.  Hurst  and  G.  H.  Whitman,  J.  Chem.  Sec.,  2864  (I960). 

H.  M.  Prey  and  R.  J.  Ellis,  J.  Chem.  Soc,  4770  (1965'. 


7.  J.  Wolinsky,  B.  Chollar,  and  M.  D.  Baird,  J.  Am.  Chem.  Soc.,  84,  2775  (1962) 

8.  S.  McLean  and  P.  Haynes,  Tet.  Let.,  2385  (1964). 

9.  W.  R.  Roth,  ibid.,  1009  (1964). 

10.  W.  D.  Shermer,  Univ.  of  111.  Org.  Seminars,  II  Semester,  1966=1967,  p.  272. 

11.  D.  S.  Glass,  R.  S.  Boikess,  and  S.  Winstein,  Tet.  Let.,  999  (1966). 

12.  E.  N.  Marvell,  G.  Caple,  and  B.  Schatz,  ibid.,  389  (1965). 

13.  Do  S.  Glass,  J.  Zirner,  and  S.  Winstein,  Proc.  Chem.  Soc.,  276  (1963). 

14.  G.  Ohioff,  Tet.  'Let.,  3795  (1965). 

15.  W.  R.  Roth,  Ann.,  671,  25  (1964). 

16.  J.  L.  Kice  and  T.  S.  Cantrell^  J.  Am.  Chem.  So:..,  8^,  2298  (1963). 


-355- 

17.  J.  M.  Conia,  F.  Leyendecker,  and  C.  Dubois -Faget,  Tet.  Let.,  129  (1-966). 

18.  K.  J.  Crowley,  Proc.  Chem.  Soc. ,  17  (1964). 

19.  R.  Srinivasan,  J.  Am.  Chem.  Soc,  84,  3982  (I962). 

20.  J.  ZIrner  and  S.  Winstein,  Proc.  Chem.  Soc.,  235  (1964). 

21.  W.  R.  Messer,  Univ.  of  111.  Org.  Seminars,  II  Semester,  1965=1966,  p.  10. 

22.  J.  L.  Sehlatmann,  J„  Pot,  and  E.  Havinga,  Ree.  Trav.  Chim. ,  &3,  1173  (1964). 

23.  R.  L.  Autry,  D.  H.  Barton,  A.  K.  Ganguly,  and  W.  H.  Reusch,  J.  Chem.  Soc,  3313 
(1961). 

24.  R.  L.  Autry,  D.  H.  Barton,  and  W.  H.  Reusch,  Proc  Chem.  Soc,  55  (1959). 

25.  A.  P.  Ter  Borg  and  H.  Kloosterziel,  Rec  Trav.  Chim.,  84,  24 1  (1965). 

26.  W.  R.  Roth,  Angew.  Chem.,  25_,  921  (I963). 

27.  R.  W.  Murray  and  M.  L.  Kaplan,  J.  Am.  Chem.  Soc,  88,  3527  (1966). 

28.  J.  C.  Gaal,  Univ.  of  111.  Org.  Seminars,  Summer,  19&3 s   P«  1°« 

29.  S.  J.  Roads  in  P.  de  Mayo,  ed.,  "Molecular  Rearrangements,"  Interscience 
Publishers,  New  York,  196.3^  Part  I,  p.  655. 

30.  W.  von  Doering  and  W.  R.  Roth,  Angew.  Chem.  Int.  Ed.,  2,  115  (I963). 

31.  W.  von  Doering  and  W.  R.  Roth,  Tetrahedron,  18,  67  (19o°2). 

32.  K.  Fukui  and  H.  Fujimcto,  Tet.  Let.,  251  (ISJSG)  . 

33.  W.  N.  White  and  B.  E.  Norcross,  J'.  Am.  Chem.  Soc,  8j5,  1968  (I96I). 

34.  Ibid.,  3265  (I96I)  . 

35.  L.  D.  Huestis  and  L.  J.  Andrews,  ibid.,  1963  (1961). 

36.  E.  N.  Marvell,  J.  L.  Stephenson,  and  J.  Ong,  ibid.,  87,  I267  (1965). 

37.  E.  N.  Marvell,  B.  J.  Burreson,  and  T.  Crandall,  J.  Org.  Chem.,  30,  IO3O  (I965). 

38.  E.  N.  Marvell,  B.  Richardson,  R.  Anderson,  J.  L.  Stephenson,  and  T.  Crandall, 
ibid.,  1032  (I965). 

39.  R.  B.  Woodward  and  R.  Hoffmann,  J.  Am.  Chem.  Soc,  87,  4390  (1965). 

40.  R.  Schmid  and  H.  Schmid,  Helv.  Chim.  Acta.,  36,  6^7  (1953)  • 

41.  P.  Fahrni  and  H.  Schmid,  ibid.,  42,  1102  (1959). 

42.  R.  J.  Ellis  and  H.  M.  Frey,  Proc.  Chem.  Soc,  221  (1964). 

43.  W.  Grimme,  Chem.  Ber. ,  £8,  756  (1965). 

44.  W.  R.  Roth  and  J.  Kflnig,  Ann.,  688,  28  (1965). 

45.  W.  R.  Roth,  ibid.,  631,' 10  (19&TJ7 

46.  W.  M.  Lauer  and  T.  A.  Johnson,  J.  Org.  Chem.,  28,  2913  (1963). 

47.  R.  M.  Roberts,  R.  G.  Landolt,  R.  N.  Greene,  and  E.  W.  Heyer,  J.  Am.  Chem.  Soc., 
8g,  l4o4  (1967). 

48.  R.  M.  Roberts  and  R.  G.  Landolt,  ibid.,  8£,  2281  (1965). 

49.  R.  M.  Roberts,  R.  N.  Greene,  R.  G.  Landolt,  and  E.  W.  Heyer,  ibid.,  87, 
2282  (1965). 

50.  D.  E.  McGreer,  N.  W.  K.  Chiu,  and  R.  S.  McDaniel,  Proc.  Chem.  Soc.,  415  (1964). 

51.  H.  A.  Staab  and  F.  VSgtle,  Tet.  Let.,  51  (I965).' 

52.  H.  A.  Staab  and  F.  Vflgtle,  Chem.  Ber.,  £80  2691  (1965). 

53.  Ibid.,  £§,  2701  (I965). 

54.  H.  A.   Staab  and  C.  Wtfnsche,  ibid.,  3479  (1965). 

55.  D.  V.  Kashelikar  and  P.  E.   Fanta,  J.  Am.   Chem.   Soc,  82^  4930  (i960). 
^6.  P.   E.   Fanta  and  M.  K.  Kathan,  J.  Heter.   Chem.,  1,  293  (1964). 


-356- 
PHOTOCHEMISTRY  OF  CYCLOBUTANONE S  AND  CYCLOBUTANEDIONES 


Reported  by  Edward  F.    Johnson 


May  15,  196? 


CYCLOBUTANONES 

The  photochemistry  of  cyclobutanone  in  the  gas  phase  has  been  reviewed  by 
Srinivasan1  and  compared  to  the  known  photochemistry  of  other  cyclic  ketones. 
In  general,  the  primary  photochemical  process  for  cyclic  ketones  is  thought  to 
be  a  ring  opening  to  a  biradical  which  then  undergoes  rapid  dissociation  to 
stable  molecules.  Carbon  monoxide  is  split  out,  leaving  cyclic  hydrocarbons  and 
various  olefins;  in  addition,  unsaturated  aldehydes  are. formed  by  abstraction 
of  a  6  hydrogen.  Cyclobutanone  undergoes  vapor  phase  reactions  similar  to  those 
of  its  higher  homologs,  with  the  exception  that  only  a  very  minor  amount  of  the 
unsaturated  aldehyde  is  formed.2 

Photolysis  of  cyclic  ketones  in  the  liquid  phase  is  often  dependent  on  the 
choice  of  solvent.3  The  first  step  in  the  photochemical  process  can  be  formally 
represented  by  the  fission  of  a  CO-^  bond  to  give,  most  commonly,  the  most  sub- 
stituted  aryl/aikyl  biradical.   The  biradical  will  then  stabilize  itself  by  one 
of  three  paths;  (a)  hydrogen  transfer  from  the  carbon  atom_a_to  the  carbonyl 
group  to  form  a  ketene,  (b)  hydrogen  transfer  to  the  carbonyl  group  to  form  an 
unsaturated  aldehyde,  or  (c)  loss  of  carbon  monoxide.  The  loss  of  CO,  well 
known  in  the  gas  phase s  is  seen  only  for  certain  ketones  in  the  liquid  phase.4 

The  photolysis  of  cyclobutanone  in  inert  solvents  has  been  found  to  give 
the  same  products  as  in  the  gas  phase;  namely  ethylene,  ketene,  propylene, 
and  cyclopropane o5-*6  This  decomposition,  both  in  vapor  and  condensed  phases, 
can  be  explained  in  terms  of  one  primary  photochemical  process  (scheme  1)  in 
analogy  to  -fhebond  cleavage  reported  for  the  vapor  phase  reactions  of  other  cyc- 
lic ketones.  First,  there  is  an  a  cleavage  from  the  n  -»  n  state  of  1  to  yield 
the  hypothetical  biradical  2  which  decomposes  by  either  cycloelimination  or  de- 
carbcnylation,  pathway  A  or  B,  respectively. 

Scheme  1 


h  V 


n  ->  it 


J> 


Path  A 


Path  B 


->      CH2=CH2       +       CH2=C=0 


■1 


A    *   A 


In  the  presence  of  a  reactive  solvent,  a  third  decomposition  pathway  is 
seen  to  be  operative.  Hcstellter7  irradiated  the  bicyclic  ketone  J  in  methanol 
and  isolated  two  products  in  nearly  equal  amounts.  He  proposed  the  carbene  h 
and  the  ketene  j?  as  intermediates  to  explain  the  products.   No  direct  evidence 
was  presented  for  the  presence  of  the  carbene. 


-357- 

Quinkert,  Cimbollek,  and  Buhr5  have  examined  the  photochemistry  of  C-lj 
epimeric  D-nor~l6-keto  steroids  6  and  7> 


2  I 

In  benzene,  6  and  2  Sive  the  decarbonylation  product  8  and  the  cycloelimination 
product  £  in  product  ratios  of  6l:39  for  6  and  8^92  for  7,  On  the  other  hand, 
acetals  are  the  major  products  if  the  steroid  is  irradiated  in  ethanol  solution 
as  shown  in  scheme  2.   Steroid  J  also  gives  products  showing  complete  retention 

Scheme  2 


hv 


+ 


OEt 


+  8  +  Q 


of  configuration.,  The  configuration  was  determined  "by  spectral  properties  and 
by  chromic  acid  oxidation  to  the  lactone.   The  resulting  lactones  were  compared 
to  authentic  samples  prepared  from  the  original  steroids  by  Baeyer-Villigar 
oxidation.   Therefore,  the  retention  of  configuration  seems  to  rule  out  a  "free" 
alkyi/acyl  biradical. 

Turro  and  Southam6  have  also  studied  this  ring  expansion  reaction.  Irrad- 
iation of  cyclobutanone  in  deutero-methanol  leads  to  nearly  exclusive  formation 
of  10.  This  product  offers  additional  support  for  the  carbene  intermediate, 
and  rules  out  a  potential  mechanism  involving  a  hydrogen  shift  and  addition  of 
methanol  across  a  double  bond.  The  cleavage  has  been  shown  to  favor  the  most 


<P 


1. 


CH30D 


f\ 


0 


4 


10 


OCH3 


CH3OD 


OCH3 

substituted  bond  a_to  the  carbonyl  group.  If  2,2-dimethylcyclobutanone  is 
irradiated,  none  of  the  acetal  resulting  from  the  cleavage  of  the  C4— -0  bond 
is  found.  Cyclopentanone,  when  irradiated  in  ethanol,  does  not  undergo  ring 
expansion,  but  forms  4-pentenal  instead.  Yates  and  Kilmurry8  report  a  case 
in  which  a  tricyclic  cyclopentanone  undergoes  ring  expansion  by  way  of  an  0x0= 
carbene.  If  d-cyclocamphanone,  11,  is  irradiated  in  cydc?ne.>sne,the  carbene  is 
trapped,  and  the  resulting  product  was  isolated  and  shown  to  be  12. 


■358- 


h-y" 


» 


HEKZOCYCLOBUIENEDIOHES 

Oxocarbene s  have  also  been  proposed  in  the  photolysis  of  benzocyclobuten- 
edione,  1^,  in  solution.   Brown  and  Solly9  have  isolated  two  dimers  in  the 
photolysis  of  1£  in  degassed  cyclohexane  using  natural  sunlight  filtered  through 
pyrex.  The  dimers,  14  and  1J?  (5.5  and  38$  yield  respectively) ,  are  proposed  to 
result  from  the  reaction  of  the  oxocarbene  16  with  1^  or  an  excited  state  of  13„ 
More  definite  evidence  for  the  oxocarbene  results  from  the  work  of  Staab  and 
Ipaktschi.10  They  irradiated  1^  in  pentane;  methylene  chloride  (2:1)  solutions 
at  20°  using  a  mercury  high  pressure  lamp  and  isolated  three  dimers^  Ik  (25%) , 
the  cis  isomer  of  14  (5$),  and  17  (h$) .      No  evidence  is  presented  to  show  whether 
1J  is  formed  as  a  result  of  a  photochemical  process  on  1^  or  as  a  result  of 
rearrangement  of  another  product .   The  oxocarbene  16  is  trapped  as  the  ethyl 
acetal  by  irradiating  13  in  refluxing  alcohols  while  a  control  experiment  carried 
on  without  irradiation  yielded  no  acetal  after  refluxing  in  alcohol  for  12  hours. 


±2 


0  ±k 


</ 


A 


16 


Xl  R2R3  R4 
18 


Irradiation  of  1£  in  excess  alkenes  (propylene,  isobutylene^,  cyclohexene ,  butadiene 
and  ethyl  vinyl  ether)  at  20°  for  2k  hours  gave  the  spiro-lactone  structure  18. 

CYCLOBUTANE-1 , 3-DIONES 

An  examination  of  the  ultraviolet  spectrum  of  tetramethylcyclobutane-1,3- 
dione^  20,  showed  an  n  -»  it  transition  in  the  30C  mu  region.12  Kosowerls  observed 
a  second  n  ->  it  transition  in  the  ^>k0  mM-  region.  The  low  intensities  of  the 
transitions  as  well  as  both  the  direction  and  the  magnitude  of  the  solvent  ef° 
fects  are  consistent  with  n  -»  «*  transitions.   The  bands  are  observed  at  j>48  mM- 
(€  18)  and  308  mi-i  (6  39)  in  iso-octane;  3kk  nm  (6  18)  and  30^  mH  (6   30)  in 
ethanol.  The  two  different  transitions  are  explained  by  an  interaction  between 
the  carbonyl  groups  in  the  excited  state  and  are  described  as  follows; 

*+  *2*     (3^0  mu  band) 


A 
B 


n  ->  it- 


n  ->  m' 


Jtg 


(300  mu  band) 


-359- 
The  excited  state  A  may  be  pictured  as  shown  below. 

+ 
h  v  \ 


20  A 

The  photolysis  of  tetramethylcyclobutane-l,3-dione  in  the  vapor  phase  had 
been  carried  out  by  Turro,  Leermakers,  Wilson,  Necker,  Byers,  and  Vesley.14 
They  found  that  complete  photolysis  at  low  pressure  yields  2.0  moles  of  carbon 
monoxide ,  O.O78  moles  of  propylene,  0.0024  moles  of  methane,  and  a  trace  of 
propane  for  every  mole  of  the  dione.  There  is  a  slight  induction  period  in  the 
evolution  of  carbon  monoxide  and  a  rather  obvious  induction  period  for  propylene. 
This  difference,  and  also  the  low  yield  of  olefins  compared  to  the  production 
of  carbon  monoxide,  is  explained  by  the  early  formation  of  high  molecular  weight 
polymers  which  were  observed  on  the  sides  of  the  reaction  vessel.  The  photolysis 
was  carried  out  using  pyrex  vessels  and  mercury  arc  lamps.  These  conditions  were 
used  for  all  the  experiments  reported  in  the  remainder  of  the  seminar,  unless 
otherwise  noted. 

Several  groups  of  workers  have  looked  at  the  photolysis  of  the  dione  in 
inert  solvents.  Cookson,  %e,  and  Subrahmanyam15  have  irradiated  20  in  benzene 
solution.  They  obtained  tetramethylethylene  (80$  yield)  and  carbon  monoxide. 
Turro  and  co-workers16  observed  that  a  ketene  is  also  a  product  of  the  reaction. 
Evidence  for  the  presence  of  dimethylketene  is  suggested  by  the  yellow  color  of 
the  reaction  mixture,  the  strong  infrared  band  at  2124  cm  1 „  and  the  disappear^ 
ance  of  both  the  yellow  color  and  the  2124  cm"1  band  when  iso=propyl  alcohol 
was  added.  The  quantum  yield  for  the  disappearance  of  20  was  0. 38  +  0.01. 

The  formation  of  an  olefin  and  carbon  monoxide  seems  to  be  a  general  reaction 
for  tetra- substituted  cyclobutane~l,3~diones.  The  exhaustive  photolysis  of  di~ 
spiro[5.1.5<>3j-"fcetradecane~7>l4-dione,  21,  in  degassed  benzene  and  methylene 
chloride  yields  bicyclohexlidene  (70$  yitJ-d)  and  CO.14  If  a  benzene  solution 


h  v   v       <    >==/   \  +  2C0 


21 

of  tetramethyl-  and  tetraethylcyclobutane-l,>dione  is  photolyzed,  high  yields 
of  the  expected  tetramethyl-  and  tetraethyl-  olefins  are  obtained.,  but  no  di- 
methyldiethylethylene  could  be  detected  (limit  of  detection  about  0.1$).  The 
yellow  color  of  a  ketene  was  also  observed  in  the  latter  reaction. 

If  the  photolyzed  solution  of  20,  after  low  conversion,  is  analyzed  by 
vapor  phase  chromatography  (vpc),  the  presence  of  a  new  component  is  detected 
in  significant  amounts.  This  new  component  corresponds  to  isopropenyl  isopropyl 
ketone,  22,  and  after  isolation  by  vpc,  was  shown  to  be  identical  to  an  authentic 
sample.  However,  an  examination  of  the  reaction  mixture  by  spectral  methods  showed 
the  ketone  22  to  be  absent.  Therefore,  it  was  thought  that  the  ketone  22  was 
formed  as  a  result  of  decomposition  of  an  unknown  precursor  upon  analysis  by  vpc. 
As  the  photolysis  of  20  proceeded,  22  reached  a  steady  state  but  dropped  to  l/4 
if  the  reaction  mixture  was  allowed  to  stand  in  the  dark  for  15  hours. 

If  dione  20  is  photolyzed  in  benzene  under  540  mm  of  oxygen  and  the  reaction 
is  followed  by  mass  spectroscopic  and  vpc  studies,  the  products  shown  in  scheme 
3  are  observed  (with  appropriate  mole  ratios).17  No  isopropenyl  isopropyl  ketone22 
was  detected  by  vpc.  It  was  shown  that  both  the  ketone  22  and  tetramethylethylene 
were  stable  to  both  light  and  oxygen  under  the  reaction  conditions.  Therefore, 


hV, 


<t>H.  05 


=360» 
Scheme  3 

B 

CH3CCH3  +  CO  +  C02  + 
1.5  0.8    0.4 


0.07 


the  precursor  of  ketone  22  forms  tetramethylethylene  under  degassed  conditions 
and  acetone  and  tetramethylethylene  oxide  in  the  presence  of  oxygen,, 

Much  evidence  has  been  presented  that  the  precursor  of  isopropenyl  isopropyl 
ketone  is  tetramethylcyclopropanone,  2J5,  or  its  excited  state ,  for  which  many 
resonance  forms  can  be  drawn  as  shown  below.   Several  workers  have  been  successful 
in  trapping  the  cyclopropane ne  intermediate  and  finally  in  isolating  it  from  the 


0  0. 

Ac  <-*  jK  <—>  etc 


0 


< — > 


< — >  J/ 


reaction  mixture .  Richey,  Richey,  and  Clagett18  irradiated  tetramethylcyclo- 
butane-l,3=dione  in  5$  solutions  cf  ethanoi  and  were  able  to  isolate  tetramethyl- 
cyclopropanone  ethyl  hemiketal,  24 ,  in  35$  yield  by  recrystallization  and  55$ 
yield  by  vpc.  Also  detected  were  the  esters  2£  (20-25$),  26 ,  and  2J  and  ketone  28. 
The  last  three  were  found  in  10-15$  yield  and  were  believed  to  result  from 


OLOCH2CH3 

A- 


24 


> 


0 
II 
C-0CH2CH3 


26 


> 


8 

■C-OCH2CH3 
21 


> 


0CH2CH3 


28 


decomposition  of  the  hemiketal.   Isolation  and  the  resulting  decomposition  of 
the  pure  hemiketal  have  shown  this  assumption  to  be  correct.19 

Turro  and  co-workers14'16  obtained  corresponding  products  from  the  photol- 
ysis of  the  dione  in  methanol.  They  also  detected  a  5$  yield  of  tetramethyl- 
ethylene.  The  quantum  yield  for  the  disappearance  of  20  in  methanol  was  found 
to  be  0.49  +  0.03.   Cookson  and  co-workers15  carried  out  the  photolysis  in  the 

presence  of  furan  and  isolated  adduct  2£  in  15$  yield  after  distil- 
lation. Turro  and  co-workers20  were  able  to  isolate  tetramethyl- 
cyclopropanone, 2J>,  and  examine  its  properties.   Saturated  pentane 
solutions  of  the  dione  20  were  photolyzed  in  a  pyrex  Hanovia  450-w 
immersion  apparatus  at  ~^E°   for  1  to  2  hours.  Longer  reaction  times 
cut  down  the  yields  of  the  cyclopropanone.  The  resulting  solution 
was  stripped  of  the  tetramethylethylene  and  pentane  until  an  approx- 
imately 10$  solution  of  2J>  was  obtained  (estimated  by  infrared  spectroscopy). 
This  remaining  pentane  solution  of  2^  was  purified  by  bulb  to  bulb  distillation 
atJL  mm  and  20°.  A  band  for  the  C=0  stretching  frequency  was  found  at  1840 
cm  ,  which  would  be  expected  for  such  a  small  ring  carbonyl  compound.  They 
found  that  the  infrared  band  disappears  when  either  furan,  oxygen,  or  methanol 
is  added  to  the  pentane  solution.   When  injected  into  the  vpc,  2J  rearranges  to 
isopropenyl  isopropyl  ketone.   The  authors  concluded  that  tetramethylcycloprop- 
anone  "is  a  fairly  stable,  distillable  compound  which  can  be  handled  in  pentane 
solution."  Furthermore,  all  the  various  trapping  experiments  to  prove  the 


-361- 

intermediacy  of  2^  in  the  photolysis  of  the  dione  20,  are  valid.  In  fact,  all 
the  reactions  reported  for  tetramethylcyclopropanone  under  the  photolysis  con- 
ditions will  also  proceed  in  the  absence  of  light  starting  with  pentane  solutions 
of  the  cyclopropanone.   The  oxidation  of  2J>  in  the  presence  of  oxygen  is  thought 
to  proceed  through  the  peroxides  ^0  and  31  which  decompose  to  the  observed  products, 
acetone  and  tetramethylethylene  oxide. 14>XT 


=0  +  CO  /"\ 


> 


By  using  pyrex  glass  the  excitation  was  primarily  located  in  the  lowest 
energy  transition  of  the  dione  ( the  3^0  mu  region) .  The  use  of  filters  to 
excite  only  the  lowest  energy  transition  did  not  appear  to  alter  the  course 
of  the  reaction.14  No  fluorescence  or  phosphorescence  from  20  was  detected  at 
77°K  in  an  ethanol: ether: isopentane  matrix.  The  lack  of  emission  from  20  implies 
an  extremely  rapid  path  of  deactivation  from  the  lowest  excited  state.  There  is 
some  indirect  evidence  which  indicates  that  the  chemically  active  state  is  the 
n  ->  n*  singlet.  Since  the  quantum  yield  for  decomposition  of  20  is  high,  the 
most  likely  paths  for  deactivation  of  the  lowest  excited  state  are  (a)  photochem- 
ical reaction,  or  (b)  intersystem  crossing  to  the  lowest  triplet^  followed  by 
decomposition.   The  photolysis  of  the  dione  is  neither  sensitized  by  benzophenone 
(E  =  69  kcal  per  mole) nor  quenched  by  1,3-pentadiene  (0.3M).  These  facts  imply 

that  the  lowest  triplet  is  not  involved.  However 5  this  result  is  not  conclusive 
because  the  triplet  of  benzophenone  may  not  be  of  sufficient  energy  to  excite 
the  dione,  and  the  decomposition  of  dione  in  the  triplet  state  may  be  faster  than 
energy  transfer  to  the  quencher. 

In  order  to  decide  what  the  primary  photochemical  processes  in  the  photol- 
ysis of  tetramethylcyclobutane-l,3-dione  were,  Haller  and  Srinivasan2x  took  the 
infrared  spectrum  of  the  reaction  mixture  as  soon  as  possible  after  photolysis 
by  using  a  cell  which  allowed  direct  infrared  analysis  at  variable  temperatures. 
The  reaction  was  studied  at  room  temperature  and  in  a  nitrogen  matrix  at  k°K,   and 
conversion  of  20  to  products  was  held  to  10$  or  less  to  insure  that  no  secondary 
processes  would  take  place.  The  workers  were  able  to  follow  the  production  of 
tetramethylcyclopropanone  by  a  band  at  1840  cm"1  and  the  dime thy lketene  by  a 
band  at  2124  cm  1.  Both  bands  were  observed  at  both  temperatures  indicating  two 
primary  photochemical  processes  were  taking  place,  as  shown  in  processes  C  and  D. 


From  the  extinction  coefficient  of  dimethylketene  (at  2124  cm"1),  it  is 
estimated  that  this  product  accounts  for  at  least  20$  of  the  disappearance  of 
the  dione.  Concurring  results  also  were  obtained  by  an  ultraviolet  study. 
The  yield  of  esters  formed  when  alcohol  reacts  with  the  dimethylketene  has  already 


-362- 

been  shown  to  be  between  20-30$, 14?1S  which  agrees  with  the  physical,  data.  How- 
ever, process  D  severely  complicates  the  photochemistry  of  20  for  the  following 
reasons;  l)   If  dimethylketene  is  irradiated  at  2537  ft  in  cyclohexane  using  a 
vessel  with  KBr  windows,  tetramethylethylene ,  carbon  monoxide,  and  tetramethyl- 
cyclopropanone are  observed.   Since  conversion  was  kept  negligibly  small,  the 
cyclopropanone  must  come  only  from  dimethylketene.  The  reaction  can  be  explained 
in  the  following  manner.   The  photolysis  of  dimethylketene  in  the  vapor  phase 


>=«  ^-^         >; 


>;    +    >— c=c 


+  CO 


>   A    +  co 


yields  dimethyl  carbene  at  either  2537  or  366O  $.,  although  the  quantum  yield  is 
much  lower  when  the  higher  wavelength  radiation  is  used.22  An  analogy  for  the 
latter  reaction  is  the  known  procedure  for  making  cyclopropanones  from  diazometh- 
ane  and  ketenes.23  Therefore,  photolytic  decomposition  of  dimethylketene  may  also 
be  a  source  of  tetramethylethylene  and  tetramethylcyclopropanone  when  the  dione 
is  photolyzed  in  an  inert  solvent.  2)   If  oxygen  is  bubbled  through  a  solution 
of  dimethylketene  in  cyclohexane,  the  band  at  2124  cm"1  disappears  and  new  bands 
are  seen  at  2324  cm""1  (C02)  and  1720  cm""1  (acetone).   3)   Dimethylketene  is  known 
to  undergo  dimerization  -  xclusively   to  tetramethylcyclobutane-l,3-dione  at 
room  temperature.20-'24 

If  oxygen  is  added  to  a  partially  photolyzed  solution  of  the  dione,  the  ke- 
tene  disappears  quickly,  but  the  cyclopropanone  is  found  to  diminish  rather  slowly.21 
This  reaction  casts  doubt  on  the  efficiency  of  oxygen  as  a  trapping  agent  for 
the  cyclopropanone.  However,  Turro  and  co-workers20  claim  that  when  oxygen  is 
added  to  a  pentane  solution  of  isolated  tetramethylcyclopropanone,  it  disappears 
very  quickly.   No  apparent  reason  is  offered  for  this  conflict. 

By  assuming  that  the  extinction  coefficient  for  cyclopropanone  is  the  same 
as  for  cyclobutanone ,  Haller  and  Srinivasan21  estimate  that  the  primary  process 
C_  is  of  no  more  importance  than  process  D.  Therefore  since  there  are  2  moles 
of  dimethylketene  for  every  mole  of  tetramethylcyclopropanone,  primary  process 
D  accounts  for  no  more  than  40$  of  the  disappearance  of  the  dione. 

These  same  workers  found  still  a  third  primary  process  which  is  important.25 
This  process  involves  the  loss  of  2  moles  of  CO  to  form  t  etramethylethylene  dir- 
ectly. Evidence  for  this  process  was  a  band  at  II76  cm  1  in  the  infrared  spectrum 


< 


+  2C0    E 


of  the  photolyzed  dione,  at  low  conversion,  in  a  nitrogen  matrix.  This  band 
was  assigned  to  the  ethylene  derivative.  Better  evidence  for  process  E  was  pro- 
vided by  an  ultraviolet  study  of  a  cyclohexane  solution  of  the  dione  20.   The 
spectrum  showed  a  marked  increase  in  absorption  in  the  region  of  240  to  200  mu 
on  photolysis,  at  low  conversion,  in  a  quartz  cell.  The  contribution  to  this 
band  by  dimethylketene  was  determined  to  be  minor,  while  the  contribution  due  to 
the  cyclopropanone  derivative  was  even  less.  In  control  experiments  it  was  shown 
that  this  absorption  was  not  due  to  products  from  the  reaction  of  ketene  with  water 
or  oxygen. 

A  summary  of  the  reactions  taking  place  when  tetramethylcyclobutane-1,3- 
dione  is  photolyzed  is  shown  in  the  following  schemes 


.565- 


>=< 


+  2C0 


20 


> 


h-r 


=c=o 


+  CO 


It  is  difficult  to  estimate  the  extent  of  processes  C  and  E.  for  several 
reasons.  First,  the  secondary  processes  give  products  identical  to  those  formed 
in  the  primary  processes.   Second,  the  nature  of  the  process  which  converts 
tetramethylcyclopropanone  to  tetramethylethylene  is  unknown.  However,  direct 
photolysis  of  the  cyclopropanone  is  unlikely  at  low  conversion  of  20,  so  that  the 
presence  of  tetramethylethylene  in  the  early  stages  of  the  reaction,  requires 
the  third  primary  process.  As  reported  earlier  in  the  seminar,  trapping  experi- 
ments showed  5%  substituted  ethylene  present  when  the  cyclopropanone  was  trapped 
by  an  alcohol  and  in  the  presence  of  a  high  pressure  of  oxygen.  However,  as  it 
has  been  pointed  out,  oxygen  may  or  may  not  be  a  good  trapping  agent. 

It  has  been  shown  earlier  that  there  is  an  interaction  between  the  carbonyl 
carbon  atoms  in  the  excited  state  of  the  dione.13  This  effect  could  decrease 
the  distance  between  the  other  two  carbon  atoms  which  would  be  required  for  the 
simultaneous  loss  of  two  carbon  monoxide  molecules. 

Process  E^  according  to  Srinivasan,25  accounts  for  40$>  of  the  disappear- 
ance of  the  dione.  However,  Turro14  maintains  that  the  trapping  of  cyclopropanone 
with  oxygen,  furan  or  an  alcohol  is  complete,  and  this  trapping  process  accounts 
for  a  majority  of  the  disappearance  of  the  dione.   Process  E  accounts  for  only 
5$.   Srinivasan21  points  out  the  trapping  experiments  may  not  be  valid  for  the 
following  reasons: 

a)   Oxygen,  furan,  or  an  alcohol  may  react  with  an  excited  state  precursor  of 
tetramethylcyclopropanone,  which  otherwise  may  lead  to  other  products  such  as  the 
substituted  ethylene,  b)   A  solvent  effect  may  change  the  course  of  the  reaction 
in  a  more  polar  solvent. 

Prolonged  photolysis  of  tetramethylcyclobutane-l,3-dione  leads  to  still  another 
product,  the  lactone  jj2,  in  low  yields.20  Product  J52  might  be  thought  of  as  a 

dimerization  product  of  dimethylketene.  However,  the  thermal 
dimerizati  on  of  dimethylketene  leads  exclusively  to  the  tet- 
ramethylcyclobutane-l,3-dione.24  A  photochemical  reaction  of 
dimethylketene  under  the  same  conditions  of  the  photolysis 
reaction  of  the  dione,  did  not  lead  to  significant  amounts  of 
the  lactone.20  A  possible  mechanism  for  the  lactone  reaction 
is  shown  below.   Cookson  and  co-workers26  have  also  reported  a  low  yield  of  the 


hv" 


V- ? 


32 


lactone,  along  with  other  products,  in  the  photolysis  of  20. 

They  have  also  shown  that  this  lactone  formation  takes  place  in  other  cyclic 


=364- 
diones.  For  example,  when  2,2,4,4-tetramethyl~cyclohexane-l,3-dione  is  irradia- 
ted^ lactone  j>4  is  found  in  35-56$  yield. 


21  & 

CONCLUSION 

CyclolxitarDne  has  "been  found  to  undergo  most  reactions  expected  for  a  cyclic  ketone  j 
however,  in  alcohol  solvents ,  a  ring  expanded  acetal  product  has  been  observed.   This 
product  is  believed  to  result  from  a  carbene  intermediate.  An  oxocarbene  intermediate 
is  also  proposed  in  the  photolysis  of  benzocyclobutenediones  to  yield  ring  expanded 
acetals.  The  photolysis  of  cyclobutane~l,3-diones  has  been  found  to  go  by  three  dif- 
ferent paths.  The  secondary  processes  combine  to  make  the  understanding  of  the  various 
primary  processes  difficult. 

BIBLIOGRAPHY 

1.  R.  Srinivasan,  Advan.  Photochem. ,  1,  83  (1963). 

2.  R.  K.  KLemm,  D.  N.  Morrison,  P.  Gilderson,  and  A.  T.  Blades,  Can.  J.  Chem., 
i£,  193^  (1965). 

3.  R.  0.  Kan,  "Organic  Photochemistry,"  McGraw-Hill,  New  York,  N.  Y.  I9660 

4.  J.  E.  Starr  and  R.  H.  Eastman,  J.  Org.  Chem.  ,  ~%ko   !593  (19o6). 

5.  G.  Quinkert,  G.  Cimbollek,  and  G.  Buhr,  Tet.  Letters,  4573  (1966). 

6.  N.  J.  Turro  and  R.  M.  Southam,  Tet.  Letters,  545  (1967). 

7.  H.  U.  Hostellter,  Tet.  Letters,  687  (1965). 

8.  P.  Yates  and  L.  KLlmurry,  Tet.  Letters,  1739  (1964);  J.  Am.  Chem.  Soc,  88, 
I563  (1966). 

9.  R.  F.  G.  Brown  and  R.  K.  Solly,  Tet.  Letters,  169  (1966). 
10.  Ho  A.  Staab  and  J.  Ipaktschi,  Tet.  Letters,  583  (1966). 

12.  E.  A.  LaLancette  and  R.  E.  Benson,  J.  Am.  Chem.  Soc,  8^,  4867  (1961). 

1>  E.  M.  Kosower,  J.  Chem.  Phy. ,  ^8,  2813  (I963). 

l4.   N.  J.  Turro,  P.  A.  Leermakers,  H.  R.  Wilson,  D.  C.  Necker,  G.  W.  Byers,  and  G.  F. 

Vesley,  J.  Am.  Chem.  Soc,  87",  2613  (1965). 
15-  R.  C.  Cookson,  N.  J.  Nye,  and  G.  Subrahmanyam,  Proc  Chem.  Soc,  144  (1964). 
l6.  N.  J.  Turro,  G„  W.  Byers,  and  P.  A.  Leermakers,  J.  Am.  Chem.  Soc,  86,  955  (1964). 
17-  P«  A.  Leermakers,  G.  V.  Vesley,  N.  J.  Turro,  and  D.  C.  Neckers,  J.  Am.  Chem. 

Soc,  86,  4213  (i$6h). 

18.  H.  G.  Richey,  J.  M.  Richey,  and  D.  C.  Clagett,  J.  Am.  Chem.  Soc,  86,  3906  (1964). 

19.  N.  Jo  Turro,  W.  B.  Hammond,  P.  A.  Leermakers,  and  H.  T.  Thomas,  Chem.  and  Ind. 

990  (1965). 
20»  N.  J.  Turro,  W.  B.  Hammond,  and  P.  A.  Leermakers,  J.  Am.  Chem.  Soc,  87, 
2774  (1965). 

21.  I.  Haller  and  R.  Srinivasan,  J.  Am.  Chem.  Soc,  87,  1144  (1965). 

22.  R.  A.  Holroyd  and  F.  E.  Blacet,  J.  Am.  Chem.  Soc,  7£,  4830  (1957). 

23.  W.  B.  DeMore,  H.  0.  Pritchard,  and  N.  Davidson,  J.  Am.  Chem.  Soc,  8l,  5874 
(1959);  W.  B.  Hammond  and  N.  J.  Turro,  J.  Am.  Chem.  Soc,  88,  2880  JV)GG)  y 

N.  J.  Turro  and  W.  B.  Hammond,  J.  Am.  Chem.  Soc,  88,  3672  (I966). 
2K      W.  E.  Hanford  and  J.  C.  Sauer/Org.  Reaction,  ^,  127  (1946)  5  R.  H.  Hasek, 

Research  (London)   l4,  74  (l§6l). 
25-  I»  Haller  and  R.  Srinivasan,  Can.  J.  Chem.,  4^,  31^5  (1965). 
26.  R.  C.  Cookson,  A.  G.  Edwards,  J.  Hudec,  and  M.  KLngsland,  Chem.  Comm. ,  98  (1965). 


■z£c 

-pop- 


THE  MECHANISM  OF  PAPAIN  CATALYSIS 
Reported  by  Paul  Elliot  Bender 
INTRODUCTION* 


May  18,  196\ 


Papain  is  a  crystalline  proteolytic  enzyme  found  in  highest  concentrations  in 
the  secreted  fluid  of  latex  vessels  under  the  skin  of  the  tree  Caprica  papaya.1 
Named  in  1879  by  Wurtz  and  Bouchut,2  who  performed  the  first  veil  controlled 
experiments  on  the  crude  extract,  it  has  played  an  important  role  in  the  establish- 
ment of  many  basic  facets  in  our  present  conceptions  of  enzyme  action.1  The 
literature  concerning  papain  has  been  extensively  reviewed  to  1962.1-?3i'4"s  A  review 
of  the  work  on  primary  structure  appeared  in  19647  and  some  findings  of  more  recent 
studies,  (I965),  have  been  outlined.8  This  seminar  will  emphasize  the  material  of 
the  last  four  years  concerning  the  mechanism  of  the  papain  catalyzed  hydrolysis  of 
synthetic  substrates. 

ACYL  ENZYME  INTERMEDIATE 

The  first  direct  evidence  of  an  acyl  papain  derivative  was  the  observation  by 
Lowe  and  Williams,9,910  in  a  difference  spectrum,,  obtained  thirty  seconds  after 
adding  methyl  thionohippurate  shown  in  Figure  I  to  the  activated  buffered  enzyme, 


NH2 

I 

C-NH 
I 

NH 

(CH2)3 

NH—  CH 

BAEE 


-OCaH5 


NH2 

C=NH 

NH 

J    s 

(CH2)3 

NH-  CH 

BAA 


■NH; 


NH2 

I 
C^O 

I 
NH 

I 
(CH2)3 

:«cnh— CH 


I 


OCH. 


BCME 


NH2 

I 

(  QH2)  4 

C6H5CH20-.CNH—  CH 


■OCgHspNOa    CeasCKHCHa-C-OCHs 


Z-L-lysine-p-nitrophenyl  ester 


methyl 
thionohippurate 


Figure  I.   Structural  Formulas 


CsHsCNHCHaC-OCaHs 

ethyl  hippurate 


of  a  single  U.V.  absorption  band  at  X         313  rap,  (log  es4.3  +  C.3)  whose  O.D.  dropped 

iiislx  ■" 

max 

to  0  in  12.5  minutes.  From  a  comparison  to  the  model  compounds  shown  in  Table  I10 

below,  the  authors  suggested  that  the  group  most  comparable  in  terms  of  X         and 


Cbromophore 

ethyl  dithioacetate 
methyl  thionohippurate 


Table 

I 

Acyl  Enzyme 

Models 

1      C1) 

-"max 
(rap) 

log  e 
0  max 

x      (' 

^max 
(mu) 

305 

4.08 

460 

230 

»K  26 

291 

4.00 

2p0 

log  e 


max 


1.25 


4.00 


*In  this  paper  a-N~benzoyl~L-arginine  ethyl  ester,  a~N-benzcyl~L~argininamide,  a-N- 
benzoyl-L-citrulline  methyl  ester,  and  a-N-benzyloxycarbonyl  will  be  abbreviated  as 
BAEE,  BAA,  BCME  arid  Z,  respectively. 


-366- 

Table 
X_     W 
(mu) 

I  (Conto) 

log  e 
&    max 

358 

1.25 

278 

k.71 

(2) 
Ghromophore  Xmax"'  log  e^       X^  log  e^ 


(mu) 

=UHS  358  1.25  268  ito05 

papain  at  pH  7 

log  e,  whose  presence  in  the  intermediate  could  account  for  the  observed  difference 
max 

spectrum,  is  the  dithioaeylester  moietyD  It  is  quite  obvious  that  both  the  substrate 

and  the  enzyme  absorb  well  below  313  mu<>  Upon  adjustment  of  the  pH  from  the  initial 

60 0  to  2.5  (where  denaturation  is  known  to  occur)11  the  313  mu  band  shifted  to 

309  mu  and  the  absorption  was  maintained. 

The  spectral  evidence  for  the  second  acyl  enzyme ,  trans -cinnamoyl  papain #12>13 

also  had  to  be  obtained  via  a  difference  spectrum,  since  for  this  acyl  intermediate , 

the  rate  of  deacylation  is  comparable  to  that  of  acylation.  To  prepare  a  partially 

acylated  enzyme,  excess  trans -cinnamoyl  imidazole  was  added  to  a  pH  3«^3  buffered 

activated  papain  solution.  After  five  minutes  the  mixture  was  chromatographed  or  a 

Sephadex  G~25  column  to  isolate  the  acyl  enzyme.  The  difference  spectrum  of  a 

fraction,  so  obtained s   containing  the  acylated  enzyme  was  scanned  spectrophoto- 

metrically  from  39O  to  2^0  mu,  revealing  an  absorption  at  X      .  326  mu  (log  e*k.k2k) . 

Comparison  of  the  X         and  log  e    of  activated  native  trans -cinnamoyl  papain  to 

that  of  trans -cinnamoyl  cysteine  (the  thiol  ester  model)  and  N«acetylserinamide 

(the  hydroxyl  ester  model)  as  shown  in  Table  II13  below  indicates  X         of  the  thiol 
x     J         J  ■  max 

Table  II 

Acyl  Enzyme  Models 

trans -cinnamoyl  derivative  X   (mu)  log  e   (mu) 

papain  326  kek2k 

N-acetylserinamide  281.5  ^0385 

Cysteine  306  h0^k 

ester  model  to  be  closer  to  that  observed  for  the  acyl  enzyme.  Upon  denaturation  of 
trans -einnamoyl  papain  in  4.8M  guanidinium  chloride  X         shifts  to  3OI-309  mu  which, 

TUSLX. 

as  in  hippuryl  papain,  is  in  good  agreement  with  the  acylated  thiol  model  in  a  more 

uniform  nonperturbed  environment.  Indication  of  the  intermediacy  of  the  observed 

species  is  provided  by  the  following  evidences  (1)  A  complete  system  showed  an 

absorption  at  35^  mu  which  first  increased  to  a  maximum  and  then  decreased  to  zero. 

(2)  The  isolated  species  has  the  characteristic  extinction  of  the  trans -cinnamoyl 

moiety  and  a  X    which  differs  from  the  reactant  trans -cinnamoyl  .imidazole  (X 

max  J  v  max 

307  mu)  and  the  product  trans -cinnamate  ion  (X         269  mu) „  (3)  The  trans -cinnamoyl 

max 

moiety  is  not  separated  from  the  native  or  denatured  enzyme  on  Sephadex  filtration, 

implying  eovalent  bonding.  (V>  Addition  of  the  specific  substrate,  BAEE  (shown  in 

Figure  I) ,  to  the  trans -cinnamoyl  enzyme  showed  that  the  rate  of  deacylation  of 

the  trans -cinnamoyl  group  was  coincident  with  the  rate  at  which  papain  catalyzed 

hydrolysis  of  the  specific  substrate  reappears.14  In  the  light  of  the  kinetic 

evidence  for  BAEE  hydrolysis  involving  a  thiol  group s   (see  section  on  kinetics)  and 

the  implication  of  a  thiol  group  at  the  active  site,  Brubacker  and  Bender  concluded 

that  observation  (If)  implied  a  bonding  of  the  trans -cinnamoyl  moiety  at  the  same 

active  site,  (probably  the  thiol  group)  responsible  for  BAEE  hydrolysis. 

Also  indicative  of  a  eovalent  acyl  enzyme  intermediate  was  the  work  of  Kirsch 

and  Katchalski.15  The  investigators  compared  the  enzyme  catalyzed  to  hydroxide 

catalyzed  ratio  of  the  hydrolysis  rate  to  the  180  exchange  rate  of  acyl-thiol 


carbonyl  labeled  180»ethyl  hippurate  (shown  in  Figure  I).  Assaying  the  remaining 
ester  vs.  time,  they  found  a  hydroxide  ion  catalyzed  ratio  of  lks   but  the  enzyme 
catalyzed  ratio  of  80  was  within  experimental  error  of  no  detectable  exchange. 
The  authors  concluded  that  the  enzymatic  pathway  for  hydrolysis  either  passes 
through  an  acyl  enzyme  intermediate  or  through  an  enzyme  substrate  complex,  which 
hinders  180  exchange  sterically. 

The  further  finding  by  Brubacher  and  Bender13  that  in  the  deacylation  of 
trans -cinnamoyl  papain  in  the  presence  of  added  nucleophiles,  the  rate  of  nucleo- 
philic  catalysis  by  amines  was  much  greater  than  for  the  oxygen  analogues,  was 
taken  by  the  authors  to  implicate  a  thiol  ester  as  the  enzyme  intermediate. 

ET^Bffi  KINETICS 

Papain  has  a  broad  specificity  toward  amides  ,  and  wiH  hydrolyze  a  polypeptide 
more  completely  than  either  pepsin  or  trypsin.4  Kinetic  investigations  of  papain 
catalyzed  hydrolysis  have  relied  mainly  upon  synthetic  substrates  such  as  N-acyl 
a -amino  derivatives  of  esters  and  amides  of  a-amino  acids.  Kinetic  studies  on  these 
substrates  have  yielded  much  significant  information  as  to  the  mechanism  of 
catalysis. 

Since  Mf.chaelis-Menton  kinetic  parameters  have  been  employed  throughout  this 
paper,  they  are  derived  below  assuming  both  the  kinetic  path  shown  and  steady  state 
conditions. 

E  +  S  ^=*   ES  — k*  E  +  P 

0  *  ka(S)(E)   -  (k=a+  kb)(ES) 

where  (E)» 

of  free  enzyme 


Ejs      (E)  +  (ES)  where  (E)«  concentration 

o 


(Ec)s  enzyme  added 
initially 


dt      (S)  +  k„a  +  kb 


a. 


V    -4*  k 

definitions^-  k  .  «  k,  9  Ks  -^=^-r- — k.,  k  -  ES  dissociation  constant  - 

a     s 

Smith  and  coworkers  g,1  studying  the  pH  and  temperature  dependencies  of  BAEE 

hydrolysis  of  the  Michaelis  Menton  kinetic  parameters,  found  k^  ,/k  to  be  a  concave 

cat  m 

downward  bell  shaped  curve  as  a  function  of  pH.  The  limbs  of  the  BAEE  curve  appear 

to  represent  the  titration  of  two  prototropic  groups  in  the  enzyme  with  pKx«4.3  and 

pK2®8. 02  at  37°o  The  shift  in  the  descending  limb  with  temperature  indicated  a 

heat  of  ionization  of  5.1  kcal/mole  at  0°,  which  could  correspond  to  either  an  a- 

amino  or  a  thiol  group  in  the  opinion  of  the  authors.  They  suggest  the  latter 

alternative  due  to  the  proven  necessity  of  a  free  thiol  group  to  enzyme  activity. 

A  lack  of  significant  shift  in  pKx  with  temperature,  implicated ,  to  the  authors,  the 

titration  of  a  carboxylate  ion  other  than  an  a-carboxyl  group,  as  shown  by  the 

negligible  heat  of  ionization. 

Although  it  was  shown  that  positively  charged  substrates  are  hydrolyzed  more 

readily  than  neutral  ones,  and  that  negatively  charged  substrates  show  inhibition, 

the  shape  of  the  k  ./k  vs.  pH  curves  for  basic  BAEE,  BAA.  (shown  in  Figure  I),  and 

cat  m 

neutral  hippur amide  were  nearly  identicals  only  the  value  of  k  ,/K  (lim)  was 

greatly  different.  This  indicated  to  the  authors  that  acylation  proceeded  by  the 
same  mechanism  in  all  of  these  eases  (i.e.,  all  dependent  upon  two  prototropic 
groups).  A  plot  of  k    vs.  pH  for  BAEE  showed  this  rate  constant  to  be  pH 

independent  down  to  pH  5.0„  Below  this  value,  a  k  ,  decrease  was  seen,  which  was 

Cav 


-368- 

th ought  to  be  indicative  of  a  single  titratable  group  of  pKa  3° 5  at  25° 9   active  in 

the  decomposition  of  an  acyl  intermediate  to  products.  However,  correcting  Smith's 

pH-Stat  data  for  the  state  of  ionization  of  the  product  at  low  pH,  Sluyterman16 

found  no  pH  dependence .  Williams17  has  also  found  no  pH  dependence  in  the  hydrolysis 

of  methyl  hippur ate . 

In  a  study  of  papain  catalyzed  hydrolysis  of  a  series  of  hippurate  esters,18 

Lowe  and  Williams  found  that  nine  of  the  eleven  esters  studied  have  essentially  the 

same  value  of  k    (2-4  sec"1).,  Kirsch  and  Igelstronr19  found  similar  k  .  independence 
cat  cat 

of  leaving  group  in  a  series  of  carbobenzoxyglycine  esters „  To  account  for  this 

observation,  in  light  of  an  acyl  enzyme  intermediate ,  Lowe  and  Williams  proposed  the 

following  kinetic  scheme  (where  k2,  k3,  Vx   and  P2  are  the  acylation  rate  constant, 

deacylation  rate  constant,  an  alcohol,  and  hippuric  acid,  respectively). 


E  +  S 


~ 


ES 


assuming  steady  state  conditions 


dt 


ES! 

•f 

Pi 


k- 


■>     E    +    Ps 


(Eo)(S) 


L$LL 


ko, 


(k^-k3) 


(k..i+ka) 


definitions  s     k 


cat 


K 


m 


k  ,/K 

cat/  m 


k.k- 


(k»!+ka) 


in  which  k2  )>^>  k3  and  therefore  k 


k3„  According  to  these  kinetics,  if  k  ,/ 


cat' 


m 
with  p 


l0  A  plot  of  log  (k.  ,/K  )  vs.  a  gave  a  good  linear  fit 

c  ax  m 


k  ,/K  to  be  ^ 

cat'  m 


»/K  where  K 

*  s       s 


'cat  ■'  "3 
=  ka/K  ,  then  k2  «  k  ,  „  A  -plot  of  log  (k.^/K ..] 

!,  indicating,  according  to  the  authors, 

a  constant  and  therefore  k2  <^  k^:i. 

The  functional  dependence  of  the  deacylation  rate  constant  upon  pH  was  determined 
by  Brubacher  and  Bender13  using  the  isolated  trans -cinnamoylated  papain  aided  to  the 
desired  buffer  solution,  and  the  absorption  of  the  difference  spectrum  at  330  mu. 
followed  with  time0  The  authors  observed  the  deacylation  rate  constant  (k3)  to 
increase  in  a  sigmoid  fashion  with  pH  to  a  k3  (lira)  value  cf  3«68  x  10"30 
Deacylation  was  apparently  dependent  upon  the  ionization  of  a  single  protctropic 
group  of  pKa  ^.69,  which  the  authors  proposed  to  be  a  carboxyl  group,, 

In  a  comparison  of  papain  catalyzed  BAA  to  BAEE  hydrolysis,  Whitaker  and 
Bender20  spectrophotometrieally  reinvestigated  the  kinetics  of  hydrolysis,  extending 
the  pH  range  employed  by  Smith  and  coworkers1  and  analyzing  the  kinetic  data  in 
terms  of  the  rate  constants  and  prototropic  equilibria  illustrated  below  by  making 


EH2 

11  Ki 

EH  + 

K2 


kiClimt 
^  k_, 


E 


EH2S 
EHS 

J[k2 

ES 


EH2S: 


Mlim); 


Ki 

EHS5 


k3(lim.) 


^  EH  +  P; 


+  Pi 


certain  assumptions s  (1)  Kg  is  pH  independent  above  pH  5.0,  (2)  BAA  and  BAEE  share 
a  common  mechanistic  pathway  in  their  papain  catalyzed  hydrolysis,  (3)  k3  is 
independent  of  pH  at  high  pH,  and  (k)    steady  state  conditions.  The  authors  observed 

/Km   vs.,  pH  dependencies  of  BAA  and  BAEE  are  identical  in  form, 

are  very  different.  They  conclude 


that  although  the  k 
the  ] 


vs, 


cat/  m 
pH  dependencies  of  the  two  substrat- 


cat 

from  this,  that  granting  assumption  (2),  two  rate  steps  are  involved  in  these 
hydrolyses,  and  of  the  two  one  is  slower  in  amide  hydrolysis  whereas  the  other  is 
slower  In  ester  hydrolysis.  A  complete  analysis  in  terms  of  the  proposed  rate 


-369- 

constants  demonstrates  the  following;  (1)  a  sigmoid  dependence  of  k3  to  pH^  (2) 

the  pKi8  for  BAEE  deacylation  is  the  same  as  that  for  BAA  deacylation ^  which  in 
conjunction  with  (1)  supports  assumption  (3) 9   and  shows  the  dependence  of  deacylation 
upon  a  group  of  pKa  3.91^  (3)  k3  (lim)  for  both  BAA.  and  BAEE  are  equal  within 
experimental  error 9   which  is  to  be  expected  for  the  deacylation  of  a  common  aeyl 
moiety °9   (4)  k2  (lim)  for  BAEE  is  3D2  fold  greater  than  k3  (lim)§  (5)  k3  (lim)  for 
BAA  is  three  fold  greater  than  ks  (lim)  3  (6)  the  values  of  pKx  and  pK2  for  the 
aeylation  step  of  BAEE,  4.29  and  8.49  respectively,,  are  essentially  indentical  to 
those  for  BAA9  and  similar  to  the  values  obtained  earlier  by  Smith  and  coworkers.1 
Employing  these  values^  internal  consistency  of  the  data  to  the  proposed  scheme  of 
equation  (2)  was  shown  by  using  the  derived  values  of  the  limiting  rate  constants s 
prototropic  dissociation  constants  and  the  value  of  the  substrate  dissociation  con- 
stant to  calculate  k  ,/K  s   k  ,  and  K  vs.  pH  profiles  which  displayed  good 

C8iU   233.    C  £tXi        lu 

correspondence  to  the  empirical  data^  except  in  the  region  below  pH  4.5  io  the  K 
and  k  ,  vs.  pH  plots  for  BAA.  Bender  and  Brubacher21  offer  an  explanation  for 

these  discrepancies  as  perhaps  representing  either  an  increase  in  K  with  pH  (in 
this  range)  due  to  the  increased  repulsion  of  the  positively  charged  substrate  by 
the  progressively  more  positively  charged  enzyme 9   or  a  perturbation  of  the  enzyme  of 
such  a  nature  as  to  shift  the  pKx  of  the  enzyme  substrate  complex  below  the  pK2  of 
the  free  enzyme. 

In  conflict  with  the  bell  shaped  pH  dependencies  of  k  .  obtained  by  Whitaker 

and  Bender^20  Sluyterman26  and  Williams17  both  reported  that  k  ,  was  pH  independent 

for  ethyl  hippurate  hydrolysis  down  to  pH  4.2  and  3=8  respectively.  Lowe  and 
Williams31  suggest  that  the  pH  dependence  of  deacylation  in  benzoylargininyl  papain 
is  due  simply  to  binding  of  the  positively  charged  guanido  group  by  a  carboxylate 
ion ^  assisting  deacylation  by  orienting  the  thiol  ester  bond  and  perhaps  modifying 
the  conformation  of  the  acyl  enzyme.  They  note  that  although  the  k3  of  benzoyl- 
argininyl papain  is  7  times  greater  than  k  ,  of  ethyl  hippurate  (taken  as 

k,    ,  +.son)  at  pH  6.0^,  the  former  is  approximately  equal  to  the  latter  at  pH  3*0.? 

where  the  oarboxyl  group  is  largely  prctonated.  This  hypothesis  would  predict  that 
the  masking  of  the  basic  group  (e„g.5  with  a  N-formyl  group)  should  reduce  k^  . 

(lim)  to  the  area  of  2.7  sec"1  (the  k  .  of  ethyl  hippurate} .  The  recent  work  by 

Bender  and  Brubacher21  has  shown  that  for  Z-L-lysine  p-nitrophenyl  ester  (shewn  in 
Figure  I) s   the  value  of  k  +  (lim)  drops  from  45  sec"1  to  32  sec"1.  The  authors 

consider  these  two  values  to  be  the  same.  Furthermore^  substitution  of  BCME  (shown 
in  Figure  I)  for  BAEE,,  (which  provides  a  substrate  with  an  isosteric  acyl  function)  s 
has  been  recently  found  by  Cohen  and  Petra23  to  exhibit  no  effect  (20.2  vs.  20.15 
sec"1)  on  k3  (lim)  calculated  similarly.  The  latter  results  imply  that  the  presence 
of  a  positive  charge  on  the  substrate  does  not  increase  the  deacylation  rate. 

Kinetic  studies  by  Bender  and  Brubacher21  with  the  series  p-nitrophenyl^,  benzyl 
and  methyl  Z-L-lysine  esters  as  a  function  of  pH^,  analyzed  in  terms  of  the  scheme 
of  equation  (2)$   similar  to  the  treatment  of  Whitaker  and  Bender^20  have  followed 
similar  k  ,s   k^  ,/K  9   and  K  vs.  pH  dependencies  to  those  shown  by  BAEE.  For  this 

09X    CQ.X   HI         HI 

series  of  esters 3   k3  was  in  all  cases  at  least  3.5  fold  smaller  than  k2.  However^ 

it  was  found  that  K  decreased  from  a  value  of  10.7mM  for  the  methyl  ester  to 

approximately  G.Q^mM  for  the  p-nitr ©phenyl  ester s   a  3.2  x  102  fold  drop.  The 

authors  state  that  a  change  in  K.,  of  this  magnitude  strongly  implies  that  the 

substrate  leaving  group  is  bound  at  some  enzymatic  site.  This  dependence  of  Ks  upon 

the  structure  of  F^   is  in  conflict  with  the  conclusion  of  Lowe  and  Williams18  based 

upon  the  linear  dependence  of  log  (k  ,/K  )  vs.  6  for  four  aryl  hippurate  esters. 

cat  m 

MECHANISTIC  PROPOSALS 

Several  mechanisms  .have  been  proposed  in  the  course  of  papain  investigations  in 
order  to  explain  some  of  the  experimental  observations.  Most  of  those  proposed  agree 
that  an  acyl  enzyme  is  formed s   which  is  later  deacylated  by  various  means.  Lowe  and 


-370- 

Williams18  based  their  suggestion  that  the  k2  step  is  subject  to  nucleophilic 
catalysis  combined  with  acid  catalysis  on  the  observations  that  (1)  for  substituted 
aryl  hippurate  esters  log  (k^  ,/K  )  gives  a  better  fit  to  Q  than  to  o*~  (p^l.2) 

indicating  relative  insensitivity  in  k2  and  (2)  for  BAA  hydrolysis  the  data  of 

Smith  and  coworkers1  show  k  ,  ,TT  ^/k  ,,_  ~,  to  be  0.8,,  which  Lowe  and  Williams 

cat(H20,r  cat(D20j         ' 

consider  to  be  an  indication  of  acid  catalyzed  ester  hydrolysis .  It  is  their  view 

that  the  alkaline  limb  of  the  k  ./k  vs.  pH  curve  represents  not  thiols  but 

C3.X  xn 

imidazolium  titration  (for  histidine  pKa  -  5.6-7.0  at  25°^  AH  ~   6. 9-7. 5  kcal/mole) 0 
Their  proposal^  then^  is  that  the  thiol  group  is  aeylated  by  proton  abstraction  from 
a  sulfhydryl  group  by  a  carboxylate  ion  with  sulfur  attack  at  the  acyl  moiety  and 
proton  donation  to  the  leaving  group  by  imidazole .  It  should  be  noted  that  there 
are  only  two  histidine  residues  in  papain^  at  106  and  175 •  Further,,  the  direct 
evidence  for  imidazole  implication  (it  is  bound  by  earboxymethylation  with  irrevers- 
ible inhibition)2*  has  not  been  observed  by  other  workers55*26  who  have  found  that 
only  cysteine  is  bound  by  earboxymethylation. 

Whitaker  and  Bender20  employed  k  ,  as  a  measure  of  k2  in  the  hydrolysis  of 

BAA.  and  reported  k„  J\.   0  =  1°35<>  They  clearly  expected  a  larger  value  in 

accordance  with  general  base  catalysis 9   but  this  low  value  lends  no  support  to  such 
catalysis.  The  authors  suggest  that  the  carboxylate  ion  acts  as  a  base  and  the 
thiol  group  as  an  acid  which  is  aeylated  in  the  k2  step. 

A  mechanism  for  the  deacylation  reaction  offered  by  Smith1  involved  carboxylate 
attack  on  the  thiol  ester  intermediate.  Lowe  and  Williams22  tested  this  proposal 
for  nucleophic  carboxylate  catalysis  on  the  ester  by  comparing  intramolecularly 
catalyzed  hydrolysis  rates  and  activation  parameters  of  some  appropriate  models 
shown  in  Figure  II  with  the  k  .  and  activation  parameters  calculated  from  the 

temperature  dependency  of  k  .  (from  four  points)  for  methyl  hippurate  hydrolysis. 

The  rate  constants  for  the  hydrolysis  of  II  and  III  are  measured  from  the  pH 

S-fl-CH«  ^S-C-CHsHHCCqHs  /w    S' 5 

OH  ™2 


I  II  III 

Figure  II „  Models  of  the  acyl  enzyme 

independent  region  of  k  ,   vs.  pH.  The  best  model  of  hippuryl  papain  (II)  is 

ODS 

3  x  10°5  fold  slower  than  the  k  .  for  the  enzyme  catalyzed  hydrolysis.  The  other 

models  are  also  10~5  fold  slower.  Activation  parameter  comparison  for  III  shows 
that  the  entropy  of  activation  for  the  enzymic  k  ,  is  20  e.u./mole  lower  than  the 

model^  but  the  enthalpy  of  activation  is  nearly  7  kcal/mole  lower  for  the  enzymic 
reaction  then  the  model.  The  authors  conclude  that  the  large  difference  in  rate 
constants  suggests  that  nucleophilic  catalysis  of  deacylation  by  a  carboxylate  ion 
is  unlikely. 

Lowe  and  Williams31  proposed  that  deacylation  is  catalyzed  by  nucleophilic 
imidazole  catalysis  at  the  thiol  ester  followed  by  a  rapid  hydrolysis  step.  Kinetic 
support  for  imidazole  participation  in  deacylation  has  come  only  from  the  results  of 
Cohen  and  Petra23  on  a~N-benzoyl<=L-eitrulline  methyl  ester  s   where  a  sigmoid  dependence 
of  k3  upon  pH  with  a  pKa  of  7  was  calculated.  The  form  of  this  dependence  implies 
the  catalytic  group  is  active  in  the  protonated  form. 

The  observation  of  large  deuterium  isotope  effects  by  Whitaker  and  Bender g20 


sat, 


/K 


at. 


-371- 
2.V?  for  BAEE  hydrolysis ,  and  Brubaeher  and  Bender ,12*2S 


"H20  ^T^O 
k„  -Vk-  0  ss  3„35  for  deacylation  of  trans  -cinnamoyl  papain,  all  support  general 

base  catalysis  on  the  thiol  ester  by  a.  carboxylate  ion  as  the  mechanism  of  the 
deacylation  step  by  implying  proton  transfer  in  the  transition  state  of  this  step. 
The  complicity  of  a  carboxylate  ion  to  deacylation  has  been  discussed  above „  In  a 
study  of  the  deacylation  of  trans -cinnamoyl  papain  by  added  amine  nucleophiles  i 


Brubaeher  and  Bender13  observed  that  a  plot  of  log  k 


amine  eat0  deacylation 


\     i.  ,,  ■."  P 


log  k*)  vs.  pKa  of  the  amine  produced  no  correlation  to  any  simple  relationship.  A 
plot  of  the  observed  deacylation  rate  constant  k    =  k3  +  k^  (added  nucleophile)  vs. 

the  added  nucleophile  concentration  gave  a  straight  line  with  no  evidence  of  binding 
of  the  nucleophile.  The  authors ,  however,  felt  that  certain  comparisons  of  k&  for 
appropriate  nucleophiles  of  simlliar  pKa  or  structure  necessitated  that  a  specific 
binding  interaction  exists,  and  that  basicity  has  little ,  if  any,  influence  upon 
k^.  The  authors  considered  this  supportive  evidence  for  general  base  catalysis  in 
the  deacylation  of  trans -cinnamoyl  papain.  They  reasoned  that  the  less  basic  the 
nucleophile  ,  the  more  readily  it  will,  release  its  proton,  but  the  more  basic  the 
nucleophile,  the  better ,  so  that  these  two  effects  would  tend  to  cancel.  In  regard 
to  the  general  applicability  of  these  observations,  it  must  be  noted  that,  both  the 
added  nucleophile  effects  and  the  largest  observed  deacylation  isotope  effects 
were  observed  in  trans =cinnamoyl  papain  hydrolysis,  where  k3  (lim)  -   3»68  x  10~3 
sec"1,  or  700  fold  smaller  than  the  k  +  for  methyl  hlppurate  hydrolysis,  and  about 

10,000  fold  smaller  than  N~benzoylargininyl  papain  deacylation.  Lake  and  Lowe27 
have  interpreted  this  as  indicative  of  either  the  involvement  of  a  different  rate 
determining  step,  or  the  employment  of  another  mechanism  In  the  hydrolysis  of  this 
nonspecific  trans -cinnamoyl  papain.  They  have  also  found  k^  rJ^-s^.   0  for  p- 

nitrophenyl  hlppurate  to  be  1„75  which  they  have  interpreted  as  a  secondary  isotope 
effect. 

A  fourth  deacylation  pathway  has  been  recently  proposed  by  Lake  and  Lowe27  as 
involving  a  slow  conformational  change  from  the  acyl  thiol  enzyme  (ES3)  to  a  con- 
former  (ES!J),  in  comparison  to  the  relatively  rapid  subsequent  breakdown  step 
giving  product.  To  test  the  pathway  involved  in  general  base  catalysis  (Scheme  I) 
against  this  fourth  proposal  ( Scheme  II) ,  the  authors  studied  the  effect  of  added 
methanol  upon  k  .  for  the  production  of  Px  from  p-nitrophenyl  hlppurate,  and  also 

for  the  production  of  hippuric  acid  (P^)  from  methyl  hlppurate.  The  k  .  for 

CQ.X 

product  formation,  according  to  each  scheme,  is  the  following %     for  Px  in  Scheme  I 


Scheme  I   E  +  S 


ES 


E  +  P2 


E  +  P3 


Scheme  II  E  +  S 


ES 


ES! 
+Pi 


Jia 


»  ES3 


E  +  P2 


E  +  P. 


3 


(assuming  k2  »  k3  +  k^  (Me  OH))  k  .  s  k3»  +  k4  (Me  OH),  in  Scheme  II  (assuming  k2  » 


k3  and  k3  «  14  +  k5  (MeCH))  k 


cat 


cat 


k3£  for  P2  in  Scheme  I  (assuming  k2  )>  k:3  +  k^ 


(Me OH))  l/k  .  ■  l/k3,  and  in  Scheme  II  (assuming  k3  «  k|  +  k5  (MeCH))  l/k. 


cat 


(k2  +  kaJ/kaks  +  (ks  +  k3)(k5  (MeOHj/ksksk^.  It  was  observed  for  p~nitrophenal  (Px) 
that  k  ,  was  independent  of  the  methanol  concentration  up  to  2M  methanol,  and 


-372- 
decreased  slightly  above  this  point.  Following  hippuric  acid. production,  . 

was  seen  to  be  linearly  related  to  the  methanol  concentration.  These  results  are 
in  agreement  with  any  pathway  of  the  kinetic  form  of  Scheme  II  with  k3  <(<(  k!4  +  k5 
(Me OH) ,  but  do  not  provide  a  description  of  the  k3  stepc  Lake  and  Lowe  state  that 
these  observations  provide  strong  support  for  the  absence  of  general  base  catalysis 
in  the  deacylation  of  hippuryl  papain,, 

SUMMARY 

A  large  collection  of  experimental  evidence  has  accumulated  to  specify  certain 
aspects  of  the  pathway  of  papain  catalyzed  hydrolysis  of  synthetic  substrates. 
Although  the  evidence  supporting  an  acyl  thio-enzyme  is  significant,  the  detailed 
mechanism  of  its  formation  and  deacylation  is  subject  to  the  uncertainty  of  con- 
flicting observations  and  interpretations.  Particularly  evident,  has  been  the 
dependence  of  the  observations  and  hence  the  proposed  mechanism,  upon  the  substrate 
employed.  It  would  appear  that  to  make  any  generalized  statement  as  to  the  nature 
of  papain  catalyzed  hydrolysis  would  at  this  point  be  premature  due  to  a  lack  of 
knowledge  concerning  possible  variations  or  discontinuities  in  the  mechanism  with 
a  spectrum  of  substrates, 

BIBLIOGRAPHY 

1,  E,  Smith  and  J.  Kimmel  in  "The  Enzymes,"  Vol,  4,  P,  Boyer,  H,  Lardy,  and  K, 
Myrback,  Ed,,  Academic  Press  Inc.,  New  York,  N.Y. ,  I960, 

2,  A,  Wurtz  and  E,  Bouchut,  Compt,  Rend,  Acad,  Sei„,  8g,  425  (l8?9)o 

3,  K,  Hwang  and  A,  Ivy,  Ann,  New  York  Acad,  Sci,,  j4,  l,6l  (1951). 
k,     J.  Kimmel  and  E,  Smith,  Advances  in  Enzymol,,  1£,  267  (1957). 

5„  E.  Smith,  R,  Hill  and  J.  Kimmel  in  "Symposium  on  Protein  Structure,"  A, 
Neuberger,  Ed,,  Methuen,  London,  1958. 

6.  E„  Smith,  A,  Light,  and  J,  Kimmel,  Symp.  Biochem,  Soc,  21,  88  (1962). 

7.  A.  Light,  R,  Prater,  J.  Kimmel,  and  E„  Smith,  Proc.  Natl,  Acad.  Sci.  U.S., 
£2,  1276  (1964) . 

8.  M.  Bender  and  F„  Kezdy,  Ann,  Rev.  Biochem,,  Vol.  34  (1965). 
9»  G„  Lowe  and  A.  Williams,  Proc.  Chem.  Soc.,  140  (1964). 

10.  G.  Lowe  and  A.  Williams,  Biochem.  J.,  96,  189  (I965), 

11.  A.  Glazer  and  E.  Smith,  J.  Biol.  Chem.,  2^5,  I9V5  (I96I). 

12.  M.  Bender  and  L,  Brubacher,  J.  Am.  Ghem.  Soc,,  86,  5333  (1964), 

13.  L,  Brubacher  and  M,  Bender,  J.  Am.  Ghem.  Soc,  B5,  5871  (1966). 

14.  M.  Bender,  et.  al. ,  J.  Am.  Ghem.  Soc.,  88,  5890T1966) . 

15.  Jo  Kirsch  and  E.  Katchalski,  Biochem,,  ¥J  884  (1965)0 

16.  L„  Sluyterman,  Biochim,  Biophys.  Acta,  B^,  305  (1964). 

17.  A.  Williams,  Doctoral  Thesis,  Oxford  University,  1964. 

18.  G,  Lowe  and  A,  Williams,  Biochem,  J,,  96,  199  (1965), 
19o  Jo  Kirsch  and  M.  Igelstrom,  Biochem.,  5,  7^3  (1966). 

20.  J,  Whitaker  and  M.  Bender,  J.  Am.  Chem?  Soc,  87,  2728  (1965). 

21.  M.  Bender  and  L.  Brubacher,  J.  Am.  Chem.  Soe.,~|8,  5880  (1966). 

22.  G,  Lowe  and.  A,  Williams,  Biochem.  J.,  96,  194  (1965). 

23.  W.  Cohen  and  P.  Petra,  Biochem.,  6,  104f  (1967). 

24.  S„  Yu~Kum  and  T.  Chen~Lu,  Sci.  Sinica,  12,  1845  (1963). 

25.  B„  Finkle  and  E.  Smith,  J.  Biol.  Chem.,  2J0,  669  (1958). 

26.  A,  Light,  Biochem.  Biophys.  Res.  Common.,  if,  781  (1964). 

27.  A.  Lake  and  G„  Lowe,  Biochem,  J.,  101,  402  (I966). 


-373= 

THE  PHOTOSENSITIZED  CIS-TRANS  ISOMERIZATIOli  OF  OLEFINS 

Reported  by  Robert  Kalish  May  22 ,  196' 

The  light  induced  c is ° trans  isomerization  of  olefins  has  long  been  known,  and 
has  been  widely  used  as  a  synthetic  tool;,  especially  for  the  preparation  of  the 
less  stable  member  of  an  isomeric  pair  of  olefins.,  It  has  only  been  within  recent 
years,  however,  that  the  mechanisms  of  these  isomerizations,  particularly  those 
occurring  under  triplet  photosensitized  (hereafter  referred  to  as  photosensitized) 
conditions ,  rave  been  intensively  studied.  This  seminar  will  be  specifically 
concerned  with  mechanistic  aspects  of  the  photosensitized  cis-trans  isomerization 
of  olefins  occurring  in  solution,  the  area  of  vapor  phase  photosensitized  iso- 
merizations  having  recently  been  reviewed.1  The  utility  of  such  studies  in  en- 
hancing understanding  of  photochemical  processes  in  general,  and  triplet  energy 
transfer2  in  particular,  will  also  be  emphasized . 

NON-CONJUGATED  OLEFINS 

Theoretical  calculations1*3  indicate  the  most  stable  configuration  of  the 
lowest  triplet  state  of  ethylene  and  other  non-conjugated  olefins  to  be  the 
twisted,  non-spectroscopic  form  (the  so-called  phantom  triplet,9  p)  in  which  the 
planes  defined  by  the  two  GHp  groups  are  orthogonal  thus  minimizing  interaction 
between  the  two  2Cp  orbitals  and  electrons. 

Such  a  twisted  triplet  state  provides  a  ready  pathway  for  the  cis -trans 
isomerization  of  simple  olefins,  as  formation  of  this  triplet  from  either  a  cis 
or  trans  olefin  is  expected  to  result  in  decay  (via  twisting  and  inter system 
crossing)  to  both  the  cis  and  trans  isomers.  The  existence  of  a  twisted  triplet 
as  a  common  intermediate  in  these  isomerizations  is  based  both  on  theoretical 
predictions, ln3i"4  and  on  the  finding  that  the  sum  of  the  quantum  yields  for  the 
cis  to  trans  (<t>+  )  and  trans  to  cjLs  (<J>  )  photosensitized  isomerization  of  many 

olefins,  in  which  the  olefin  triplet  is  formed  with  unit  quantum  efficiency,  is 
one.5  Calculations**  indicate  that  4>   +  $   equals  one  only  if  isomerization 

occurs  solely  from  a  common  twisted  triplet.  If.  however,  isomerization  occurs 
from  two  non-interconvertible  spectroscopic  triplets,  0\$  ,  +  <1>  <^2.  The  fact 

that  4>   +  <t>   has  been  found  to  equal  but  never  to  exceed  one  is  strong  pre- 
sumptive evidence  for  a  common-intermediate  process. 

Two  types  of  triplet  energy  transfer  processes.,  vertical  and  non-vertical, 
can,  in  principle,  occur  via  a  coupled  energy  transfer  process2  of  the  form  shown 

ml  C<) 

in  eq  1,  S   being  the  triplet  energy  donor  ( sensitizer) ,  and  A   being  the 
(olefin)  acceptor.  Vertical  (classical)  energy  transfer  occurs  in  accordance  with 

ml      qO         qO      ml 

(1)  S~  +  Ab— ►  SS  +  AT 

the  Franck~Condon  principle,7  producing  vibrationally  excited,  spectroscopic 

olefin  triplets  which  then  relax  vibrationally  (via  twisting)  to  the  twisted 

triplet  with  subsequent  decay  to  the  isomeric  ground  state  olefins.  Non-vertical 

(non-classical)  energy  transfer4,5'3'99  directly  to  the  twisted  triplet,  although 

forbidden  by  the  Franck-Condon  principle  for  radiative  processes,  can  occur  in 

a  coupled  triplet  energy  transfer  process.  Vertical  and  non-vertical  energy 

transfer  are  competitive  processes,  the  less  probable  and  hence  less  efficient 

non-vertical  process  being  unimportant  relative  to  the  vertical  process  when  the 

S 
triplet  energy  of  the  sensitizer,  E  (measured  by  the  0»Q  band)  exceeds  the 

triplet  energy  of  the  acceptor,  E2,  by  ^3  or  mere  kcal/mole.  In  such  a  case 

vertical  energy  transfer  is  diffusion-controlled,2  occurring  on  every  collision. 

When  E  is  less  than  E2  the  classical  energy  transfer  process  is  endothermic. 


Non-vertical  energy  transfer  can  now  effectively  camp    &th  the  vertical  process 
which,  although  intrinsically  more  efficient,  has  a  higher  activation  energy 
than  the  non-vertical  process,,  When  energy  transfer  is  endothermic  and  hence 
not  diffusion- controlled,  the  time  required  for  this  process  to  occur  has  been 
calculated  to  be  at  least  10  8  sec    It  is  thus  not  surprising  that,  the  lower 
energy  non~ vertical  transitions  can  occur,  as  the  time  required  for  radiative 
transitions j,  for  which  the  Franck-Condon  principle  was  formulated,  is  about  10~15 
sec  Little  is  known  about  the  detailed  nature  of  non-vertical  processes ,  however, 
which  will  be  discussed  further  in  connection  with  the  c is -trans  isomerization 
of  the  stilbeneso 

The  acetone  sensitized  cis^trans  isomerization  of  2-pentene  has  been  studied 
in  detail  in  a  variety  of  solvents  by  Borkman  and  Kearnso10  These  workers  found 
that  for  concentrations  of  2-pentene  greater  than  1„0  M,  <t>  .  +  $ .   equalled  loO 

+  Ool  in  accord  with  isomerization  via  a  common 9 twisted  triplet  intermediate. 
The  quantum  yield  for  energy  transfer  from  acetone  to  2-pentene  was  determined 
to  be  loO  +  Ool  indicating  that  every  acetone  triplet  eventually  transfers  its 
excitation  energy  to  a  2-pentene  molecule |  this  finding  implies  nothing  about  the 
efficiency  of  energy  transfer  during  a  single  collision,  however,, 

It  was  found  that  although  2-pentene  completely  quenches  the  phosphorescence 
of  acetone,  it  does  not  affect  the  fluorescence  of  acetone  under  the  same  conditions 
in  which  the  acetone  sensitized  2-pentene  isomerization  occurs  with  100$  efficiency,, 
Indicating  that  energy  transfer  from  acetone  to  2-pentene  is  not  vibrational  in 
nature  and  proceeds  from  the  triplet  state  of  acetone  (ivlCf  6  sec)  rather  than 
from  the  shorter-lived  singlet  state  (t^2«5  X  10"d  sec) 9   a  fact  which  has  long  been 
assumed  in  ketone  photosensitized  olefin  isomerizations,  but  never  before  rigor- 
ously proven.  Although  attempts  to  detect  2-pentene  triplets  spectroscopically 
(e,g0 „  by  esr  in  rigid  glasses  at  77°  K)  under  conditions  of  the  isomerization 
were  unsuccessful,  perhaps  owing  to  a  short  triplet  lifetime,  it  is  quite  likely 
that,,  in  accord  with  the  Wigner  spin  conservation  rule,  it  is  the  2-pentene  trip- 
let which  Is  formed  upon  energy  transfer „ 

From  a  study  of  the  initial  rate  of  the  isomerization  as  a  function  of  the  e©n- 
centration  of  2-pentene,,  Borkman  and  Kearns  were  able  to  calculate  the  quenching 
constant^,  K  ,  for  the  triplet  energy  transfer  proce.s    tere  K.  =^r_k.  „  T  being 

the  acetone  triplet  lifetime  in  solution  and  k  being  the  overall  bimolecular 


rate  constant  for  energy  transfer  from  the  acetone  triplet  to  2-pentene,,  From 

the  experimentally  determined  value  of  K  ,  k  was  calculated  to  be  ^lO7  m"1  sec"1, 

a  value  about  103  times  less  than  the  theoretically  predicted  diffusion-controlled  one, 

This  finding  led  Borkman  and  Kearns  to  conclude  that  the  energy  transfer  step, 

known  to  involve  close  contact  between  sensitizer  triplet  and  acceptor,2  is  best 

T1         $r 
written  as  two  distinct  steps  (eq  2),  [S   *  « olefin  ]  being  a  collision  complex 

£l  cD       &       ml  gO      k         O  ml 

(2)      S   +  olefin0  f=±=±[SL  ..«,  olefin  ]     >  S*  +  olefin 


in  which  2-pentene  is  adjacent  to  the  acetone  triplet  but  in  which  energy  transfer 
has  not  yet  occurred $  this  collision  complex  can  revert  back  to  2-pentene  without 
energy  transfer  (k,  )  or  undergo  dissociation  with  concurrent  energy  transfer  (k 


c 


Kinetic  analysis  of  eq  2  Indicates  that  k,=k  k  /(k  +  k,  )»  If  k  ,  the  nearest 
tf       *  t  a  c   c    d       c 

neighbor  rate  of  energy  transfer,  is  large  compared  with  k,  then  k. -k  ,  i.e., 

d      t  a 

the  process  is  diffusion-controlled,,  If  k  <C<(k,  then  k, -k  k  /k,  and  k.  is  less 

than  diffusion-controlled  as  found  for  2-pentene «  Furthermore  if  it  Is  assumed 
that  the  collision  complex  has  negligible  stability,  l<,e«,  3   £H°  of  formation  *^-0_9 


-375= 

then  k  /k    should  be  temperature  independent  indicating  that  k  will  he  a  function 
a  d  ' 

of  temperature  only  if  k  is0  Variable -temperature  kinetic  studi.es  in  the  25° 

to  =78°  range  did  indicate  a  temperature  dependence  in  k.  (k  )  from  which  an 

™  t   c 

activation  energy  of  ^  ka 3  kcal/mole  was  calculated  for  the  energy  transfer  stepD 

The  E  of  acetone  has  been  estimated  from  phosphorescence  studies  to  be 
^v^80  kcal/mole10  (a  value  disputed  by  Cundall11  who  claims  a  value  of  ^75  kcal/ 
mole) ,  whereas  the  0-0  triplet  energy  of  2-pentene  has  been  taken  by  Kearns  to 
be  equal  to  that  of  ethylene , 4 /v 82  kcal/mole  (a  value  which  may  be  slightly 
too  high1^  o  Triplet  energy  transfer  from  acetone  to  2-pentene  is  thus  predicted 
to  be  endothermic  by  ™  2  kcal/mole v  in  reasonably  good  agreement  with  the  exper- 
imentally determined  activation  energy,  thus  accounting  for  the  observed  rate  of 
energy  transfer „ 

Morrison,  et„  al„  ,  have  used  the  photosensitized  cis-trang  isomerization 
of  olefins  as  a  tool  to  investigate  intramolecular  triplet  energy  transfer,12 
i0eo  y  light  absorbed  in  one  part  of  a  molecule  results  in  a  chemical  reaction 
at  another,  non-conjugated  part  of  the  same  molecule  via  energy  transfer  from  the 
initially  excited  chromophore  to  the  reacting  center » 

The  most  conclusive  evidence  for  intramolecular  triplet  energy  transfer  comes 
from  a  study  of  the  irradiation  of  trans-  and  cj^-l-phenyl~2~huter.ie,12  "  light 
of  230-280,^1  being  used  to  insure  that  only  the  phenyl  chromophore  is  excited 
to  the  jc^rt  triplet  state;  the  only  observable  reaction  was  cis-trang  isomer- 
ization with  <&n4.=$,    =0o21  +  O0OI60  The  mechanism  suggested  by  Morrison  for  this 


photoisomerization  is  given  in  eqs  3=60 12  s 


It  is  assumed  that  the  phenyl  donor 


S°   hv 


a 


k 


„0    &3        -K-4     rO 

(6)      tfa  « p  — VcD 

and  double  bond  acceptor  chromophore s  are  non-interacting  in  the  ground  and  sing- 
let excited  states,  and  also  that  the  rates  of  intersystem  crossing  (k_._)  and  rad- 

iationless  decay  (kj  of  the  phenyl  donor  are  the  same  for  either  isomer;  kx 

and  k2  are  the  rates  of  energy  transfer  from  the  phenyl  triplet  to  the  cis  and  trans 

olefinic  linkage,  respectively,,  giving  a  common,  twisted  triplet,  p0  Since  E_6  5" 

(  ^83  kcal/mole)  is  at  least  2«3  kcal/mole  greater  than  the  0-0  triplet  excitation 
energy  of  either  isomer,  ki=k2  and  energy  transfer  probably  occurs  at  close  to 
the  diffusion-controlled  rate  to  give  vibrationally  excited,  spectroscopic 
olefinic  triplets  which  decay  to  p  which  undergoes  subsequent  decay  with  concom- 
itant cis-trans  isomerization  (k3.k4)0 

Kinetic  analysis  of  eqs  3^6  leads  to  the  prediction  that  the  photo stationary 
state  composition,  ( cis) _/trans}  ,  is  equal  to  (k2/kj)^fk4/k3) „  Since  kx=k2, 

(SjLOL)   AiSH£§J  "^4/^3,?  ii£ej  "k^e  photo  stationary  state  composition  is  determined 

by  the  decay  ratio  of  p,  as.  intrinsic  property  of  the  excited  state  in  a  given 
solvent  at  a  given  temperature.,16  Moreover,  this  expression  predicts  that  ex- 
citation of  the  double  bond  by  intermolecuiar  energy  transfer  from  benzene  should 
lead  to  the  same  (cis) ./(trans)  as  the  intramolecular  process;  in  accordance 

with  this  prediction,  (cis)  /(trans)  was  found  to  equal  1„0  for  both  types  of 

sensitization  processes. 


Evidence  that  the  observed  intramolecular  energy  transfer  is  electronic  and 
not  vibrational.,  i.e, .,  from  a  vibrationally  excited  ground  state  of  the  phenyl 
group  leading  to  thermal  cis -trans  isomerlzatloc.,,  comes  from  the  observation  that 
the  quantum  yield  for  the  trans  to  cis  isomerization  in  cyclopentane  is  reduced 
from  0o21  to  0.15  in  the  presence  of  an  equimolar  amount  of  the  triplet  quencher 
trans °2°hexene ,  indirectly  indicating  that  intramolecular  transfer  of  electronic 
energy  does  occur o 

Cis-trans  isomerizations  have  also  been  ob served,  to  occur  upon  irradiation 
of  trans-4°hexen"2"One12  and  tranS"5~hepten=2"Oneo12   Formation  of  the  cis 
isomer  upon  excitation  of  the  carbonyl  group  of  trans =S-hepten-2~one  was  taken 
by  Morrison  as  fairly  good  evidence  for  intramolecular'  energy  transfer  from  the 
carbonyl  triplet  to  the  double  bond,  although  the  possibility  of  vibrational 
energy  transfer  was  not  ruled  out.  The  situation  with  regard  to  the  isomerization 
of  trans~4"hexen=2"One  is  unclear,,  products  arising  from  acetyl  and  2-butenyl 
radicals  being  observed. 

Recent  studies  of  the  benzene,,  toluene  and  xylene  sensitized  irradiation 
of  a  series  of  l~alkylcycloalkenes  have  indicated  the  possible  formation  of  the 
highly  strained  trans  isomers,.13  Irradiation  of  (+)-3~carene  (l)  was  found  to 
give  (+)"3(10)=-carene  (2)  |  in  the  presence  of  methanol,  ethers  jj  and  4  were  also 
formed  (eq  ?) •  Deuterium  labeling  studies  indicated  the  rearrangement  to  the 


h*v   v 
"7" 


CH30H, xylene 


-OCH3 
(12#) 


exocyciic  olefin  to  be  intermolecular  with  respect  to  the  proton  shift « 

These  observations  of  proton  incorporation,,  ether  formation^  rearrangement 
and  Markovnikov  addition  are  suggestive  of  ionic  processes s   and  led  Kropp  to 
suggest  that  decay  of  the  twisted  cycloalkene  triplet^  formed  via  triplet  energy 
transfer o  leads  to  both  cis  and  trans  olefin,  the  highly  strained  trans  isomer 
undergoing  protonation  with  relief  of  strain  to  give  a  carbonium  ion  which  leads 
to  the  observed  products,,13   The  presence  of  olefin  triplets  in  these  reactions 
was  suggested  by  the  finding,  of  Carroll  and  Marshall  that  the  reaction  rate  is 
decreased  by  added  oxygen..     'These  workers  propose  a  mechanism  involving  direct 
protonation  of  the  olefin  triplet  to  give  a  carbonium  ion  and  the  observed  products » 
Kropp  believes  this  mechanism  to  be  less  likely  in  view  of  the  finding  that 
exocyciic  and  acyclic  olefins 3   the  trans  isomers  of  which  are  not  highly  strained,, 
undergo  neither  photoinduced  double  bond  migration  nor  ether  formation,, 

Other  examples  of  photosensitized  cis^trans  isomerizations  of  simple  olefins 
include  the  isomerization  of  cis,,  trans »  trar;.s-l55i>9=-yclcdodecatriene  to  the  cis, 
cis,  trans  and  trans v  trans,  trans  isomer s/^and  the  isomerization  of  methyl 
oleate  to  methyl  elaidateT1*' 

CONJUGATED  OLEFINS 

The  photosensitized  cis°trans  isomerization  of  the  stilbenes  and  1^,2-diphenyl- 
propenes  in  benzene  solution  has  been  studied  in  considerable  detail sle"2°a:m.  will  be 
discussed  with  particular  reference  to  the  stilbenes 9   the  results  obtained  for 
the  lP2=diphenylpropenes  being  similar  except  when  otherwise  notedo 

The  mechanism  represented  by  eqs  8=12  will  serve  as  a  starting  point  for 
discussion  of  the  isomerizationo  In  this  mechanism  only  vertical  energy  transfer 

(8)  S^_i^ssli^s^ 


-377- 
(9)     S1"  +  £>  _J^  4J*  +  s^     (10)    sl1,/^/^^ 


ll 


(11)     cTl »  tTl  (22)   (a)  t^-fJES-  tTl  -^>c^  (b) 

T1. 

to  cis-  and  trans- stilbene  yielding  the  planar,  spectroscopic  transoid  (t  )  and 

ml  jl       rpl 

cisoid  (c  )  triplets  is  proposed?,  conversion  of  £   to  t   (eq  H)  has  been 

T1 
shown  to  occur  as  will  be  seen  shortly »  Eq  12b,  decay  of  t   to  cis-stilbene, 

is  a  non-vertical,  radiationless  decay  process9  completely  analogous  to  the 

non-vertical  excitation  processes  previously  discussed^  its  importance  will  be 

assayed  later.  The  light  used  was  filtered  to  avoid  direct  excitation  of  either 

of  the  stilbene  isomers,  and  no  indication  of  ojay   singlet  state  reaction  of  the 

stilbenes  (e0g0  ,  phenanthrene  formation)  was  foundo16  Furthermore,  for  high-energy 

sensitizers,  i0ec ,  e3  vertical  excitation  energy  of  either  stilbene  isomer  by 

1  iTll 

at  least  3-5  kcal/mole,  radiationless  decay  of  S   was  found  to  be  negligible , 

ml 

deactivation  of  S   occurring  only  via  energy  transfer  to  the  stilbenes0 

Kinetic  analysis  of  eqs  8-12  leads  to  the  expression  for  the  photostationary 
state  composition  given  by  eg  1J.  For  high-energy  sensitizers  ka,  and  k2  will  be 

( 13)      ( cis)  «/( trans)  „  =  ( kx/ks)  °  ( k4/k3) 

equal  and  diffusion-controlled,  and  the  photostationary  state  composition  will 
be  constant,  being  determined  solely  by  the  decay  ratio,  k4/k3o  Under  such  con- 
ditions the  quantum  yields  are  related  to  (cis) J{ trans)  by  eq  lk.     The  0-0  triplet 

s  s 

{lk)  Kj*n+  =  k4As  s  (£is)  J{ trans) 


2ia,b 


tc'  ct    *  °   v-— '   s'  *zses=/  s 
excitation  energies  of  the  stilbenes  are^v5Y  kcal/mole  for  the  cis  isomer 

and/^50  kcal/mole  for  the  trans  isomer,.21   As  E_  approaches  and  falls  below 

57  kcal/mole,  kg  is  predicted  to  decrease  below  the  diffusion-controlled  limit, 

ki  remaining  constant .,  thus  leading  to  an  increasingly  cis-rich  photostationary 

state0  As  E  approaches  and  falls  below  £Q  kcal/mole  k±   should  also  decrease,, 

Below  50  kcal/mole  the  excitation  ratio  ki/k^  for  the  endothermic  energy  transfer 
process  is  predicted  to  become  constant,  being  determined  solely  by  the  difference 

c       t 

in  the  0-0  triplet  energies  of  the  two  isomers,  i.a  e,  9   kx/ka  =  expt [ ( E--E^) 

^v6  X  106  at  25°.  !Ehis  predicts  a  limiting  photostationary  state  of  essentially 
pure  cis- stilbene. 

The  experimentally  determined  variation  of  the  photostationary  state  com- 
position,8'16 extrapolated  to  infinite  dilution  with  respect  to  the  sensitizer 

g 
concentration,  with  E  for  the  stilbenes,  fig  1,  does  not  completely  follow  these 

predictions,  however.  The  predicted  high-energy  region  of  constant  photostationary 

state  and  subsequent  increase  in  $  cis- stilbene  as  E  decreases  below ™62 

kcal/mole  are  observed^  the  other 
predictions  are  incorrect .  With 

(cis)        q  £    f  ^  low-energy  sensitizers  (Err<(^62  kcal/ 


(trans)      Q  |  6y        '^-v-  ,   mole)  the  value  of  k2  (cis- stilbene 

hfu^1  *  '4o  '  'r1-'  "fcfT"  fr ' '  '4^' '  '^q"  exci'fca'fcion)  falls  off  much  more 
+}     5  J     55  ou  op  (L,     o     slovlj  than  predicted,  a  finding 


E  (kcal/mole)  which  led  Hammond  to  propose  that 

fig  1  triplet  energy  transfer  to  cis- stilbene 

occurs  via  non-vertical  excitation 


-378= 

ml 

with  synchronous  distortion  to  give  the  twisted  phantom  triplet  (p)  and/or  t 


mX 

processes  requiring  less  energy  than  transition  to  c  ;  since  trans- stilbene  is 
more  stable  than  eis-st liberie  by»6  kcal/mole,22  non- vertical  excitation  of  the 

cis  isomer  to  p  or  t   should  only  become  endothermic  when  E  <(^W-  kcal/mole  in 

agreement  with  variable -temperature  kinetic  studies26  which  indicated  no  activation 
energy  for  this  process  in  the  53-60  kcal/mole  range.  Inclusion  of  eq  15  in  the 
kinetic  scheme  leads  to  the  photo stationary  state  composition  given  by  eq  16. 
For  high-energy  sensitizers  kx   ~  k/s^fes  and  eq  16  reduces  to  eq  13.  Below  ^62 

(15)    ST  +  c — ^->tT  ,p  +  SS     (16)    (cis)  /(trans)  -  kxk4/(k2  +  k5)k3 
_  s      "™  s 

kcal/mole  k2  decreases,  kj.  and  k3  remaining  constant  in  accord  with  the  observed 

increase  in  (cis)  /(trans)  .  Below/v 51-52  kcal/mole,  kx  also  decreases  sharply, 

,:_L "  S   '        £» 

k3  remaining  essentially  constant?,  ki/(k2  +  k5)  thus  decreases  explaining  the  general 
shape  of  fig  1  in  the  ^5-50  kcal/mole  region. 

It  is  observed  from  fig  1  that  the  low-energy  sensitizers  eosin  (j>)  and  9j 
10-dibromoanthracene  (6)  establish  a  photo stationary  state  of  essentially  pure 
trans-stilbene „  recent  studies18  have  shown  that  photolysis  of  these  sensitizers 
produces  bromine  atoms  which  cause  thermal  equil.ibiatn  :-r.  of  the  stilbenes.1 

Sensitizers  of  E  <(  53  kcal/mole  were  found  to  give  pronounced  concentration 

effects  in  the  phot ©stationary  state  composition,  (cisj  /(trans)  decreasing 

as  the  concentration  of  the  sensitizer  increased?,  this  finding  was  attributed  to 
reversible  energy  transfer  to  trans -stilbene  (eq  17;  »16  Replacement  of  eq  9  by 
eq  17  in  the  kinetic  scheme  leads  to  eq  18  which  is  fit  by  the  experimental  data. 

(17)    t ,   +  ST   rjp-*-  tT-t  +  S        (18)  (cis)  J^fcrans)  ^k1k4/(k2+k5)(k3+k„1[S   I) 

"1 

T1 
In  further  accord  with  this  idea  of  reversible  energy  transfer  to  t  ,  it  was 

found  that  inclusion  of  the  triplet  quencher  azulene,  az  (£^29=^2  kcal/mole) 

in  the  reaction  mixture  also  gave  photostatlonary  states  richer  in  trans -stilbene 
indicating  eq  19  to  be  operative .  The  effect  of  added  azuiene  was  the  same  for 

ml 

all  sensitizers  usedo  Further  evidence  for  quenching  of  t   comes  from  examination 
of  the  <t>  /<t>  ,  ratio  obtained  for  high-energy  sensitizers  .  While  eq  Ik  was  found 

"CC  CX> 

to  hold  for  the  1,2-diphenylpropene  isomerization,  it  was  not  satisfied  by  the 

stilbenesc  This  deviation  from  eq  Ik  was  found  to  be  due,  at  least  in  part,  to  the 

T1 
existence  of  self-quenching  of  t   (eq  20). 16 

>pl  c<Q  cO  ml  rpl  c<0  qO 

(19)    t   +  az  — — ^  t   +  az      (20)    t   +  t   ->»  2t 

These  quenching  studies  indicate  that  energy  transfer  to  cis-  or  trans- 
stilbene  leads  to  the  ultimate  production  of  the  same  triplet  species  which  is  then 
quenched  to  give  trans-stilbene ,,  i.  e. ,  energy  transfer  to  cjLs-stilbene  produces 
a  triplet  whichican  be  deactivated  to  tr ans " stilbene  by  quenching.  However, 
quenching  of  £   could  not  be  detected.  To  explain  these  findings,  Hammond 

ml       rpl 

postulated  the  rapid,  irreversible  decay  of  £   to  t   (eq  11);  this  interconversion 

rpl 

probably  occurs  via  p.,  believed  to  be  in  equilibrium  with  t   ( eq  21) .  These 

rpl  rpl 

(21)    £   _ — ^   p  ^=±  t 

findings  are  also  evidence  for  the  production  of  an  electronically  excited  triplet( s) 
upon  sensitization  of  cis-stilbene  by  low-energy  sensitizers  indicating  that 
non-vertical  energy  transfer  does  not  involve  transfer  of  vibrational  energy. 
Quenching  effects  were  not  observed  for  the  1,2-diphenylpropene  system,  indicating 
that  conversion  of  both  the  transoid  and  cisoid  triplets  to  p  is  rapid  and  irreversible, 
isomerization  occurring  only  from  p.ls 

Becent  studies  by  Hammond  and  Herkstroeter  have  provided  direct  evidence  as  to  the 


79- 

nature  of  the  energy  transfer  processes. 1:y  These  workers  used  flash  spectroscopy 
to  study  the  rate  of  decay  of  various  sensitizer  triplets  using  cis-  and  trans- 
stilbene  and  1,2-diphenyipropene  as  quenchers.  Attempted  direct  study  of  the 
stilbene  and  1,2-diphenylpropene  triplets  was  unsuccessful  due  to  their  short 
lifetimes.  From  these  quenching  studies  the  rates  of  energy  transfer,  L,  ^  from 

the  sensitizer  triplets  to  the  olefin  acceptors  were  determined..  The  results 
obtained  confirm  what  has  already  been  said  about  the  nature  of  the  energy  transfer 

processes.  Classically,,  if  E  is  less  than  E2,  energy  transfer  will  require  an 

activation  energy  equal  to  E^-E  ,  the  decrease  in  transfer  efficiency  as  a  function 

of  E  being  given  by  eq  22. 

(22)    d(iog  k^/dCE^  -  f  1/2.303  FT 

The  experimentally  observed  quenching  curve  for  trans- stilbene  had  the  slope 

indicated  by  eq  22  for  sensitizers  of  E  ^_48  kcal/mole,  indicating  that  non- vertical 

excitation  of  trans-stilbene  to  p  does  not  occur.  The  behavior  of  cis- stilbene  towards 

sensitizers  of  E_  ~  42-5$  kcal/mole  did  not  fit  eq  22 s   excitation  of  cis-stilbene 

o 

being  quite  efficient  even  for  sensitizers  of  Em  10  kcal/mole  too  low  to  effect 

vertical  excitation,  in  agreement  with  the  postulated  non-vertical  process.  As 
also  inferred  from  photostationary  state  studies,  neither  cis-  nor  trans-1,2" 
diphenylju  rpene  was  found  to  exhibit  classical  behavior  as  a  triplet  quencher 
indicating  that  both  of  these  isomers  undergo  efficient  non-vertical  excitation. 

From  these  studies  it  was  concluded  that  trans- stilbene  can  find  no  excitation 
pathway  of  substantially  lower  energy  requirement  than  vertical  excitation  to 

ml  ml  ml 

t  ,  indicating  p  to  be  close  to  isoenergetic  with  t   but  below  £   in  energy. 
Since  neither  of  the  1,2-diphenylpropenes  behaves  classically,  both  the  cis  and 
trans  isomers  are  apparently  able  to  undergo  transitions  to  one  or  more  twisted 
states  of  lower  energy  than  either  the  planar  cisoid  or  transoid  triplets,  This 
is  probably  a  consequence  of  the  relief  (via  twisting)  of  the  steric  strain  that 
exists  in  both  the  cisoid  and  transoid  triplets  owing  to  mnbonded  interaction 
between  a  phenyl  group  and,  the  phenyl  or  methyl  group  cis  to  it. 

The  results  of  the  quenching  studies  as  well  as  the  observed  non-vertical 
energy  transfer  to  cis -stilbene  may  be  explained  either  by  the  assumption  that  the 

rpl 

two  triplets,  t   and  p,  are  in  equilibrium  (eq  21) 9   or  that  only  a  single  triplet , 

ml  ml 

t  ,  exists.,  non- vertical  excitation  of  cis- stilbene  giving  only  t  .  The 
latter  hypothesis,  although,  not  rigorously  disprove^  appears  to  be  much  less 
likely.  In  order  to  accommodate  this  hypothesis  it  must  be  assumed  that  whereas 

ml 

t   is  selectively  deactivated  to  trans-stilbene  in  quenching  reactions,  spontaneous 

ml 

decay  of  t   yields  both  cds-  and  trans-stilbene »  Furthermore,  inclusion  of  p 

in  the  isomerization  mechanism  allows  a  rationalization  of  the  very  short  lifetime 

of  the  stilbene  triplets,  estimated  from  the  azulene  quenching  studies  to  be 

<£__  To?  X  10  8  sec.  If,  as  seems  reasonable,  p  is  assumed  to  have  a  twisted 

configuration  with  an  angle  of  twists  it/2,  it  may  be  very  close,  both  in  energy 

and  configuration,  to  a  point  on  the  potential  surface  of  the  ground  singlet  state 

(fig  2)|  intersystem  crossing  from  the  triplet  to  the  ground  state  is  expected  to 

be  very  rapid  under  such  conditions.   Saltiel  believes  that  there  may  be  an  actual 

crossing  of  the  singlet  and  triplet  states  in  this  region  as  shown  in  fig  2^2°  ' 

the  energy  of  the  twisted  ground  state  has  been  estimated  from  thermal  isomerization 

studies  to  be~^9  kcal/mole  above  that  of  trans-stilbene . 23 

Recent  work  by  Saltiel  has  provided  additional  evidence  as  to  the  nature  of  p 
and  the  decay  processes  involved  in  the  r^llbene  isomerization.20  The  photostationary 
state  composition  obtained  from  isomerization  of  trans- stilbene-d.12  was  found  to  equal 
that  obtained  from  undeuterated  trans-stilbene  for  sensitizers  of  E^j  =  48-69  kcal/mole 5 


Energy 
(kcal/mole) 


T1 triplet  excited  Tf 


.580= 


63 


angle  of  twist 

Stilbene  Energy  Profile 

fig  2 


the  quantum  yields  of  the  two 
isomeriz?.tions  were  also  identical,, 
Deuteration  is  known  to  decrease  the 
rate  of  triplet  to  singlet  radiationless 
decay,,  the  effect  diminishing  as  the  energy 
separation  between  the  two  states  decreas- 
es.24 If  p  is  of  lover  energy  than  t  f 
deuteration  is  predicted  to  affect  decay 
of  tTl  to  trans - stilbene  more  than  decay  of 
p  to  a  twisted  ground  state 9   assuming  decay 
from  both  tT1  and  p  to  be  operative.  The 
absence  of  such  a  deuterium  effect  led 


Saltiel  to  conclude  that  decay  from  t   is  negligible  for  both  trans -stilbene  and  trans- 
stilbene -di2«  He  believes  p  to  be  of  lower  energy  than  tTl  as  shown  in  fig  2  with  vir- 
tually all  decay  to  the  ground  state  occurring  from  p^  a  point  of  view  not  fully  shared 
by  Hammond.18  Although  decay  to  the  ground  state  from  txl(eq  12)  has  not  been  rigorous- 
ly ruled  out;,  decay  of  the  very  short-lived  cisoid  triplet  to  ground  state  stilbene  is 
not  believed  to  occur ^  internal  conversion  to  p  and/or  t^  being  faster.  If  formation 
of  cis-stilbene  is  assumed  to  occur  mainly  from  p^  then  the  earlier  finding  that  the 
photo stationary  state  becomes  richer  in  cis-stilbene  at  higher  temperatures  can  be  at- 
tributed to  the  existence  of  an  activation  energy  in  the  interconversion  between  p  and 
tT1  as  shown  in  fig  2. 

Photosensitized  isomerization  of  the  stilbenes  by  the  low-energy  sensitizer  phenan- 
thraquinone  (PAQ,  E^  =  48.8  kcal/mole)  has  recently  been  studied  in  benzene  solution  by 
Bohning  and  Weiss 02*  Formation  of  adduct  J  occurred  competitively  with  isomerization. 

(n  %n     The  kinetic  results  obtained  are  in  substantial  agreement  with  the  mech- 
-\Y^  "^f^CsHs  anism  already  discussed  for  the  isomerization,  the  only  major  difference 
•>lks  ^T'CgHs  being  the  inclusion  of  a  short-lived  complex^  X_,  formed  via  triplet  ener- 
\  gy  transfer  to  either  cis-  or  trans -stilbene ,  decay  of  X  being  partit- 

X        ioned  between  collapse  to  2  and  decay  to  p  and  PAQ.  The  existence  of  X 
as  a  common  intermediate  is  strongly  suggested  by  the  observation  that  both  cis-  and 
trans-stilbene  give  the  same  adduct.  The  geometrical  changes  leading  to  p  are  postula- 
ted to  occur  in  X^  in  which  there  is  believed  to  be  freedom  of  torsional  motion  in  the 
stilbene  accounting  for  the  formation  of  7  from  both  the  cis  and  trans  isomers.  Clas- 
sical energy  transfer  to  trjns~stiibene(but  not  to  eis-stilbsne)  was  also  invoked j  for 
in  its  absence  the  quantum  yields  of  adduct  formation  become  independent  of  (cis) s/ 
(transj  s .,  whereas  such  a  dependence  was  observed. 

I^ie  cis-trans  isomerization  of  stilbene  has  been  used  to  examine  the  steric  re- 
quirements of  triplet  energy  transfer26   Theory  predicts  that  this  process  should  be 
subject  to  steric  hindrance  by  bulky  substituents  on  the  donor  or  acceptor.2  In  agree- 
ment with  this  idea.,  it -was  found  that  whereas  the  high-energy  sensitizers  2^3i-5;6-tet- 
ramethyl-M -methoxybenzophenone  (E§  -  70.2  kcal/mole)  and  2^„6-trimethyl»4t-methoxy- 
benzophenone  (E^  =  68.  k  kcal/mole)  produced  the  same  photo  stationary  state  as  benzopl: 
one  (E§  =  68.5  kcal/mole) ^  indicating  diffusion-controlled  energy  transfer,  2^6- 
triisopropyl-^'-methoxybenzophenone  (E§  =  69.9  kcal/mole)  and  2^6-triisopropylhenzo- 
phe"none  (E§  =  68.7  kcal/mole)  ,  compounds  which  are  more  hindered  about  the  carbonyl 
group  where  the  triplet  energy  is  believed  to  be  localized^2  gave  cis  rich  photosta- 
tionary  states  indicating  that  energy  transfer  to  cis-stilbene  is  less  efficient  than 
to  trans-stilbene.  To  insure  that  the  observed  results  were  not  due  to  selective  energy 
transfer  from  the  low-energy  triplets  of  the  photoenols  of  structure  8  formed  from  the 

sensitizer  triplets,  the  rates  of  quenching  of  the  photoenolization 
of  the  triplets  of  2^4s6-trimethyl-4J-methoxybenzophenone  and  2  .'■>-   6« 
triisopropyl-M-methoxybenzophenone  by  added  cis-  and  trans-stil- 
bene were  studied!  the  rate  of  decrease  of  photoenolization  (and 
hence  rate  of  triplet  energy  transfer)  of  2^6-trimethyl-ii-'1-meth- 
oxybenzophenone  was  about  15  times  as  great  as  that  of  2,4,6= 
triiscpropyl-M-methoxybenzophenone^  trans  -  stilbene  being  the 
better  quencher  in  each  case,  confirming  the  existence  of  a  steric  effect  to  energy 
transfer o 

The  photosensitized  cis-trans  isomerization  of  the  piperylenes  is  mechanistically 


>phen= 


-38l- 

similar  to  that  of  the  stilbenes  and  has  been  reviewed  by  Turro,27  and  recently  studied 
on  a  silica  gel-benzene  matrix28  and  under  ferrocene  photosensitized  conditions.29  In 
addition,  the  piperylene  isomerization  has  found  many  uses  in  photochemistry  including 
measurement  of  the  inter system  crossing  quantum  yields  of  sensitizers  whose  triplet 
states  can  effect  the  isomerization,5  and  measurement  of  the  triplet  energies  of  sensi- 
tizers for  which  spectral  data  is  not  available.16-927  Owing  to  their  relatively  low 
triplet  energies  j,  both  cis-  and  trans -piperylene  (E^  =  56.9  and  58.8  kcal/mole,  respec- 
tively) have  been  extensively  used  as  triplet  quenchers  in  mechanistic  studies. 

The  photosensitized  conversion  of  cis „cis-l , 3-cyclooctadiene  to  bicyclo[4.2.0]~ 
oct-7-ene  has  been  found  by  Liu30  to  proceed  via  the  isolable  cis,trans-l,3"-cycloocta- 
diene  which  is  then  thermally  converted  into  bicyclo[4.2.QJ-oct-7-ene  in  accord  with 
the  Woodward-Hoffmann  rules o 

BIBLIOGRAPHY 

lo  Ro  Bo  Cundall,  Progr.  Reaction  Kinetics ,  2,  165  (196*0. 

2.  For  a  review,  see  N.  J.  Turro,  "Molecular  Photochemistry „"  W.  A.  Benjamin, 
Inc,  New  York,  N.  Y.  ,  I965.?  chapter  5° 

3c  R.  S.  Mulliken  and  C.  C.  J.  Roothaan,  Chem.  Rev.  ,  4l,  219  (19^7)  • 

4.  D.  F.  Evans,  J.  Chem.  Soc,  1735  (1960). 

5.  (a)  A.  A.  Lamola  and  G.  S.  Hammond.,  J.  Chem.  Fhys.,  4j5,  2129  (1965)5  (b)  M.  A. 
Golub,  et  al. ,  J.  Chem.  Phys.,  4£,  1503  (1966). 

6.  Z.  R.  Grabowski  and  A.  Bylina^  Trans.  Faraday  Soc,  60,  1131  (1964). 

7.  Ref  2,  pp  30-42. 

8.  G.  S.  Hammond,  Kiagaku  to  Kogyo  ( Tokyo)  ,  18,  1464  (1965). 

9.  (a)  G.  S.  Hammond  and  J.  Saltiel,  J.  Am.  Chem.  Soc,  8j>,  25l6  (1963)5  (b)  ref  2, 
pp  182-3. 

10.  R.  F.  Borkman  and  D.  R.  Kearns,  J.  Am.  Chem.  Soc. ,  88,  3467  (1966). 

11.  R.  B.  Cundall  and  A.  S.  Davies,  Proc.  Roy.  Soc.  (London),  A290,  563  (1966). 

12.  (a)  H.  Morrison,  Tetrahedron  Letters,,  3653  (1964)5  (b)  H.  Morrison,  J.  Am. 
Chem.  Soc,  8j,  932  (1965)5  (c)  H.  Morrison,,  et  al.  ,  Abstracts,  153rd 
National  Meeting  of  the  American  Chemical  Society,  Miami  Beach,  Fla. ,  I967.9 
p  01375  (d)  H.  Morrison,  private  communication. 

13.  (a)  P.  J.  Kropp,  J.  Am.  Chem.  Soc,  88,  4091  (1966);  (b)  J.  A.  Marshall  and 
R.  D.  Carroll,  ibid.  ,  88,  4092  ( 196617 

14.  H.  Nozaki,  et  al.,  Tetrahedron  Letters,  2l6l  (1965). 

15.  A.  C.  Testa,  J.  Org.  Chem. ,  2£,  246l  (1964). 

16.  G.  S.  Hammond,  et  al. .  J.  Am.  Chem.  Soc,  86,  3197  (1964). 
17-   S.  Malkin  and  E.  Fischer,  J.  Phys.  Chem.,  fsB,  1153  (1964). 

18.  G.  S.  Hammond,  private  communication. 

19.  W.  G.  Herkstroeter  and  G.  S.  Hammond.  J.  Am.  Chem.  Soc,  88,  4769  (I966). 

20.  (a)  J.  Saltiel,  ibid8  8£,  IO36  (1967)5  (b)  private  communication. 

21.  (a)  D.  F.  Evans,  J.  Chem.  Soc,  1351  (1957)  J  (b)  R.  H.  Dyck  and  D.  S.  McClure, 
J.  Chem.  Phys.,  j>6,  2326  (1962). 

22.  R.  B.  Williams,  J.  Am.  Chem.  Soc,  64,  1395  (.1942). 

23.  G.  B.  Kastiakowsky  and  W.  R.  Smith,  ibid„  ^6,  638  (1934). 

24.  Ref  2,  pp  69-70. 

25.  J.  J.  Bohning  and  K.  Weiss,  J.  Am.  Chem.  Soc,  88,  2893  (1966). 

26.  G.  S.  Hammond,  et  al.,  J.  Am.  Chem,  Soc,  88,  4777  (1966). 

27.  Ref  2,  pp  178-81. 

28.  P.  A.  Leermakers,  et  al. ,  J.  Am.  Chem.  Soc,  88,  3176  (1966). 

29.  J.  J.  Dannenberg  and  J.  H.  Richards,  ibid,  87,  1626  (1965). 

30.  R.  S.  H.  Liu,  ibid,  89,  112  (1967). 


xfto. 


Reported  by  James  E. 
INTRODUCTION 


THE  ABNORMAL  CLAISEN  REARRANGEMENT1 
Shaw 


May  25,   196? 


In  I.936  Lauer  and  Filbert  reported  that  rearrangement  of  y-ethylallyi  phenyl 
ether  (I)  in  N,N-diethyls.niline  at  201-225°C  did  not  result  in  the  normal  Claisen 
rearrangement  product,  £-(a-ethy!aIlyl) phenol  (III),  but  instead  gave  £- (a ,7 -dimethyl - 
allyl)  phenol  (II) 


.1-3 


0~CH2CH~CHCH2CH3 
i-  a    3  y  b    e 


E 


^ 


CH3 

^CHCH-CHCH^ 
a  3  y 


I 


II 


CHCHsCH2 


III 


The  abnormal  product  II  appears  to  arise  from  attachment  of  the  p-or  5 -carbon  of  the 
allyl  chain  to  the  ortho  carbon  of  the  benzene  ring.  The  normal  Claisen  rearrange- 
ment involves  7 -•attachment,   Lauer  and  Filbert  found  that  ozonolysis  of  their 
product  gave  small  amounts  of  formaldehyde  in  addition  to  the  expected  acetaldehyde j, 
however,  they  failed  to  attribute  the  formaldehyde  as  possibly  arising  from  the 
normal  product  III.  It  was  later  shown  by  Hurd  and  Pollack4  that  under  the  same 
conditions  both  the  normal  and  abnormal  products  are  formed  in  a  1.3  :  1  ratio. 
These  workers  also  shewed  that  the  aliphatic  analogue,  7-ethylallyl  vinyl  ether 
(IV),  rearranged  to  give  some  abnormal  product  V  in  addition  to  the  major  normal 
Claisen  rearrangement  product  VI. 


CH2=CH-0-CH2CH*=CHCH2CH3 
IV 


??o0r  P3 

->    CH3CH=CHGHCHaCHO 


sealed, 
be 


+  CH* 


CH2CH3 
=CHCHCH2CHO 

VI  :  9W 


or 


Several  other  examples  of  the  abnormal  Claisen  rearrangement  are  given  in  the 
literature,5"'10  In  all  of  these  cases  the  allyl  phenyl  ether  which  undergoes 
rearrangement  contains  a  7 -substituted  methyl,  ethyl,  or  n-propyl  group.  No 
abnormal  Claisen  rearrangements  have  been  reported  for  a-  or  3-alkyl  substituted 
7-aryl  substituted  allyl  phenyl  ethers.11'12  It  appears  that  7 -secondary  or  7- 
tertiary  alkyl  substituted  allyl  phenyl  ethers  have  not  been  investigated ,  The 

:-   of  this  seminar  will  be  to  examine  the  mechanism  of  the  abnormal  Claisen 
rearrangement  in  both  aromatic  and  aliphatic  systems, 

MECHANISM 

For  the  rearrangement  of  7-ethylallyl  phenyl  ether  Hurd  and  Pollack  proposed 
the  following  cyclic  mechanism  which  involves  attachment  of  the  5-carbon  of  the  cis 
configuration  of  the  allyl  chain  to  the  ortho  carbon  of  the  benzene  ring.4 

CH3\_ 


JIT 

?H3 


^ 


II 


However,  the  abnormal  rearrangements  of  ethyl  £-(7-n-propylallyloxy)benzoate  (VII) 
and  7,7-dimethylallyl  estrone  ether  (VIII)   have  shown  that  this  mechanism  is 

,  and  that  the  abnormal  attachment  of  the  allyl 

group.5'10     For  compound  VII  the  abnorma         oduct  DC  could  result  from  attachment 

he  p-  or  e-carbon.      If  the  abnormal  rearrangements  of  both  the  7 -propyl  and 
7~eT  Lyl  phenyl  ethers  a  oceeding  by  a  single  mechanism,  3 -attachment  must 


OCH2CH=CHCH2CH2CH3 

213-24l°C 


H3 

HCH=CHCH2CH3 


-0  mm. 


COCX^sHs 
VII 

be  involved.  The  abnormal  rearrangement  of  ether  VIII  would  result  in  product  XI 
if  ^-attachment  occurred,  and  product  XII  is  there  was  6 -attachment. 


CH2CH2CH3 
' HCH=CHp 


VIII 


XI 


XII 


Since  the  observed  product  was  XI ,  it  appears  that  the  abnormal  rearrangement 
involves  attachment  of  the  p -carbon  of  the  allyl  chain  to  the  benzene  ring. 

Lauer  and  co-workers  investigated  the  possibility  that  the  rearrangement  of 
the  crotyl  phenyl  ether  XIII,  the  simplest  of  the  /-substituted  allyl  phenyl  ethers, 
was  proceeding  by  the  abnormal  path  involving  (3 -attachment  of  the  allyl  group  rather 
than  by  the  normal  route  involving  /-attachment. 13  In  both  cases  the  product  would 
be  the  same.  However,  if  the  /-substituted  methyl  group  were  tagged  with  14C,  the 
normal  rearrangement  would  give  XIV  while  the  abnormal  rearrangement  should  give  XV. 


OCH2CH=CHCH3 


220-235°C 
sealed 
tube 


* 

CH3 

CHCH«CH2 


OOCaHs 
XIII 


C00C2H5 
XIV 


CH3  %. 

£hch=ch; 


Ozonolysis  of  the  product,  followed  by  a  study  of  the  radioactivity  of  the 
formaldehyde  produced,  showed  that  15-29$  of  the  product  had  been  produced  by  the 
abnormal  rearrangement  process.  To  account  for  the  abnormal  rearrangement,  Lauer 
proposed  the  following  mechanism  which  involves  (3 -attachment  of  the  cis  configuration 
of  the  allyl  chain. 


CH2    CHp 
lCH=^CH 


CH3  * 
CHCH=CH; 


XV 


COOC^ 


COOC^s 


However,  in  Lauer' s  work  it  was  not  shown  whether  ether  XIII  was  cis  or  trans .   Schmid 
and  co-workers14'15  studied  the  rearrangement  of  the  cis  isomer  of  /-14C-methylallyl 
P_-tolyl  ether  (XVI).  Heating  the  ether  for  three  hours  at  230°C  in  N,N-diethyl- 
aniline  gave  a  rearrangement  product,  2-(a-methylallyl) -4-methylphenol,  which  showed 
by  examination  of  the  formaldehyde  produced  upon  ozonolysis,  that  kCffo   of  the 
reaction  was  apparently  proceeding  by  the  abnormal  route  in  comparison  to  Lauer1 s 
15-29$. 


,/ 


-38*i— 

* 

CH2^ 

CH3 


XVI 


Marvell  and  co-workers16  were  the  first  to  show  that  the  abnormal  Claisen 
product  was  not  formed  directly  from  the  ally!  phenyl  ether,  but  instead  was  due  to 
further  rearrangement  of  the  initially  formed  normal  Claisen  product.  The  rearrange- 
ment  of  7 -ethylallyl  phenyl  ether  (I)  in  N,N-diethylaniline  at  195°C  was  followed 
by  infrared  measurements  and  gave  the  data  shown  in  the  graph  below. 


.......  Ether 

Normal  product  III 

Abnormal  product  II 


Mole/L. 


10     20     30     kO 

Time  (hours) 


50 


Furthermore,  if  the  normal  Claisen  product  III  was  heated  in  N,N-diethylaniline  or 
neat  at  200-225°C,  it  slowly  rearranged  to  the  abnormal  product  II„  Roberts  and 
Landolt17  have  found  that  heating  either  XVII  or  XVIII,  the  normal  and  abnormal 
products  of  7 -ethylallyl  g-tolyl  ether,  results  in  the  same  equilibrium  mixture  of 
the  two  with  the  abnormal  product  being  favored  by  a  ratio  of  2k    1   1. 


CH2CH3 
CHCH=CH2 


XVII 


200°C, 


"pit 


3 


H3 
HCH^CHCH3 


XVIII 


Jefferson  and  Scheinmann10  have  reported  that  rearrangement  of  7,7-dimethylallyl 
estrone  ether  (VIII)  in  N,N-diethylaniline  produces  only  the  abnormal  product  XI. 
However,  if  the  rearrangement  is  carried  out  in  diethylaniline  containing  butyric 

anhydride,  the  normal  Claisen  product  can  be 
trapped  as  its  butyric  ester  XIX.  The  normal 
Claisen  product  rapidly  isomerized  to  the  abnormal 
product  XI  when  it  was  heated  in  diethylaniline. 
On  the  basis  of  this  evidence,  Marvell,  Roberts,  and 
Jefferson  conclude  that  the  abnormal  Claisen 
rearrangement  is  really  the  result  of  two  consecutive 
processes?  normal  Claisen  rearrangement  of  the  7- 
alkylallyl  aryl  ether  to  the  o-(a-alkylallyl) phenol, 
followed  by  rearrangement  of  this  phenol  to  produce  the  isomeric  phenol.  Although 
the  evidence  strongly  supports  this  conclusion,  it  cannot  rule  out  another  possibility 
in  which  the  normal  product  revcrsibly  forms  the  allyl  aryl  ether  which  then  reacts 


CHrjCHgCH^C 


XIX 


-385- 

by  some  mechanism  to  give  the  abnormal  product  directly.  However,  in  equilibration 
experiments  such  as  that  involving  XVII  and  XVIII ,  no  ether  has  ever  been  reported 
found  014'17 

Marvell  and  co-workers16  also  reported  that  the  methyl  ether  of  o~(a-ethyl- 
allyl) phenol  (III)  was  recovered  unaltered  upon  heating  it  under  the  conditions  which 
converted  o-(a-ethylallyl) phenol  to  the  abnormal  product  II0  This  indicated  that 
the  rearrangement  of  the  normal  product  to  the  abnormal  product  depends  on  the 
phenolic  hydroxy!  group.  Also,,  2 ,6-dimethyl-4-(a-ethyla!lyl) phenol  was  stable  under 
the  same  conditions  showing  that  the  allyl  side  chain  must  be  ortho  to  the  hydroxyl 
group  in  order  for  rearrangement  to  occur.  Marvel!  proposed  the  following  mechanism 
for  the  rearrangement  between  the  normal  and  abnormal  products. 


►  CH2 
Of    CH 


X^chc: 


HCH2CH3 


6" 


H3 
HCH-CHCH3 


1.4,1 


XX 


II 


XXI 


Infrared  studies  of  o-allylphenols  have  revealed  that  the  phenolic  proton  is  hydrogen 
bonded  to  the  ally!  it-bond,  thus  indicating  that  the  molecule  is  in  the  proper 
conformation  for  reaction.18  When  the  spirodienone  intermediate  XX  is  formed  from 
III,  the  ethyl  group  could  also  be  trans  to  the  methyl  group,  but  this  configuration 
would  not  allow  further  rearrangement  to  II  by  the  intramolecular  process  shown. 
q         The  intermediate  XX  is  similar  to  spiro[205]octa-I,^--dien-3-one 
(XXI)  isolated  by  Winstein  and  Baird.19'20  Since  this  compound 
decomposed  upon  standing  at  room  temperature,  it  is  quite  unlikely 
that  a  spirodienone  intermediate  such  as  XX  could  be  Isolated 
under  the  conditions  of  the  abnormal  Claisen  rearrangement. 

Marvell 3s  mechanism  is  strongly  supported  by  experiments 
dealing  with  14C  labeling,  deuterium  incorporation,  and  cis -trans 
isomerization  of  substituted  o-allylphenols .  Lauer  and  Johnson^'1 
have  shown  that  heating  y-14C-methyla!lyl  g-carbethoxyphenyl  ether  (XIII)  for  280 
hours  at  220°C  results  in  a  50  s SO  distribution  of  14C  between  the  two  positions 
shown  in  XIV  and  XV.  This  is  the  result  predicted  by  Marvel! !s  mechanism  since  as 
shown  in  intermediate  XX  (replace  ethyl  by  methyl) 9   the  14C  labeled  methyl  group  can 
become  one  of  fcwo  symmetrically  equivalent  methyl  groups.  This  should  allow  equal 
distribution  of  the  radioactive  label  between  the  methyl  and  methylene  positions  in 
XIV  and  XV.  Similar  results  have  been  obtained  by  Schmid.14 

Schmid  and  co-workers25*'23  have  studied  the  incorporation  of  deuterium  into  2- 
(a-methylailyl)  -if- methylphenol  (XXII).  As  shown  below  the  deuterated  phenol  XXI.Ia 
can  form  either  the  cis  spirodienone  intermediate  XXIII  or  the  trans  intermediate 
XXIV  by  intramolecular  transfer  of  the  deuterium  atom.  The  trans  intermediate  would 
probably  be  formed  more  often  because  it  is  very  likely  that  the  cis  intermediate  is 
of  higher  energy,  due  to  steric  factors.  In  the  cis  intermediate,  proton  transfers 
from  the  methyl  groups  can  produce  phenols  XXIIb  or  XXIIc.  However,  in  the  trans 
intermediate  only  phenol  XXIIb  can  be  formed  by  an  intramolecular  proton  transfer. 
Therefore,  if  2-(a-methylally!) -4-methylphenol  (XXII)  is  heated  in  D20,  deuterium 
should  be  initially  incorporated  more  rapidly  at  the  methylene  position  ( sCH2)  than 
at  the  a -methyl.  However,  at  equilibrium  the  amount  of  deuterium  incorporated  at 
the  methylene  carbon  should  be  statistically  equal  to  that  at  the  methyl.  In 
other  words,  if  n  equals  the  amount  of  deuterium  on  the  methylene  carbon  and  m  the 
amount  on  the  methyl,  then  at  equilibrium  3n  should  equal  2m  or  3n/2m  should  equal 
one  assuming  no  deuterium  isotope  effects. 


-386- 
OH 
^^^CHCH-CHD 

V 


CH3 


XXIIb 


CHaD 


HCH< 


M 


XXIIa 


CHCH=CH2 
CH2D 


XXIIc 


By  heating  2-(a-methylallyl) -4-methylphenol  (XXII)    in  D20  at  200°C  for  various  periods 
of  time  and  determining  the  deuterium  content  and  distribution  by  combustion  and  nmr, 
Schmid  obtained  the  results  shown  in  Table  Ie 

Table   I 


Time  heated 
(hours) 


fo  D  in  side  chain 
100/o  =  5  D 


k 
2k 
48 


k.e 
k9.k 

68.6 


3n/2m 

1.78 
1.18 


No  deuterium  was  incorporated  at  the  a   or  p  positions  of  the  allyl  group.  The  fact 
that  3n/2m  equals  approximately  five  after  four  hours  heating  indicates  that 
deuterium  is  initially  incorporated  more  rapidly  at  the  methylene  position  as 
predicted »  When  the  deuterated  phenol  which  possessed  a  3n/2m  value  of  1.18  was 
heated  in  water  for  ten  hours,  the  3n/2m  value  changed  to  O.67,  showing  that 
deuterium  is  more  rapidly  removed  from  the  methylene  carbon  than  the  methyl  carbon 
as  would  be  predicted. 

Schmid  also  studied  deuterium  incorporation  into  3>5-dimethyl-2-(a-methylallyl) 
phenol  (XXV).22'23 

H  0    TT  CH3 

■113 

HCH2sCH2 


XXV 


XXVI 


In  this  case  the  trans  intermediate  XXVI  should  be  disfavored  due  to  repulsion 
between  the  cyclopropyl  methyl  group  and  the  methyl  group  at  the  3-position  of  the 
benzene  ring.   It  would  be  expected  therefore  that  an  increased  amount  of  cis 
intermediate  possessing  symmetrically  equivalent  methyl  groups  should  help  to 
equalize  the  rates  at  which  deuterium  is  incorporated  at  the  methylene  and  methyl 
positions.   In  other  words,  the  value  of  3n/2m  should  approach  a  value  of  one  much 


more  quickly  for  XXV  than  it  did  for  XXII.  It  was  found  that  when  XXV  was  heated  in 
D20  at  200°C  for  only  twelve  hours,  3n/2m  equaled  1.04.   If  this  result  is  compared 
with  those  for  phenol  XXII  in  Table  I,  it  is  apparent  that  the  formation  of  the 
trans  intermediate  XXVI  was  significantly  suppressed. 

By  the  Marvell  mechanism,  any  _o-allylphenol  should  be  able  to  form  a  spiro- 
dienone  intermediate  similar  to  XX.  For  this  reason  compounds  XXVII  and  XXVIII  could 
be  expected  to  incorporate  deuterium  at  the  7-carbon  of  the  allyl  group. 22*23 


a  p  7 

H2CH*sCH2 

OH 


CH2C —  CH2 
CH3 


XXVII 


XXVIII 


Heating  l-allyl-2-naphthol  (XXVII)  in  D20  at  200°C  for  48  hours,  resulted  in  almost 
complete  incorporation  of  deuterium  at  this  position  only.  When  XXVIII  was  heated 
in  D20  at  200°C  for  48  hours,  I.85  deuterium  atoms  were  found  on  the  7-carbon  and 
less  than  0.1  on  the  6 -methyl  group. 

The  thermal  interconversion  of  cis  and  trans  o-( 7 -alkylallyl) phenols  can  also 
be  explained  by  the  Marvell  mechanism.  Schmid  and  Frater24  found  that  heating  either 
cis-  or  trans -XXIX  in  diethylaniline  at  200°C  for  several  days  resulted  in  the  same 
equilibrium  mixture  of  the  two  isomers. 


200°C 


CH3   \/    CH3 

cis -XXIX 


K,     /  .  -3.6 
trans/ cis 


:h3 

trans -XXIX 


The  methyl  ethers  of  cis-  or  trans -XXIX  did  not  interconvert  under  these  conditions, 
showing  that  the  hydroxyl  group  is  necessary  for  the  isomerization.  Marvell 's 
mechanism  can  account  for  this  cis -trans  interconversion,  since  the  7-carbon  of  the 
allyl  group  becomes  a  saturated  center  in  the  spirodienone  intermediate.  Marvell25''26 
has  shown  that  heating  either  cis  or  trans  o-(a-7-dimethylallyl) phenol  (II)  at  210°C 
produces  the  same  equilibrium  mixture  of  22^>  cis-II  and  78$  trans -II.  The  isomeriza- 
tion of  cis-II  in  water  and  D20  at  205°C  followed  first  order  kinetics  with  k^/k^ 
equal  to  2.8.  The  rate  of  formation  of  trans -II  from  cis-II  in  D20  was  found  to  be 
equal  to  the  rate  of  olefinie  deuterium  incorporation  as  measured  by  nmr.  The  nmr 
of  trans -II  which  was  isolated  after  0.7,?  2.0,  3«6,  and  4.8  half  lives  showed  the 
presence  of  one  olefinie  proton.  The  7 -methyl  appeared  as  a  clean  singlet.  This 
showed  that  there  was  one  deuterium  at  the  7-carbon  of  each  molecule  of  trans -II. 
These  data  show  that  every  molecule  of  cis-II  which  is  converted  to  trans -II 
incorporates  one  deuterium  into  the  allyl  side  chain  at  the  7-carbon  as  shown  below. 

CH3 


\ 

rVicH    H 

CHa 


0 


XXX 


Therefore,  the  proton  transfer  in  intermediate  XXX  to  give  the  trans  isomer  must  be 
completely  stereoselective  within  the  limits  of  the  experiment. 


-388- 


ALIFHATIC  ANALOGUES 


Marvell' s  mechanism  can  account  for  the  formation  of  the  abnormal  product  V  in 
the  rearrangement  of  7-ethylallyl  vinyl  ether  (IV)  as  shown  below. 


normal 
Claisen  ) 


of*N 


0 


^ 


H 


21 


II 


XXXII 


XXXIII 


The  enols  XXXI  and  XXXII  are  directly  analogous  to  the  o-allylphenols  obtained  in 
the  rearrangement  of  7-ethylallyl  phenyl  ether  (I). 

That  the  above  mechanism  is  indeed  involved  in  aliphatic  systems  is  supported 
by  deuterium  isotope  experiments  by  Roberts  and  co-workers. 27f2B     These  workers 
heated  ^-pentenophenone-2-d2(XXXIV)and  3-methyl-4-pentenophenone-2-d2(XXXV)  in  sealed 
tubes  at  202°C  and  followed  the  changes  in  the  deuterium  distribution  by  nmr. 


C6H5-C- 


(a)    (b)    (c)(d) 
■CD2-CH2-CH«CH2 

XXXIV 


(  (a)  (b)(c)(d) 
CgH5-C-CD2-CH-CH=CH2 
"  CH3(e) 

XXXV 


By  applying  the  Marvell  mechanism,  it  can  be  predicted  that  at  equilibrium  compound 
XXXIV  should  have  one  deuterium  at  each  of  positions  (a)  and  (d)  and  no  deuterium 
at  positions  (b)  and  ( c) .  Compound  XXXV  at  equilibrium  should  have  0,57  deuteriums 
at  each  of  positions  (a)  and  (d) ,  0.86  deuteriums  at  position  (e),  and  no  deuterium 
at  positions  (b)  and  (c).  The  experimental  results  are  given  in  Table  II. 

Table  II 


Compound 


Time  heated 
(hours) 


Deuterium  (g-atom) 
CH2  (a)      CH2  (d)      CH3  (e) 


XXXIV 


XXXV 
11 


17 

72 

1^5 

12 

hQ 

121 


1.73 
1.23 

0.97 

1.27 
O.83 
O.58 


0.20 
0.7^ 

0.90 

0.60 
0.63 
0.63 


0.12 
0.52 
0.66 


Both  compounds  had  no  deuterium  in  positions  (b)  and  (c).  These  results  are  in 
fairly  good  agreement  with  those  predicted  by  the  Marvell  mechanism.  However,  the 
results  are  complicated  by  the  fact  that  mass  spectroscopy  of  the  product  obtained 
after  heating  XXXIV  for  72  hours  showed  the  presence  of  d0,  d3,  and  d4  molecules 
indicating  that  intermolecular  deuterium  exchange  had  also  occurred. 

Other  experimental  work  by  Roberts  and  co-workers  supports  the  intermediacy 
of  the  cyclopropyl  compound  XXXIII  in  the  Marvell  mechanism.28'29  l-Acetyl-2,2- 
dimethylcyclopropane  (XXXVI)  smoothly  rearranged  to  5-methyl-5-hexen-2-one  (XXXVII) 
at  temperatures  above  150°C.  The  reaction  follows  first  order  kinetics  and  has  an 
activation  energy  of  30  kcal/mole  and  an  entropy  of  activation  of  -10  eu.  The 


o  ox  uv^  wux«~  oi  t-^mp^una  juui«i  Xt>  oUt_n 


Hie  methyl  group  must  be  cis  to  the  carbonyl 


CH3C=0 


163°C 
sealed 
tube 


fi       £-3 

CH3C  -CHgCHgC  -CH2 


XXXVII 


group  and  thus  in  a  favorable  position  for  the  intramolecular  1,5 -hydrogen  shift  as 
shown  in  intermediate  XXXIII.  That  a  cis  relationship  is  required  between  the 
carbonyl  and  methyl  groups  in  order  for  the  rearrangement  to  occur  is  shown  by  the 
thermal  isomerization  of  the  eis  and  trans  isomers  of  l-acetyl-2-methylcyclo- 
propane.28'29  At  l60°C,  the  cis  isomer  rearranged  almost  completely  in  twelve  hours 
to  the  expected  5-hexen»2-one,  but  at  l80°C,  the  trans  isomer  decomposed  only 
slightly  over  a  period  of  2k   hours  and  the  product  was  not  5-hexen-2-one.   Other 
examples  of  this  same  type  of  rearrangement  are  reported  in  the  literature.30""32 
The  rearrangement  of  l-aeetyl-2,2-dimethylcyclopropane  (XXXVI)  is  analogous  to  the 
rearrangement  of  cis -1 -methyl- 2-vinylcyclopropane  (XXXVIII)  involving  the  homo- 
dienyl-l,5»hydrogen  shift.  33-»34 


sjx 


£H< 


160°C 


XXXVIII 

This  rearrangement  follows  first  order  kinetics  and  has  an  energy  of  activation  of 
30  kcal/mole  and  an  entropy  of  activation  of  -12  eu. 9  which  are  very  similar  to 
those  values  previously  mentioned  for  XXXVI.  trans °l-Msthyl-2-vinylcyclopropane  is 
stable  at  250°C.   Ohloff31  has  studied  the  thermal  rearrangement  of  cyclopropane 
XXXIX  which  contains  both  vinyl  and  carbonyl  groups  cis  to  methyl  groups. 


300°C 

50  ram 


+ 


X*S 


CH.3  CH3 
XXXIX 


XL 


XLI 


The  major  product  XL  was  due  to  reaction  involving  the  carbonyl  group.  The  other 
product  XLI,  which  was  due  to  reaction  of  the  vinyl  group,  was  formed  in  approximately 
ICffo   yield. 

SUMMARY 

The  abnormal  Claisen  rearrangement  has  been  identified  as  the  result  of  two 
consecutive  processes;  normal  Claisen  rearrangement  of  the  7-alkylallyl  aryl  ether 
to  an  £-(a-alkylallyl) phenol,  followed  by  rearrangement  of  the  side  chain  of  this 
phenol  to  produce  an  isomeric  phenol.  The  mechanism  of  the  secondary  rearrangement 
has  been  formulated  as  involving  a  substituted  spiro[2.5]octa-4,6-dien-3-one  inter- 
mediate. Considerable  experimental  work  has  provided  strong  support  for  this 
mechanism.  Abnormal  Claisen  rearrangements  involving  aliphatic  systems  appear  to 
be  closely  analogous  to  the  aromatic  case. 

BIBLIOGRAPHY 

1.  W.  M.  Lauer  and  W.  F.  Filbert,  J.  Am.  Chem.  Soc,  £8,  1388  (1936). 

2.  W.  M.  Lauer  and  H.  E.  Ungnade,  ibid. ,  58,  1392  (1935). 


_xon_ 

3.  W.  M.  Lauer  and  L.  I.  Hansen,  ibid.  ,  6l,  3039  (1939). 

4.  C.  D.  Hurd  and  M.  A.  Pollack,  J.  Org.  Chem. ,  3,  550  (1939). 

5.  W.  M.  Lauer  and  R.  M.  Leekley,  J.  Am.  Chem.  Soc,  6l,  3043  (1939). 

6.  W.  M.  Lauer  and  H.  E.  Ungnade,  ibid. ,  6l,  3047  (1939). 

7.  C.  D.  Hurd  and  M.  D.  Puterbaugh,  J.  Org.  Chem.,  2,  381  (1938). 

8.  W.  L.  Lauer  and  0.  Moe,  J.  Am.  Chem.  Soc,  65,  2B9  (I9U3). 

9.  W.  L.  Lauer  and  P.  A.  Sanders,  ibid.  ,  6^,  198  (19^-3). 

10.  A.  Jefferson  and  F.  Scheinmann,  Chem.  Commun. ,  239  (1966). 

11.  D.  S.  Tarbell,  Org.  Reactions,  2,  1  (1944) . 

12.  D.  S.  Tarbell,  Chem.  Rev.,  27,  ^95  (19^5). 

13.  W.  M.  Lauer,  G.  A.  Doldouras,  R.  E.  Hileman  and  R.  Liepins,  J.  Org.  Chem.,  26, 
W5  (1961). 

14.  A.  Habich,  R.  Barner,  W.  von  Philipsborn,  and  H.  Schmid,  Helv.  Chim.  Acta., 
4^,  19^3  (1962). 

15.  H.  Schmid,  Gazz.  Chim.  Ital. ,  92,  968  (1962). 

16.  E.  N,  Marvell,  D.  R.  Anderson,  and  J.  Ong,  J.  Org.  Chem.,  27,  IIO9  (I962). 

17.  R.  M.  Roberts  and  R.  G.  Landolt,  ibid.,  ^1,  2699  (I966). 

18.  A.  W.  Baker  and  A.  T.  Shulgin,  J.  Am.  Chem.  Soc,  80,  5358  (I958). 

19.  S.  Winstein  and  R.  Baird,  ibid.,  7£,  4238  (1957). 

20.  S.  Winstein  and  R.  Baird,  ibid.,  B|,  567  (1963). 

21.  W.  M.  Lauer  and  T.  A.  Johnson,  J.  Org.  Chem.,  28,  2913  (1963). 

22.  A.  Habich,  R.  Barner,  W.  von  Philipsborn,  and  H.  Schmid,  Helv.  Chim.  Acta.,  48, 

1297  (1965). 

23.  H.   Schmid,   Oesterr.   Chemiker-Ztg. ,  6^,  109  (1964). 

24.  G.   Frater  and  H.   Schmid,  Helv.   Chim.  Acta.,  4£,  1957  (1966). 

25.  E.   N.   Marvell  and  B.   Schatz,  Tetrahedron  Letters,  67  (1967). 

26.  E.   N.   Marvell,  J.  L.   Stephenson,  and  J.   Ong,  J.  Am.   Chem.   Soc,  87,  1267  (1965), 

27.  R.  M.   Roberts,  R.   N.    Greene,  R.   G.   Landolt,  and  E.   W.  Heyer,  ibid.,  87,  2282 
(1965). 

28.  R.  M.   Roberts,  R.   G.   Landolt,  R.   N.   Greene,  and  E.   W.   Heyer,  ibid.  ,  8£,  l404 

(1967). 

29.  R.  M.  Roberts  and  R.  G.  Landolt,  ibid.,  87,  2281  (I965). 

30.  D.  E.  McGreer,  N.  W.  K.  Chiu,  and  R.  S.  McDaniel,  Proc  Chem.  Soc,  415  (1964). 

31.  G.  Ohloff,  Tetrahedron  Letters,  3795  (I965). 

32.  K.  von  Auwers  and  0.  Ungemach,  Ann.,  511,  152  (1934). 

33.  R.  J.  Ellis  and  H.  M.  Frey,  J.  Chem.  Soc,  5578  (1964). 

34.  D.  S.  Glass,  R.  S.  Boikess,  and  S.  Winstein,  Tetrahedron  Letters,  999  (I966).