Zn-ProPhenol Catalyzed Enantio- and Diastereoselective Direct

Dec 3, 2017 - We report a Zn-ProPhenol catalyzed reaction between butenolides and imines to obtain tetrasubstituted vinylogous Mannich products in goo...
1 downloads 22 Views 1MB Size
Subscriber access provided by RMIT University Library

Communication

Zn-ProPhenol Catalyzed Enantio- and Diastereoselective Direct Vinylogous Mannich Reactions Between #,#- and #,#-Butenolides and Aldimines Barry M. Trost, Elumalai Gnanamani, Jacob S. Tracy, and Christopher Kalnmals J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b11361 • Publication Date (Web): 03 Dec 2017 Downloaded from http://pubs.acs.org on December 3, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Zn-ProPhenol Catalyzed Enantio- and Diastereoselective Direct Vinylogous Mannich Reactions Between α,β- and β,γ-Butenolides and Aldimines Barry  M.  Trost*,  Elumalai  Gnanamani,  Jacob  S.  Tracy,  Christopher  A.  Kalnmals   Department  of  Chemistry,  Stanford  University,  Stanford,  California  94305   Supporting  Information  Placeholder ABSTRACT:   We   report   a   Zn-­‐ProPhenol   catalyzed   reaction   between   butenolides   and   imines   to   obtain   tetrasubstituted   vinylogous  Mannich  products  in  good  yield  and  diastereose-­‐ lectivity   with   excellent   enantioselectivity   (97   to   >99.5%   ee).   Notably,   both   α,β-­‐   and   β,γ-­‐butenolides   can   be   utilized   as   nucleophiles  in  this  transformation.  The  imine  partner  bears   the   synthetically   versatile   N-­‐Cbz   group,   avoiding   the   use   of   the   specialized   aryl   directing   groups   previously   required   in   related  work.  Additionally,  the  reaction  can  be  performed  on   gram   scale   with   reduced   catalyst   loading   as   low   as   2   mol%.     The  functional  group-­‐rich  products  can  be  further  elaborated   using  a  variety  of  methods.  

Nitrogen-­‐containing   butenolides   are   common   motifs   in   a   variety   of   natural   products   and   pharmaceutical   compounds,   1 and  are  also  useful  synthetic  intermediates.  For  example,  (-­‐)-­‐ securinine  is  a  GABAA  antagonist,  and  its  analogs  have  been   2 studied  for  their  anticancer  properties.  Rugulovasines  A  and   3 B  exhibit  hypotensive  properties,  and  a  nitrogen-­‐containing   butenolide   was   employed   as   a   key   synthetic   intermediate   in   4   the  preparation  of  an  NK1  receptor  antagonist.

Figure   1.   Nitrogen-­‐containing   biologically  important  targets.  

butenolides  

bearing  an  ortho-­‐  hydroxyl  group  and  the  report  was  limited   to  only  a  single  butenolide.    In  fact,  simply  switching  from  α-­‐ angelica   lactone   to   the   regioisomeric   β-­‐angelica   lactone   re-­‐ sulted   in   complete   loss   of   reactivity   (eq.   2).     Shibasaki   has   reported   excellent   work   on   the   related   addition   of   bu-­‐ tenolides   lacking   substitution   at   the   5-­‐position   into   both   6   aldimines  and  ketimines. A   non-­‐direct   vinylogous   Mannich   reaction   of   butenolides   was  reported  earlier  by  Martin  and  Lopez  and  later  improved   7,8 upon   by   Hoveyda   and   Snapper   (eq.   3).     Both   reports   re-­‐ quire   pre-­‐activation   of   the   butenolide   as   the   siloxyfuran   as   well   as   cryogenic   reaction   temperatures.   In   each   of   these   non-­‐direct   Mannich   reactions,   an   N-­‐aryl   imine   bearing   a   chelating  functional  group  on  the  aromatic  ring  was  required   to   obtain   good   enantioselectivities.     Additionally,   the   scope   of  the  butenolide  partner  is  limited;  substitution  is  only      

in  

  Given   the   prevalence   of   butenolides   in   bioactive   targets   and   their   utility   as   synthetic   intermediates,   it   is   surprising   that  there  is  only  one  reported  example  of  5-­‐substituted  bu-­‐ tenolides   used   as   nucleophiles   in   a   direct   asymmetric   Man-­‐ nich  reaction.    In  this  report  by  Feng  et  al,  α-­‐angelica  lactone   is  shown  to  couple  with  aldimines  via  a  chiral  Sc(III)  catalyst   5 (eq.   1).     While   high   selectivities   are   observed,   the   imine   cou-­‐ pling   partner   required   a   specialized   N-­‐aryl   protecting   group  

ACS Paragon Plus Environment

 

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

 tolerated   at   the   5-­‐position   and   is   limited   to   methyl.   In   this   work,   we   disclose   the   first   direct   vinylogous   Mannich   reac-­‐ tion  which  utilizes  easily  deprotected  N-­‐Cbz  imines  and  both   α,β-­‐  and  β,γ-­‐butenolides  bearing  a  variety  of  substituents  and   substitution  patterns  (eq.  4).  

Page 2 of 6

Scheme  1.  Scope  of  imines  with  α-­‐angelica  lactonea-­‐d  

Our  group  has  demonstrated  that  Zn-­‐ProPhenol  complex-­‐ 9 es   are   useful   for   a   variety   of   asymmetric   transformations,   10 including  a  number  of  Mannich  reactions.  Due  to  the  limi-­‐ tations  of  existing  methods  for  vinylogous  Mannich  reactions   involving  butenolides  (vide  supra),  we  wondered  if  ProPhenol   could   overcome   these   issues.   Initial   results   proved   surpris-­‐ ingly   promising.   Using   α-­‐angelica   lactone   and   10   mol%   Zn-­‐ ProPhenol   in   toluene   afforded   the   desired   product   3a   in   64%   yield  and  94%  ee.    Various  solvents  were  screened  to  further   improve   upon   these   results   and   with   the   exception   of   diox-­‐ ane,  all  of  the  solvents  examined  afforded  Mannich  adduct  3a   in   excellent   enantioselectivity   and   >30:1   dr.     THF   gave   the   best  result  (79%  yield,  >99.5%  ee)  and  was  used  for  all  subse-­‐ quent  reactions.    

Table  1.  Optimization  of  the  reaction  conditions    

a

  Reaction   Conditions:   1   equiv   of   lactone,   1.2   equiv   of   imine,   b 10   mol%   of   Zn-­‐ProPhenol   at   rt   in   solvent   (0.3   M)   for   8   h.     c Isolated   yields   are   given.     ee   was   determined   using   HPLC   d 1 analysis.    dr  was  determined  by  crude   H  NMR  .    

and  99%  ee  and  2-­‐furyl  imine  1i  afforded  similar  results,  giv-­‐ ing  3i  in  51%  yield  and  99%  ee.  

b

c

d

entry  

solvent  

yield  

ee  

dr  

1  

Toluene  

64  

94  

>30:1  

2  

DCM  

65  

98  

>30:1  

3  

THF  

79  

>99.5  

>30:1  

4  

Ether  

61  

98  

>30:1  

5  

Dioxane  

16  

34  

4:1  

a  

Reaction   Conditions:   1   equiv   of   lactone,   1.2   equiv   of   imine,   b   10  mol%  of  Zn-­‐ProPhenol  at  rt  in  solvent  (0.3  M)  for  8  h.   c   Isolated  yields  are  given.     ee  was  determined  using  HPLC   d   1 analysis.   dr  was  determined  by  crude   H  NMR.   With   optimized   conditions   in   hand,   a   variety   of   imines   were  evaluated  (Scheme  1).  Substitution  at  the  ortho-­‐,  meta-­‐   and  para-­‐  positions  is  well  tolerated  and  has  little  impact  on   the  yield,  regio-­‐,  diastereo-­‐,  or  enantioselectivity.     Phenyl   imine   1b   produced   the   corresponding   vinylogous   Mannich   adduct   3b   in   76%   yield   with   >99.5   ee,   and   tolyl   imine  1c  afforded  the  desired  product  in  75%  yield  and  >99.5   ee.   Surprisingly,   the   less   electrophilic   4-­‐methoxy   imine   1d   also   gave   excellent   results,   affording   3d   in   91%   yield   and   >99.5%  ee.  1-­‐  and  2-­‐Naphthyl  imines  gave  3e  and  3f  in  98%  ee   and   >99.5%   ee,   respectively,   with   nearly   identical   yields.   In-­‐ troducing   a   sterically   demanding   substituent   at   the   ortho-­‐   position  had  no  effect  on  the  course  of  the  reaction,  yielding   3g in  76%  yield  and  99%  ee.  Notably,  heteroaryl  imines  were   also  well  tolerated;  thiophene  3h  was  obtained  in  69%  yield    

During   our   initial   optimization,   we   observed   that   β-­‐ angelica   lactone   afforded   the   same   results   as   α-­‐angelica   lac-­‐ tone,   albeit   with   slightly   longer   reaction   times.   Given   this   result,  we  were  curious  to  see  whether  other  α,β-­‐butenolides   would   participate   in   the   reaction,   particularly   since   sub-­‐ strates   of   this   type   were   unreactive   in   previously   reported   vinylogous  Mannich  reactions  (vide  supra).  Additionally,  due   to   conjugation   with   the   carbonyl   group,   α,β-­‐butenolides   are   more   easily   synthesized   and   stored   than   the   analogous   β,γ-­‐   compounds.   Under  our  optimized  conditions,  we  are  pleased  to  report   that   a   variety   of   nucleophiles   can   be   utilized.   Commercially   available   furanone   2b   reacted   with   both   electron-­‐deficient   (1a)   and   electron-­‐rich   (1d)   imines,   affording   3ab   and   3db   in   >99.5%  and  >99.5%  ee,  respectively,  with  good  yields.  Nota-­‐ bly,   only   a   single   diastereomer   is   observed   despite   the   pres-­‐ ence   of   a   highly   epimerizable   alpha   proton.   α,β-­‐butenolides   with  sterically  demanding  alkyl  substituents  at  the  5-­‐position   gave   incomplete   conversion   (2c   and   2d),   but   still   afforded   excellent   selectivities   in   good   yields.     (Trimethylsilyl)methyl   butenolide  2c  reacted  with  1d  to  form  3dc  in  99%  ee  and  15:1   dr.   Poor   conversion   for   these   bulky   substrates   could   be   im-­‐ proved   by   doubling   the   catalyst   loading,   with   3ac   being   formed   in   69%   yield,   99.5%   ee,   and   16:1   dr   when   20   mol   %   catalyst  was  used.  An  isobutyl  group  (2d)  was  also  tolerated,   and  3ad  and  3dd  were  obtained  with  excellent  ee  and  slightly   reduced  dr.  Introducing  a  bromo  substituent  adjacent  to  the   nucleophilic   site   had   no   deleterious   effects   on   reactivity   or   selectivity,  and  3ae  was  obtained  in  74%  yield  with  excellent   enantio-­‐  (>99.5%)  and  diastereoselectivity  (18:1).      

ACS Paragon Plus Environment

Page 3 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Scheme  2.  Scope  of  α,β-­‐  and  β,γ-­‐  butenolidesa-­‐d  

  In   addition   to   providing   high-­‐value   products   in   near-­‐ perfect   enantioselectivity,   this   reaction   can   be   easily   per-­‐ formed   on   gram   scale   at   decreased   catalyst   loading   without   impacting   yield   or   selectivity   (eq.   5).   Using   2   mol%   Zn-­‐ ProPhenol,  α-­‐angelica  lactone  reacted  with  benzene  imine  1b   to   afford   3b   in   79%   yield   and   >99.5%   ee.   As   with   the   small   scale   reaction,   the   product   was   obtained   with   >30:1   dr.   It   is   worth   noting   that   the   ProPhenol   catalyst   can   be   recovered   after   the   reaction,   further   reducing   the   effective   catalyst   10   loading.

 

 

a  

Reaction   Conditions:   1   equiv   of   lactone,   1.2   equiv   of   imine,   10   mol%   of   Zn-­‐ProPhenol   at   rt   in   solvent   (0.3   M)   for   over-­‐ b   c   night.   Isolated  yields  are  given.     ee  was  determined  using   d   1 19 HPLC  analysis.   dr  was  determined  by  crude   H  NMR  or   F   e NMR.    20  mol  %  dinuclear  zinc-­‐ProPhenol  used.   We   further   expanded   the   scope   of   the   reaction   to   included   phenyl-­‐   and   thiophenyl-­‐substituted   butenolides.   Unlike   2c   and  2d,  which  had   bulky  alkyl  groups  at  the  5-­‐positon,  5-­‐aryl   butenolides  gave  full  conversion  under  the  optimized  condi-­‐ tion.  α-­‐Phenyl-­‐β-­‐angelica  lactone  gave  3af  in  60%  yield  with   98%   ee,   and   the   analogous   thiophenyl-­‐substituted   com-­‐ pound  gave  3ag  in  62%  yield  and  97%  ee  with  3:1  dr.  Finally,   bicyclo[4.3.0]   lactone   2h   gave   excellent   results   with   4-­‐ fluorophenyl   imine   1a,   producing   3ah   in   85%   yield   and   >99.5%   ee.   To   unambiguously   determine   the   absolute   con-­‐ figuration  of  our  vinylogous  Mannich  products,  we  obtained   a  crystal  structure  of  3ah.  The  configuration  was  determined   to   be   (S,S),   which   corresponds   to   the   syn-­‐   Mannich   adduct.   The   stereochemistry   of   all   other   products   was   assigned   by   analogy.  

Figure 2. ORTEP diagram of 3ah.

Through   judicious   choice   of   the   reaction   conditions,   we   were   able   to   effect   a   variety   of   selective   reduc-­‐ tion/deprotection   reactions   on   adduct   3b.   Treatment   of   3b   with  catalytic  palladium  and  1,4-­‐cyclohexadiene  removed  the   Cbz   group,   liberating   free   amine   4b   in   52%   yield   without   reducing   the   enoate   alkene.   Under   hydrogenation   condi-­‐ tions,   deprotection   of   the   Cbz   group   was   accompanied   by   reduction  of  the  enoate  double  bond,  followed  by  spontane-­‐ ous   cyclization   to   lactam   5b   in   73%   yield,   an   impressive   re-­‐ sult   for   three   transformations   in   one   pot.   This   result   is   par-­‐ ticularly   noteworthy,   since   the   2-­‐alkyl-­‐3-­‐hydroxy   piperidine   unit  is  present  in  a  variety  of  biologically  active  targets,  such   as   neurokinin   substance   P   receptor   antagonist   L-­‐733,060   11       (6).

Scheme  3.  Selective  reductions  of  Mannich  products.  

 

This   reduction-­‐lactamization   cascade   could   provide   rapid   access  to  similar  compounds,  as  well  as  analogs  with  tertiary   alcohols   at   the   3-­‐position.   Finally,   treatment   of   3b   with   12   NaBH4  and  NiCl2 selectively  reduced  the  butenolide  alkene   to  give  saturated  lactone  7b  while  leaving  the  Cbz  group  in-­‐ tact.  

       

 

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme   4.   Further   derivatizations   of   Mannich   prod-­‐ ucts.  

A   Mannich   product   with   an   appropriately   tethered   cin-­‐ namate   ester   was   treated   with   Cs2CO3   to   afford   isoindoline   8g   via   an   intramolecular   aza-­‐Michael   addition   reaction   13 14 (Scheme  4).  Given  the  biologically  importance  and  recent   15   interest   in   the   synthesis   of   isoindoline   derivatives, this   transformation   is   particularly   notable.   Additionally,   our   products   can   be   further   elaborated   using   cross-­‐coupling   chemistry,  as  demonstrated  by  the  Sonogashira  reaction  per-­‐ formed  on  3ae.  

Scheme  5.  Proposed  mechanism.  

Page 4 of 6

tions   between   the   right   hand   diphenylprolinol   unit   and   the   bulk  of  the  ring  system.   In   conclusion,   our   Zn-­‐ProPhenol   system   efficiently   cata-­‐ lyzes   the   direct   addition   of   a   variety   of   substituted   bu-­‐ tenolides   to   various   imines   with   excellent   enantio-­‐   (up   to   >99.5%   ee)   and   diastereoselectivity   (up   to   >30:1   dr).   Addi-­‐ tionally,  this  method  overcomes  many  of  the  hurdles  associ-­‐ ated   with   using   butenolides   as   nucleophiles   in   vinylogous   Mannich   reactions.   This   is   the   first   report   of   such   a   process   16 that  does  not  require  chelating  aromatic  imines.  The  benzyl   carbamates  in  our  products  can  be  easily  cleaved  to  unmask   the  free  amines.  Furthermore,  a  broad  range  of  nucleophiles   are   employed   for   the   first   time;   both   nonactivated   α,β-­‐   and   β,γ-­‐butenolides   are   viable   reaction   partners,   and   alkyl,   aryl,   and   halogen   substitution   is   tolerated.   The   Mannich   adducts   we   obtain   are   densely   functionalized,   and   can   be   further   elaborated  into  a  variety  of  potentially  interesting  molecules.  

ASSOCIATED CONTENT Supporting Information Experimental   procedures,   characterization   data,   NMR   spec-­‐ tra  for  3a-­‐3i,  3ab-­‐3ah,  3db-­‐3dd,  4b,  5b,  7b,  8g,  and  9ae,  and   crystallographic  data.     The  Supporting  Information  is  available  free  of  charge  on  the   ACS  Publications  website.  

AUTHOR INFORMATION Corresponding Author *E-­‐mail:  [email protected]    

Notes The  authors  declare  no  competing  financial  interests.  

ACKNOWLEDGMENT We  thank  the  NSF  (CHE-­‐1360634)  and  the  NIH  (GM-­‐033049)   for  financial  support  of  our  programs.  We  also  acknowledge   Prof.  Allen  Oliver  (University  of  Notre  Dame)  for  X-­‐ray  crys-­‐ tallographic  analysis.    

REFERENCES

 

Based   on   the   configuration   of   our   vinylogous   Mannich   products,   we   propose   the   following   mechanism.   Following   generation   of   the   dinuclear   metal-­‐ligand   complex   (I)   from   ProPhenol   and   diethylzinc,   coordination   and   deprotonation   of   the   butenolide   occurs   to   generate   zinc   dienolate   II.     Based   on   previous   studies   on   other   Zn-­‐ProPhenol   catalyzed   Man-­‐ nich   reactions,   we   propose   two-­‐point   binding   of   the   imine,   giving   rise   to   complex   III,   which   directs   the   addition   of   the   butenolide   to   the   re   face   of   the   imine.   To   explain   the   ob-­‐ served   diastereoselectivity,   we   propose   that   the   butenolide   favors   the   conformation   shown   to   minimize   steric   interac-­‐

1.   Ottow,   E.   A.;   Brinker,   M.;   Teichmann,   T.;   Fritz,   E.;   Kaiser,   W.;   Brosche,   M.;   Kangasjarvi,   J.;   Jiang,   X.;   Polle,   A.   Populus   Euphratica   Displays   Apoplastic   Sodium   Accumulation,   Osmotic   Adjustment   by   Decreases  in  Calcium  and  Soluble  Carbohydrates,  and  Develops  Leaf   Succulence  under  Salt  Stress.  Plant  Physiol.  2005,  139,  1762.     2.  (a)Perez,  M.  T.;  Ayad,  P.;  Maillos,;  Poughon,  V.;  Fahy,  J.;  Ratove-­‐ lomanana-­‐Vidal,   V.   ACS   Med.   Chem.   Lett.   2016,   7,   403.   (b)   Rognan,   D.;   Boulanger,   T.;   Hoffmann,   R.;   Vercauteren,   D.   P.;   Andre,   J.-­‐M.;   Durant,   F.;   Wermuth,   C.-­‐G.   Structure   and   molecular   modeling   of   GABAA  receptor  antagonists.  J.  Med.  Chem.  1992,  35,  1969.  (c)  Beut-­‐ ler,   J.   A.;   Karbon,   E.   W.;   Brubaker,   A.   N.;   Malik,   R.;   Curtis,   D.   R.;   Enna,   S.   J.   Securinine   Alkaloids:   A   new   class   of   GABA   receptor   an-­‐ tagonist.  Brain  Res.  1985,  330,  135.     3.  Abe,  M.;  Ohmomo,  S.;  Ōhashi,  T.;  Tabuchi,  T.  Agr.  Biol.  Chem.   1969,  33,  469.   4.  Raubo,  P.;  Kulagowski,  J.  J.;  Swain,  C.  J.  Synlett  2003,  2021.   5.  Zhou,  L.;  Lin,  L.;  Ji,  J.;  Xie,  M.;  Liu,  X.;  Feng,  X.  Org.  Lett.  2011,   13,  3056.   6.(a)   For   aldimines,   see:   Yamaguchi,   A.;   Matsunaga,   S,   Shibasaki,   M.   Org.   Lett.   2008,   10,   2319-­‐2322.   (b)   For   ketimines   see:   Yin,   L.;  

ACS Paragon Plus Environment

Page 5 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Takada,  H.;  Humagai,  N.;  Shibasaki,  M.  Angew.  Chem.  Int.  Ed.  2013,   52,   7310-­‐7313.   (c)   For   related   ketimine   work,   see:   Nakamura,   S.;   Yamaji,  R.;  Hayashi,  M.  Chem.  Eur.  J.  2015,  21,  9615-­‐9618.   7.  Martin,  S.  F.;  Lopez,  O.  D.  Tetrahedron  Lett.  1999,  40,  8949.   8.   (a)   Carswell,   E.   L.;   Snapper,   M.   L.;   Hoveyda,   A.   H.   Angew.   Chem.,  Int.  Ed.  2006,  45,  7230.    (b)  Mandai,  H.;  Mandai,  K.;  Snapper,   M.  L.;  Hoveyda,  A.  H.  J.  Am.  Chem.  Soc.  2008,  130,  17961.  (c)  Wieland,   L.  C.;  Vieira,  E.  M.;  Snapper,  M.  L.;  Hoveyda,  A.  H.  J.  Am.  Chem.  Soc.   2009,  131,  570.   9.  Trost,  B.  M.;  Bartlett,  M.  J.;  Acc.  Chem.  Res.,  2015,  48,  688.   10.  Trost,  B.  M.;  Hitce,  J.  J.  Am.  Chem.  Soc.,  2009,  131,  4572.   11.   Baker,   R.;   Harrison,   T.;   Hollingworth,   G.   J.;   Swain,   C.   J.;   Wil-­‐ liams,   B.   J.   EP   0   528,   495A1,   1993.   (b)   Harrison,   T.;   Williams,   B.   J.;   Swain,  C.  J.;  Ball,  R.  G.  Bioorg.  Med.  Chem.  Lett.  1994,  4,  2545.   12.  Guo,  Y.  –L.;  Bai,  J,  -­‐F.;  Peng,  L.;  Wang,  L,  -­‐L.;  Jia,  L.  N.;  Luo,  X.   Y.;  Tian,  F.;  Xu,  X.  –Y.;  Wang,  L.  –X.  J.  Org.  Chem.  2012,  77,  8338.   13.  Trost,  B.  M.;  Gnanamani.  E.;  Hung,  C.-­‐I.  Angew.  Chem.  Int.  Ed.   2017,  56,  10451.   14.  a)  Leonard,  M.  S.  ARKIVOC  2013,  1.  (b)  Kukkola,  P.  J.;  Bilci,  N.   A.;   Ikler,   T.;   Savage,   P.;   Shetty,   S.   S.;   DelGrande,   D.;   Jeng,   A.   Y.   Bioorg.  Med.  Chem.  Lett.  2001,  11,  1737  (c)  Portevin,  B.;  Tordjman,  C.;   Pastoureau,   P.;   Bonnet,   J.;   De   Nanteuil,   G.   J.   Med.   Chem.   2000,   43,   4582.  (d)  Ewing,  D.  F.;  Len,  C.;  Mackenzie,  G.;  Petit,  J.  P.;  Ronco,  G.;   Villa,  P.  J.   Pharm.   Pharmacol.  2001,  53,  945.  (e)  Stuk,  T.  L.;  Assink,  B.   K.;   Bates,   R.   C.;   Erdman,   D.   T.;   Fedij,   V.;   Jennings,   S.   M.;   Lassig,   J.   A.;   Smith,  R.  J.;  Smith,  T.  L.  Org.  Process  Res.  Dev.  2003,  7,  851.  (f)  Esti-­‐ arte,  M.  A.;  Johnson,  R.  J.;  Kaub,  C.  J.;  Gowlugari,  S.;  O’Mahony,  D.  J.   R.;  Nguyen,  M.  T.;  Emerling,  D.  E.;  Kelly,  M.  G.;  Kincaid,  J.;  Vincent,   F.;  Duncton,  M.  A.  J.  MedChemComm  2012,  3,  611.   15.Takizawa,  S.;  Sako,  M.;  Abozeid,  M.  A.;  Kishi,  K.;  Watsala,  H.  D.   P.;   Hirata,   S.;   Murai,     K;   Fujioka,   H.;   Sasai.   H.   Org.   Lett.   2017,   19,   5426.   16.   At   present   it   appears   that   this   process   is   limited   to   aromatic   and  heteroaromatic  imines.  

Graphical diagram:

 

 

 

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 6

 

ACS Paragon Plus Environment

6