mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical...

129
Mechanochemical scission of transition metal-ligand bonds in coordination polymers Citation for published version (APA): Balan, A. (2015). Mechanochemical scission of transition metal-ligand bonds in coordination polymers. Technische Universiteit Eindhoven. Document status and date: Published: 11/11/2015 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 12. Sep. 2020

Upload: others

Post on 20-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Mechanochemical scission of transition metal-ligand bonds incoordination polymersCitation for published version (APA):Balan, A. (2015). Mechanochemical scission of transition metal-ligand bonds in coordination polymers.Technische Universiteit Eindhoven.

Document status and date:Published: 11/11/2015

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 12. Sep. 2020

Page 2: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Mechanochemical  Scission  of  Transition  Metal-­‐Ligand  Bonds  in  Coordination  Polymers  

 

 

 

PROEFSCHRIFT  

 

 

 

ter  verkrijging  van  de  graad  van  doctor  aan  de  Technische  Universiteit  Eindhoven,  op  gezag  van  de  rector  magnificus  prof.dr.ir.  F.P.T.  Baaijens,  voor  een  commissie  aangewezen  door  het  College  voor  Promoties,  in  het  openbaar  te  verdedigen  op  woensdag  11  november  2015  om  16:00  uur  

 

 

door    

 

Abidin  Balan  

 

 

geboren  te  Malatya,  Turkije    

Page 3: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Dit  proefschrift  is  goedgekeurd  door  de  promotoren  en  de  samenstelling  van  de  promotiecommissie  is  als  volgt:      voorzitter:       prof.dr.ir.  J.C.  Schouten  1e  promotor:           prof.dr.  R.P.  Sijbesma    2e  promotor:     prof.dr.  E.W.  Meijer  leden:         prof.dr.  J.H.N.  Reek  (UvA)           prof.dr.  G.  Cravotto    (Politecnico  di  Torino)             prof.dr.ir.  E.J.M.  Hensen                   prof.dr.  A.P.H.J.  Schenning      

 

 

 

 

 

Het  onderzoek  of  ontwerp  dat  in  dit  proefschrift  wordt  beschreven  is  uitgevoerd  in  overeenstemming  met  de  TU/e  Gedragscode  Wetenschapsbeoefening.  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

To  my  wife,  

Gizem’e  

 

                   

Imagination  will  often  carry  us  to  worlds  that  never  were,  but  without  it  we  go  nowhere.    

(Carl  Sagan)    

 

The  highest  activity  a  human  being  can  attain  is  learning  for  understanding,                                                because  to  understand  is  to  be  free.  

(Baruch  Spinoza)  

 

 

 

 

 

 

Page 4: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Dit  proefschrift  is  goedgekeurd  door  de  promotoren  en  de  samenstelling  van  de  promotiecommissie  is  als  volgt:      voorzitter:       prof.dr.ir.  J.C.  Schouten  1e  promotor:           prof.dr.  R.P.  Sijbesma    2e  promotor:     prof.dr.  E.W.  Meijer  leden:         prof.dr.  J.H.N.  Reek  (UvA)           prof.dr.  G.  Cravotto    (Politecnico  di  Torino)             prof.dr.ir.  E.J.M.  Hensen                   prof.dr.  A.P.H.J.  Schenning      

 

 

 

 

 

Het  onderzoek  of  ontwerp  dat  in  dit  proefschrift  wordt  beschreven  is  uitgevoerd  in  overeenstemming  met  de  TU/e  Gedragscode  Wetenschapsbeoefening.  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

To  my  wife,  

Gizem’e  

 

                   

Imagination  will  often  carry  us  to  worlds  that  never  were,  but  without  it  we  go  nowhere.    

(Carl  Sagan)    

 

The  highest  activity  a  human  being  can  attain  is  learning  for  understanding,                                                because  to  understand  is  to  be  free.  

(Baruch  Spinoza)  

 

 

 

 

 

 

Page 5: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

 

 

 

 

Abidin  Balan  

Mechanochemical  Scission  of  Transition  Metal-­‐Ligand  Bonds  in  Coordination  Polymers  

Eindhoven  University  of  Technology,  2015  

 

 

Printing:  Ridderprint  BV,  the  Netherlands  

Cover:  ‘’Jardine  Media’’,  Ridderprint  BV    

 

 

A  catalogue  record  is  available  from  the  Eindhoven  University  of  Technology  Library  

ISBN:  978-­‐90-­‐386-­‐3946-­‐8  

 

This  work  has  been  financially  supported  by  the  Netherlands  Organization  for  Scientific  Research,  Chemical  Sciences  (NWO-­‐CW).  

 

Copyright  ©  2015  by  Abidin  Balan  

 

 

 

 

 

 

 

 

Table  of  Contents  

Chapter  1  Introduction:  Polymer  Mechanochemistry  ..............................................................................  1  

History  and  background  .................................................................................................................  2  Mechanophores  .............................................................................................................................  7  Mechanoresponsive  materials  .......................................................................................................  9  Mechanical  release  of  small  molecules  in  polymer  matrices  .......................................................  11  Mechanochemical  catalysis  ..........................................................................................................  13  Aim  and  outline  of  this  thesis  .......................................................................................................  16  References  ....................................................................................................................................  18  

Chapter  2  Mechanochemical  chain  scission  in  NHC-­‐Pd  centered  coordination  polymers  .......................  21  

Introduction  ..................................................................................................................................  22  Results  and  Discussions  ................................................................................................................  24  Synthesis  .......................................................................................................................................  24  Chain  Scission  in  Pd(NHC-­‐pTHF)2Cl2  .............................................................................................  26  Determination  of  limiting  molecular  weight  Mlim  .........................................................................  29  Scission  Rates  for  M(NHC-­‐pTHF)2Cl2  .............................................................................................  32  Chain  scission  mechanism  ............................................................................................................  39  Conclusions  ...................................................................................................................................  43  Experimental  ................................................................................................................................  44  References  ....................................................................................................................................  46  

Chapter  3  Mechanical  scission  of  Pd(NHC)2Cl2  complexes  probed  with  chemiluminescence  ..................  49  

Introduction  ..................................................................................................................................  50  2-­‐Coumaranones  ..........................................................................................................................  51  Results  and  discussions  ................................................................................................................  52  Chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  complexes  in  the  presence  of  coumaranone  ......................  52  Light  emission  ...............................................................................................................................  58  Conclusions  ...................................................................................................................................  63  Experimental  ................................................................................................................................  64  References  ....................................................................................................................................  66  

 

 

 

 

Page 6: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

 

 

 

 

Abidin  Balan  

Mechanochemical  Scission  of  Transition  Metal-­‐Ligand  Bonds  in  Coordination  Polymers  

Eindhoven  University  of  Technology,  2015  

 

 

Printing:  Ridderprint  BV,  the  Netherlands  

Cover:  ‘’Jardine  Media’’,  Ridderprint  BV    

 

 

A  catalogue  record  is  available  from  the  Eindhoven  University  of  Technology  Library  

ISBN:  978-­‐90-­‐386-­‐3946-­‐8  

 

This  work  has  been  financially  supported  by  the  Netherlands  Organization  for  Scientific  Research,  Chemical  Sciences  (NWO-­‐CW).  

 

Copyright  ©  2015  by  Abidin  Balan  

 

 

 

 

 

 

 

 

Table  of  Contents  

Chapter  1  Introduction:  Polymer  Mechanochemistry  ..............................................................................  1  

History  and  background  .................................................................................................................  2  Mechanophores  .............................................................................................................................  7  Mechanoresponsive  materials  .......................................................................................................  9  Mechanical  release  of  small  molecules  in  polymer  matrices  .......................................................  11  Mechanochemical  catalysis  ..........................................................................................................  13  Aim  and  outline  of  this  thesis  .......................................................................................................  16  References  ....................................................................................................................................  18  

Chapter  2  Mechanochemical  chain  scission  in  NHC-­‐Pd  centered  coordination  polymers  .......................  21  

Introduction  ..................................................................................................................................  22  Results  and  Discussions  ................................................................................................................  24  Synthesis  .......................................................................................................................................  24  Chain  Scission  in  Pd(NHC-­‐pTHF)2Cl2  .............................................................................................  26  Determination  of  limiting  molecular  weight  Mlim  .........................................................................  29  Scission  Rates  for  M(NHC-­‐pTHF)2Cl2  .............................................................................................  32  Chain  scission  mechanism  ............................................................................................................  39  Conclusions  ...................................................................................................................................  43  Experimental  ................................................................................................................................  44  References  ....................................................................................................................................  46  

Chapter  3  Mechanical  scission  of  Pd(NHC)2Cl2  complexes  probed  with  chemiluminescence  ..................  49  

Introduction  ..................................................................................................................................  50  2-­‐Coumaranones  ..........................................................................................................................  51  Results  and  discussions  ................................................................................................................  52  Chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  complexes  in  the  presence  of  coumaranone  ......................  52  Light  emission  ...............................................................................................................................  58  Conclusions  ...................................................................................................................................  63  Experimental  ................................................................................................................................  64  References  ....................................................................................................................................  66  

 

 

 

 

Page 7: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter  4  Determination  of  ligand  exchange  dynamics  and  sonication  induced  ligand  exchange  rate  in  Imidazole-­‐Pd  centered  coordination  polymers  ..................................................................  69  

Introduction  ..................................................................................................................................  70  Results  and  Discussions  ................................................................................................................  73  Synthesis  of  Pd-­‐Imidazole  complexes  ...........................................................................................  73  Ligand  exchange  between  EtIm  and  Pd(EtIm)2Cl2  ........................................................................  75  Sonication  induced  ligand  exchange  ............................................................................................  79  Rate  of  sonication  induced  chain  scission  ....................................................................................  80  Conclusions  ...................................................................................................................................  85  Experimental  ................................................................................................................................  86  References  ....................................................................................................................................  88  

Chapter  5  Mechanochemically  induced,  directed  ligand  exchange  in  polymeric  Pd(II)  complexes  .........  89  

Introduction  ..................................................................................................................................  90  Results  and  discussions  ................................................................................................................  92  Mechanochemical  synthesis  of  hetero-­‐complexes  ......................................................................  95  Mechanochemical  synthesis  of  block  copolymers  .......................................................................  96  Conclusions  ...................................................................................................................................  99  Experimental  ..............................................................................................................................  100  References  ..................................................................................................................................  102  

Chapter  6  Transition  metal  bearing  supramolecular  polymer  networks:  Towards  self-­‐healing  applications  ........................................................................................................................  103  

Introduction  ................................................................................................................................  104  Results  and  discussions  ..............................................................................................................  106  Conclusions  .................................................................................................................................  109  Experimental  ..............................................................................................................................  110  References  ..................................................................................................................................  111  

Summary  ............................................................................................................................  113  

Curriculum  Vitae  .................................................................................................................  117  

Acknowledgements  ............................................................................................................  119  

 

Page 8: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter 1Introduction: Polymer Mechanochemistry

The past 10 years have seen a resurgence of interest in the field of polymer mechanochemis-

try. Whilst the destructive effects of mechanical force on polymer chains have been known

for decades, it was only recently that researchers tapped into these forces to realize more

useful chemical transformations. Current developments in mechanochemistry involve the

introduction of mechanophores in polymer chains that can undergo useful reactions upon

mechanical activation. History and fundamental aspects underlying mechanochemical chain

scission in polymers together with elegant examples from recent literatures were described

in this chapter. At the end the aim and outline of this thesis are given.

Chapter  4  Determination  of  ligand  exchange  dynamics  and  sonication  induced  ligand  exchange  rate  in  Imidazole-­‐Pd  centered  coordination  polymers  ..................................................................  69  

Introduction  ..................................................................................................................................  70  Results  and  Discussions  ................................................................................................................  73  Synthesis  of  Pd-­‐Imidazole  complexes  ...........................................................................................  73  Ligand  exchange  between  EtIm  and  Pd(EtIm)2Cl2  ........................................................................  75  Sonication  induced  ligand  exchange  ............................................................................................  79  Rate  of  sonication  induced  chain  scission  ....................................................................................  80  Conclusions  ...................................................................................................................................  85  Experimental  ................................................................................................................................  86  References  ....................................................................................................................................  88  

Chapter  5  Mechanochemically  induced,  directed  ligand  exchange  in  polymeric  Pd(II)  complexes  .........  89  

Introduction  ..................................................................................................................................  90  Results  and  discussions  ................................................................................................................  92  Mechanochemical  synthesis  of  hetero-­‐complexes  ......................................................................  95  Mechanochemical  synthesis  of  block  copolymers  .......................................................................  96  Conclusions  ...................................................................................................................................  99  Experimental  ..............................................................................................................................  100  References  ..................................................................................................................................  102  

Chapter  6  Transition  metal  bearing  supramolecular  polymer  networks:  Towards  self-­‐healing  applications  ........................................................................................................................  103  

Introduction  ................................................................................................................................  104  Results  and  discussions  ..............................................................................................................  106  Conclusions  .................................................................................................................................  109  Experimental  ..............................................................................................................................  110  References  ..................................................................................................................................  111  

Summary  ............................................................................................................................  113  

Curriculum  Vitae  .................................................................................................................  117  

Acknowledgements  ............................................................................................................  119  

 

Page 9: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

History  and  background  

Mechanical   activation   of   chemical   bonds   offers   an   alternative   to   conventional   processes  

such   as   thermal   or   photochemical   activation.1   As   stated   by   Kauzmann   and   Eyring,2  

mechanical  work  done  by  an  external  force  (W(r)  =  F  ×  r)  modifies  the  Morse  potential  (U’(r)  

=  U(r)   –  W(r))   and   lowers   the  energy  barrier   for  bond  dissociation   to   such  an  extent   that  

thermal   fluctuations   can   exceed   this   barrier   at   room   temperature,   in   contrast   to   the  

unstretched  bond.3,4    

 

Figure  1:  Morse  energy  potentials  representing  a  chemical  bond  in  unperturbed  state  (dash-­‐dotted  

line)   and   of   a   bond   under   stress   (solid   line)   showing   the   mechanochemical   activation   of   bond  

according  to  the  TABS  theory.  The  energy  input  due  to  mechanical  work  is  shown  as  a  dashed  line.4  

Polymer   mechanochemistry   was   first   described   by   Staudinger   in   the   1930s,   when   he  

interpreted   the   decrease   in   molecular   weight   upon   mastication   of   a   polymer   as   the  

mechanical   rupture  of  macromolecules.5   Since   then,   the  use  of   force   to  activate   chemical  

bonds   and   to   initiate   chemical   reactions   in   polymers   has   become  of   interest   because   the  

reaction  pathways  and  outcome  of  a  mechanochemical  reaction  can  be  completely  different  

from   its   thermal   analogue.   For   instance,   thermal   ring-­‐opening   reaction   of   cyclobutanes  

(CBs),   an  orbital-­‐controlled   reaction,   is  only  allowed  via  a   conrotatory  pathway,  yielding  a  

different   isomer  of   the   ring-­‐opened  product   (E,Z  or  E,E)  depending  on   the  starting   isomer  

(cis-­‐  or  trans-­‐CB,  respectively).  However,  when  activated  by  mechanical  force,  regardless  of  

the   starting   isomer,   the   ring-­‐opening   always   yields   the   E,E   isomer   as   a   product.   This  

intriguing  difference  between  the  outcomes  of  mechanical  and  thermal  reactions  is  a  result  

of   the   fact   that   mechanical   work   involved   with   force,   in   contrast   to   thermal   energy,   is  

anisotropic  (i.e.  force  has  a  direction).6–10  

 

 

Figure   2:   Electrocyclic   ring   opening   of   benzocyclobutenes.   The   intermediate   resulting   from   the  

electrocyclic   ring   opening  of   benzocyclobutenes   depends  on   the  method  of   activation   and  on   the  

geometry  of  the  molecule.6    

Nature  also  uses  mechanochemistry,  for   instance,   in  the  regulation  of  the  activity  of  some  

enzymes  and  proteins  by  force-­‐induced  structural  changes.  Proteins  fold   in  many  different  

conformations,  which  changes  the  activity  of  enzymes  associated  to  them.  The  free  energy  

difference  between   folded  and  unfolded  states   is  usually   lower   than  energy  of  a   covalent  

bond  and  this  makes  it  possible  to  ‘force’  the  protein  into  a  different  conformation  and  alter  

its  enzymatic  activity  even  with   low  stress   loadings.11  Klibanov  et  al.  have   investigated  the  

effect  of  mechanical  stretching  on  enzyme  activity  by  covalently  attaching  chymotrypsin  and  

Chapter 1

2

Page 10: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

History  and  background  

Mechanical   activation   of   chemical   bonds   offers   an   alternative   to   conventional   processes  

such   as   thermal   or   photochemical   activation.1   As   stated   by   Kauzmann   and   Eyring,2  

mechanical  work  done  by  an  external  force  (W(r)  =  F  ×  r)  modifies  the  Morse  potential  (U’(r)  

=  U(r)   –  W(r))   and   lowers   the  energy  barrier   for  bond  dissociation   to   such  an  extent   that  

thermal   fluctuations   can   exceed   this   barrier   at   room   temperature,   in   contrast   to   the  

unstretched  bond.3,4    

 

Figure  1:  Morse  energy  potentials  representing  a  chemical  bond  in  unperturbed  state  (dash-­‐dotted  

line)   and   of   a   bond   under   stress   (solid   line)   showing   the   mechanochemical   activation   of   bond  

according  to  the  TABS  theory.  The  energy  input  due  to  mechanical  work  is  shown  as  a  dashed  line.4  

Polymer   mechanochemistry   was   first   described   by   Staudinger   in   the   1930s,   when   he  

interpreted   the   decrease   in   molecular   weight   upon   mastication   of   a   polymer   as   the  

mechanical   rupture  of  macromolecules.5   Since   then,   the  use  of   force   to  activate   chemical  

bonds   and   to   initiate   chemical   reactions   in   polymers   has   become  of   interest   because   the  

reaction  pathways  and  outcome  of  a  mechanochemical  reaction  can  be  completely  different  

from   its   thermal   analogue.   For   instance,   thermal   ring-­‐opening   reaction   of   cyclobutanes  

(CBs),   an  orbital-­‐controlled   reaction,   is  only  allowed  via  a   conrotatory  pathway,  yielding  a  

different   isomer  of   the   ring-­‐opened  product   (E,Z  or  E,E)  depending  on   the  starting   isomer  

(cis-­‐  or  trans-­‐CB,  respectively).  However,  when  activated  by  mechanical  force,  regardless  of  

the   starting   isomer,   the   ring-­‐opening   always   yields   the   E,E   isomer   as   a   product.   This  

intriguing  difference  between  the  outcomes  of  mechanical  and  thermal  reactions  is  a  result  

of   the   fact   that   mechanical   work   involved   with   force,   in   contrast   to   thermal   energy,   is  

anisotropic  (i.e.  force  has  a  direction).6–10  

 

 

Figure   2:   Electrocyclic   ring   opening   of   benzocyclobutenes.   The   intermediate   resulting   from   the  

electrocyclic   ring   opening  of   benzocyclobutenes   depends  on   the  method  of   activation   and  on   the  

geometry  of  the  molecule.6    

Nature  also  uses  mechanochemistry,  for   instance,   in  the  regulation  of  the  activity  of  some  

enzymes  and  proteins  by  force-­‐induced  structural  changes.  Proteins  fold   in  many  different  

conformations,  which  changes  the  activity  of  enzymes  associated  to  them.  The  free  energy  

difference  between   folded  and  unfolded  states   is  usually   lower   than  energy  of  a   covalent  

bond  and  this  makes  it  possible  to  ‘force’  the  protein  into  a  different  conformation  and  alter  

its  enzymatic  activity  even  with   low  stress   loadings.11  Klibanov  et  al.  have   investigated  the  

effect  of  mechanical  stretching  on  enzyme  activity  by  covalently  attaching  chymotrypsin  and  

Polymer Mechanochemistry

3

1

Page 11: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

trypsin   to   nylon,   human   hair,   and   viscose   fibers.12   They   showed   that   stretching   of   these  

polymer  supports  induced  a  deformation  on  protein  molecules  and  so  a  3-­‐fold  decrease  in  

enzyme   activity.   Upon   relaxation,   the   initial   level   of   activity  was   reached   instantaneously  

revealing  that  enzyme  activity  can  be  altered  reversibly  by  stretch-­‐relax  process.  

 

Figure  3:  Dependence  of  relative  activity  of  enzymes  covalently  bound  to  different  elastic  supports  

on   the   degree   of   stretching   of   the   supports.   (a)   chymotrypsin   on   protein   coated   nylon   fiber;   (b)  

trypsin  on  human  hair;  (c)  chymotrypsin  on  viscose  (cellulose)  fiber.    Schematic  representation  of  the  

deformation  of  enzyme  bound  to  a  mechanically  stretched  elastic  fiber.12  

One   of   the  most   efficient   ways   to   exert   force   on   a   polymer   in   solution   is   by   the   use   of  

sonication  since  the  strain  rates  accessible  with  sonochemistry  are  greater  than  those  with  

other   solution-­‐based   flow   techniques   such   as   opposed   jets   or   cross   slots,   allowing  

mechanical  activation  to  be  obtained  in  polymers  of  lower  molecular  weight  and  at  greater  

scission  rates.13  Upon  sonication  of  solution,  cavitation  leads  to  strong  elongational  stresses  

around  collapsing  bubbles.14  The  part  of  the  polymer  chain  closest  to  the  collapsing  bubble  

wall  is  pulled  at  a  higher  velocity  than  the  far  end,  and  this  velocity  gradient  creates  stress  

along  the  backbone.  This  force  is  accumulated  through  the  polymer  backbone  and  reaches  a  

maximum  value  at   the  center  of   the  chain  because  the   flow  field   is  centrosymmetric  with  

respect  to  the  molecule.  As  a  consequence  chain  scission  occurs  preferentially  around  the  

chain  midpoint.15    

 

 

Figure   4:  Mechanism   for  ultrasound-­‐induced  polymer   chain   scission:   (a)   gradual  bubble   formation  

results   from  pressure  variations   induced  by   the  acoustic   field;   (b)   rapid  bubble   collapse  generates  

solvodynamic  shear;  (c)  small  molecules  undergo  pyrolytic  cleavage  to  form  radical  byproducts  upon  

bubble   collapse,   while   polymer   chains   do   not   undergo   pyrolytic   cleavage   because   they   do   not  

penetrate  the  bubble  interface.13  

Implosion  of  cavitation  bubbles  is  essentially  an  adiabatic  process,  which  leads  to  formation  

of  local  hotspots  within  the  bubble  in  which  temperature  and  pressure  increase  drastically.  

The  content  of   cavitation  bubble  pyrolyzes  under   these  extreme  conditions  and   results   in  

formation   of   reactive   species,   such   as   radicals   and   persistent,   protic   secondary  

byproducts.16.  Recent  studies  in  our  group  have  shown  that  heat  capacity  of  gas  dissolved  in  

solution   influences   the   formation   of   sonochemical   impurities.   For   instance,   the   use   of  

methane   (CH4)   instead   of   argon   (Ar)   decreases   the   production   of   radicals   significantly.17  

Higher  heat  capacity  and  possible  energy  dissipation  due  to  increased  degrees  of  freedom  in  

CH4,   compared   to   those   of   monoatomic   Ar,   decrease   the   temperature   in   hotspots   and  

suppress  the  reactive  impurity  formation.  On  the  other  hand,  solubility  of  the  saturation  gas  

may  also  have  a  negative   influence  on  ultrasound   induced  mechanical   chain   scission.   Iso-­‐

Chapter 1

4

Page 12: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

trypsin   to   nylon,   human   hair,   and   viscose   fibers.12   They   showed   that   stretching   of   these  

polymer  supports  induced  a  deformation  on  protein  molecules  and  so  a  3-­‐fold  decrease  in  

enzyme   activity.   Upon   relaxation,   the   initial   level   of   activity  was   reached   instantaneously  

revealing  that  enzyme  activity  can  be  altered  reversibly  by  stretch-­‐relax  process.  

 

Figure  3:  Dependence  of  relative  activity  of  enzymes  covalently  bound  to  different  elastic  supports  

on   the   degree   of   stretching   of   the   supports.   (a)   chymotrypsin   on   protein   coated   nylon   fiber;   (b)  

trypsin  on  human  hair;  (c)  chymotrypsin  on  viscose  (cellulose)  fiber.    Schematic  representation  of  the  

deformation  of  enzyme  bound  to  a  mechanically  stretched  elastic  fiber.12  

One   of   the  most   efficient   ways   to   exert   force   on   a   polymer   in   solution   is   by   the   use   of  

sonication  since  the  strain  rates  accessible  with  sonochemistry  are  greater  than  those  with  

other   solution-­‐based   flow   techniques   such   as   opposed   jets   or   cross   slots,   allowing  

mechanical  activation  to  be  obtained  in  polymers  of  lower  molecular  weight  and  at  greater  

scission  rates.13  Upon  sonication  of  solution,  cavitation  leads  to  strong  elongational  stresses  

around  collapsing  bubbles.14  The  part  of  the  polymer  chain  closest  to  the  collapsing  bubble  

wall  is  pulled  at  a  higher  velocity  than  the  far  end,  and  this  velocity  gradient  creates  stress  

along  the  backbone.  This  force  is  accumulated  through  the  polymer  backbone  and  reaches  a  

maximum  value  at   the  center  of   the  chain  because  the   flow  field   is  centrosymmetric  with  

respect  to  the  molecule.  As  a  consequence  chain  scission  occurs  preferentially  around  the  

chain  midpoint.15    

 

 

Figure   4:  Mechanism   for  ultrasound-­‐induced  polymer   chain   scission:   (a)   gradual  bubble   formation  

results   from  pressure  variations   induced  by   the  acoustic   field;   (b)   rapid  bubble   collapse  generates  

solvodynamic  shear;  (c)  small  molecules  undergo  pyrolytic  cleavage  to  form  radical  byproducts  upon  

bubble   collapse,   while   polymer   chains   do   not   undergo   pyrolytic   cleavage   because   they   do   not  

penetrate  the  bubble  interface.13  

Implosion  of  cavitation  bubbles  is  essentially  an  adiabatic  process,  which  leads  to  formation  

of  local  hotspots  within  the  bubble  in  which  temperature  and  pressure  increase  drastically.  

The  content  of   cavitation  bubble  pyrolyzes  under   these  extreme  conditions  and   results   in  

formation   of   reactive   species,   such   as   radicals   and   persistent,   protic   secondary  

byproducts.16.  Recent  studies  in  our  group  have  shown  that  heat  capacity  of  gas  dissolved  in  

solution   influences   the   formation   of   sonochemical   impurities.   For   instance,   the   use   of  

methane   (CH4)   instead   of   argon   (Ar)   decreases   the   production   of   radicals   significantly.17  

Higher  heat  capacity  and  possible  energy  dissipation  due  to  increased  degrees  of  freedom  in  

CH4,   compared   to   those   of   monoatomic   Ar,   decrease   the   temperature   in   hotspots   and  

suppress  the  reactive  impurity  formation.  On  the  other  hand,  solubility  of  the  saturation  gas  

may  also  have  a  negative   influence  on  ultrasound   induced  mechanical   chain   scission.   Iso-­‐

Polymer Mechanochemistry

5

1

Page 13: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

butane,  for  instance,  a  gas  with  higher  solubility  decreases  the  intensity  of  cavitation  effects  

so   it   leads   to   lower   scission   rates.   Therefore,   the   selection  of   saturation   gas   is   crucial   for  

both  scission  rate  and  sonochemical  impurity  formation.  

 

Figure   5:   Radical   production   rate   and   percentage   scission   of   supramolecular   polymer   complex  

Ag(NHC7k)2PF6  after  5  min  of  sonication  using  argon,  nitrogen,  methane  and  isobutane  as  saturation  

gases.17  

It   is   well   established   that   mechanochemical   scission   of   polymers   only   occurs   above   a  

molecular   weight   threshold   (Mlim).18   Below   Mlim,   no   scission   takes   place   since   polymer  

chains  are  too  short  to  accumulate  the  force  required  to  break  chemical  bonds.19  The  rate  of  

mechanically   induced  scission   in  covalent  polymers  has  been  shown  to  be  proportional   to  

[MW   -­‐   Mlim]   when   the   initial   molecular   weight   is   greater   than   Mlim.4,20   Madras   recently  

reported   that   the   chain   scission   rate   coefficient   ksc   is   linearly   dependent   on   the   initial  

molecular  weight  of  polymers  based  on  experimental  data  for  ultrasonic  chain  scission  of  a  

number  of  polymers,  including  poly(ethylene  oxide)  (PEO),  polyacrylamide  (PAa),  poly(butyl  

acrylate)  (PBA),  and  poly(methyl  acrylate)  (PMA).  ksc  is  assumed  to  follow  empirical  formula;    

ksc  =  kd  (MW  -­‐  Mlim  )λ    

where  kd    is  the  degradation  coefficient,  MW  is  the  initial  molecular  weight  of  the  polymer,  

Mlim    is  the  limiting  molecular  weight,  and  the  exponent  λ  is  a  value  between  0  and  3.  When  

λ  =  1,  the  experimental  data  was  fitted  to  the  theory  by  linear  regression.21    

Mechanophores  

Limiting   molecular   weight   (Mlim)   and   the   rate   of   chain   scission   (ksc)   are   influenced  

significantly  if  a  mechanically  responsive  labile  bond,  mechanophore,  is  incorporated  into  a  

polymer  chain.  Encina  et.  al.   found   that   the  ksc  of  poly(vinylpyrrolidone)   increased  10-­‐fold  

when   peroxide   linkages   were   randomly   incorporated   into   the   polymer   backbone.   The  

peroxide  bond  with  bond  dissociation  energy  of  51  kcal/mol   is  much  weaker  than  the  C-­‐C  

(88   kcal/mol)   and   the   C-­‐O   (91   kcal/mol)   bonds   and   makes   the   polymer   susceptible   to  

mechanical  energy  input.22    

Furthermore,  the  polymer  with  a  mechanophore  can  be  degraded  at  specific  locations  since  

mechanical   energy   is   selective   for   the   weakest   bond   on   mechanophore.   Moore   and   co-­‐

workers  synthesized  a  linear  PEG  with  a  single  weak  azo  link  positioned  at  the  center  of  the  

chain.   They   demonstrated   that   the   mechanically   induced   cleavage   was   localized   almost  

exclusively   to   the  homolytic  extrusion  of  nitrogen   from  the  azo  group  since   the  energy  of  

activation  for  this  process  is  24-­‐  30  kcal/mol,  much  lower  than  that  of  the  C-­‐C  and  C-­‐O  bonds  

in  the  PEG  backbone.23  

Chapter 1

6

Page 14: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

butane,  for  instance,  a  gas  with  higher  solubility  decreases  the  intensity  of  cavitation  effects  

so   it   leads   to   lower   scission   rates.   Therefore,   the   selection  of   saturation   gas   is   crucial   for  

both  scission  rate  and  sonochemical  impurity  formation.  

 

Figure   5:   Radical   production   rate   and   percentage   scission   of   supramolecular   polymer   complex  

Ag(NHC7k)2PF6  after  5  min  of  sonication  using  argon,  nitrogen,  methane  and  isobutane  as  saturation  

gases.17  

It   is   well   established   that   mechanochemical   scission   of   polymers   only   occurs   above   a  

molecular   weight   threshold   (Mlim).18   Below   Mlim,   no   scission   takes   place   since   polymer  

chains  are  too  short  to  accumulate  the  force  required  to  break  chemical  bonds.19  The  rate  of  

mechanically   induced  scission   in  covalent  polymers  has  been  shown  to  be  proportional   to  

[MW   -­‐   Mlim]   when   the   initial   molecular   weight   is   greater   than   Mlim.4,20   Madras   recently  

reported   that   the   chain   scission   rate   coefficient   ksc   is   linearly   dependent   on   the   initial  

molecular  weight  of  polymers  based  on  experimental  data  for  ultrasonic  chain  scission  of  a  

number  of  polymers,  including  poly(ethylene  oxide)  (PEO),  polyacrylamide  (PAa),  poly(butyl  

acrylate)  (PBA),  and  poly(methyl  acrylate)  (PMA).  ksc  is  assumed  to  follow  empirical  formula;    

ksc  =  kd  (MW  -­‐  Mlim  )λ    

where  kd    is  the  degradation  coefficient,  MW  is  the  initial  molecular  weight  of  the  polymer,  

Mlim    is  the  limiting  molecular  weight,  and  the  exponent  λ  is  a  value  between  0  and  3.  When  

λ  =  1,  the  experimental  data  was  fitted  to  the  theory  by  linear  regression.21    

Mechanophores  

Limiting   molecular   weight   (Mlim)   and   the   rate   of   chain   scission   (ksc)   are   influenced  

significantly  if  a  mechanically  responsive  labile  bond,  mechanophore,  is  incorporated  into  a  

polymer  chain.  Encina  et.  al.   found   that   the  ksc  of  poly(vinylpyrrolidone)   increased  10-­‐fold  

when   peroxide   linkages   were   randomly   incorporated   into   the   polymer   backbone.   The  

peroxide  bond  with  bond  dissociation  energy  of  51  kcal/mol   is  much  weaker  than  the  C-­‐C  

(88   kcal/mol)   and   the   C-­‐O   (91   kcal/mol)   bonds   and   makes   the   polymer   susceptible   to  

mechanical  energy  input.22    

Furthermore,  the  polymer  with  a  mechanophore  can  be  degraded  at  specific  locations  since  

mechanical   energy   is   selective   for   the   weakest   bond   on   mechanophore.   Moore   and   co-­‐

workers  synthesized  a  linear  PEG  with  a  single  weak  azo  link  positioned  at  the  center  of  the  

chain.   They   demonstrated   that   the   mechanically   induced   cleavage   was   localized   almost  

exclusively   to   the  homolytic  extrusion  of  nitrogen   from  the  azo  group  since   the  energy  of  

activation  for  this  process  is  24-­‐  30  kcal/mol,  much  lower  than  that  of  the  C-­‐C  and  C-­‐O  bonds  

in  the  PEG  backbone.23  

Polymer Mechanochemistry

7

1

Page 15: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   6:   Examples  of  mechanophores   incorporated   into  polymer   chains.  A)  Poly(vinylpyrrolidone)  

with  random  peroxide  linkages.22  B)  Chain  scission  of  silver-­‐carbene-­‐based  polymer  upon  sonication  

and  subsequent   trapping  of   the  carbene  with  water.24  C)  Ultrasound-­‐Induced  cleavage  of  dicyano-­‐

substituted   cyclobutane   ring   yields   cyanoacrylate-­‐terminated   polymers.25   D)   Mechanochemical  

scission  of  heteronuclear  supramolecular  polymers  in  the  presence  of  a  scavenger  complex  resulted  

in  the  formation  of  Palladium  and  Platinum  heterocomplexes.26  

Coordination  bonds,  weaker  than  covalent  bonds  on  polymer  backbone  break  more  easiliy  

and   result   in   even   lower  Mlim   compared   to   their   covalent   counterparts.26   Recently   in   our  

group,  high-­‐molecular-­‐weight  linear  coordination  polymers  of  diphenylphosphine  telechelic  

polytetrahydrofuran  with   palladium(II)   dichloride  were   developed.27  Molecular  weights   of  

these   polymers   could   be   altered   reversibly   by   ultrasound   and   it   has   been   shown   that  

polytetrahydrofuran   chains   remain   intact   during   sonication.28   This   implies   that   only   the  

reversible   palladium–phosphorus   bonds   are   broken   and   coordinatively   unsaturated  

palladium   complexes   were   produced   by   the   application   of   mechanical   forces   on   these  

coordination   polymers.29   Furthermore,   polymers   which   include   both   PdII   and   PtII   were  

OO

OO

N N NO O O

m n x

NN pTHF*OMe

Ag⁺0X*

N N pTHF*OMe

ultrasound

NN pTHF*OMe

NN pTHF*OMe

H₂O

H

+

+0AgX

X:0Cl*0or0PF₆*

X:0AgCl₂0or0PF₆Mn0(pTHF):06.70kDA

BrO

O

O

OO

Br

O O

O

OOOON N

n n

ultrasound

BrO

O

OO

nO

ON

CH₃CN,06*90KC

Pd Pt pTHFpTHFpTHF

Cl

Cl

Cl

Cly

n

C₁₂H₂₅ Pd C₁₂H₂₅

Cl

Cl

ultrasoundPt0Scission

Pd0Scission

Pt C₁₂H₂₅

Cl

Cl

C₁₂H₂₅ Pt

Cl

Cl

pTHF pTHF

Pd C₁₂H₂₅

Cl

Cl

C₁₂H₂₅ Pd

Cl

Cl

pTHF pTHF

A)

B)

C)

D)

x

sonicated  and   shown   that   force   selectively  breaks   the  weaker  Pd-­‐Phosphine  bonds  which  

were  randomly  distributed  along  the  polymer  backbone.26    

Following   the   work   on   metal–phosphine   coordination   polymers,   our   group   started  

investigating  mechanical  dissociation  of  silver(I)-­‐coordination  complexes  with  N-­‐heterocyclic  

carbene   (NHC)   functionalized   polymers.24   Groote   et   al.   investigated   mechanochemical  

scission   of   metal-­‐ligand   bonds   in   Ag(NHC)2   supramolecular   polymer   complexes   by  

ultrasound   using   viscosity   measurements   and   molecular   dynamics   simulations   (MD)  

combined   with   constrained   geometry   optimization   calculations   (COGEF).4   Calculations  

indicated  that  the  force  required  to  break  metal−ligand  bond  is  between  400  and  500  pN,  

much  lower  than  the  force  that  is  typically  required  to  break  covalent  bonds  (several  nN).  It  

has   been   shown   that   polymers   with   a   Ag(NHC)2   coordination   complex   in   the   pTHF  main  

chain  have  significantly  lower  Mlim  values.16  

Mechanoresponsive  materials  

Mechanical  activation  of  chemical  bonds  on  polymers  promises  to  provide  opportunities  to  

detect   and   repair   damage   in   polymeric   materials.   Incorporation   of   mechanophores   into  

polymers  results   in  materials  sensitive  to  mechanical  stimuli  and  leads  to  useful  molecular  

transformations   under   stress.   In   solution,   mechanically   induced   chain   scission   has   been  

used  to  release  reactive  end  groups  such  as  cyanoacrylates25  and  trifluorovinyl  ethers30  from  

mechanophore  precursors.  Furthermore,   transition  metal   complexes   located   in   the  center  

of  polymer  chains  have  been  dissociated  by  breaking  the  coordination  bonds  mechanically,  

as  an  alternative  to  thermal  activation  for  latent  catalysts.1,31,32      

An   inspiring  example  of  mechanically   induced  transformations  on  polymeric  materials  was  

reported  recently  for  the  mechanochromic  spiropyran  (SP)  to  merocyanine  (MC)  transition.  

Initially  by  ultrasound  and  later  by  the  tensile  experiments  on  elastomeric  materials,  it  was  

shown   that   the   colorless   SP   is   ring-­‐opened   to   form   the   red   MC   dye   upon   mechanical  

activation.8,33   Therefore,   linear   or   crosslinked   SP-­‐functionalized   PMA   matrix   changes   its  

color   before   failure   upon   stretching.   In   later   studies,   SP-­‐based   mechanochromic   probes  

were   embedded   in   polyurethanes   and   polycaprolactone   to   monitor   the   effect   of  

temperature  and  plasticizers.25,34,35      

Chapter 1

8

Page 16: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   6:   Examples  of  mechanophores   incorporated   into  polymer   chains.  A)  Poly(vinylpyrrolidone)  

with  random  peroxide  linkages.22  B)  Chain  scission  of  silver-­‐carbene-­‐based  polymer  upon  sonication  

and  subsequent   trapping  of   the  carbene  with  water.24  C)  Ultrasound-­‐Induced  cleavage  of  dicyano-­‐

substituted   cyclobutane   ring   yields   cyanoacrylate-­‐terminated   polymers.25   D)   Mechanochemical  

scission  of  heteronuclear  supramolecular  polymers  in  the  presence  of  a  scavenger  complex  resulted  

in  the  formation  of  Palladium  and  Platinum  heterocomplexes.26  

Coordination  bonds,  weaker  than  covalent  bonds  on  polymer  backbone  break  more  easiliy  

and   result   in   even   lower  Mlim   compared   to   their   covalent   counterparts.26   Recently   in   our  

group,  high-­‐molecular-­‐weight  linear  coordination  polymers  of  diphenylphosphine  telechelic  

polytetrahydrofuran  with   palladium(II)   dichloride  were   developed.27  Molecular  weights   of  

these   polymers   could   be   altered   reversibly   by   ultrasound   and   it   has   been   shown   that  

polytetrahydrofuran   chains   remain   intact   during   sonication.28   This   implies   that   only   the  

reversible   palladium–phosphorus   bonds   are   broken   and   coordinatively   unsaturated  

palladium   complexes   were   produced   by   the   application   of   mechanical   forces   on   these  

coordination   polymers.29   Furthermore,   polymers   which   include   both   PdII   and   PtII   were  

OO

OO

N N NO O O

m n x

NN pTHF*OMe

Ag⁺0X*

N N pTHF*OMe

ultrasound

NN pTHF*OMe

NN pTHF*OMe

H₂O

H

+

+0AgX

X:0Cl*0or0PF₆*

X:0AgCl₂0or0PF₆Mn0(pTHF):06.70kDA

BrO

O

O

OO

Br

O O

O

OOOON N

n n

ultrasound

BrO

O

OO

nO

ON

CH₃CN,06*90KC

Pd Pt pTHFpTHFpTHF

Cl

Cl

Cl

Cly

n

C₁₂H₂₅ Pd C₁₂H₂₅

Cl

Cl

ultrasoundPt0Scission

Pd0Scission

Pt C₁₂H₂₅

Cl

Cl

C₁₂H₂₅ Pt

Cl

Cl

pTHF pTHF

Pd C₁₂H₂₅

Cl

Cl

C₁₂H₂₅ Pd

Cl

Cl

pTHF pTHF

A)

B)

C)

D)

x

sonicated  and   shown   that   force   selectively  breaks   the  weaker  Pd-­‐Phosphine  bonds  which  

were  randomly  distributed  along  the  polymer  backbone.26    

Following   the   work   on   metal–phosphine   coordination   polymers,   our   group   started  

investigating  mechanical  dissociation  of  silver(I)-­‐coordination  complexes  with  N-­‐heterocyclic  

carbene   (NHC)   functionalized   polymers.24   Groote   et   al.   investigated   mechanochemical  

scission   of   metal-­‐ligand   bonds   in   Ag(NHC)2   supramolecular   polymer   complexes   by  

ultrasound   using   viscosity   measurements   and   molecular   dynamics   simulations   (MD)  

combined   with   constrained   geometry   optimization   calculations   (COGEF).4   Calculations  

indicated  that  the  force  required  to  break  metal−ligand  bond  is  between  400  and  500  pN,  

much  lower  than  the  force  that  is  typically  required  to  break  covalent  bonds  (several  nN).  It  

has   been   shown   that   polymers   with   a   Ag(NHC)2   coordination   complex   in   the   pTHF  main  

chain  have  significantly  lower  Mlim  values.16  

Mechanoresponsive  materials  

Mechanical  activation  of  chemical  bonds  on  polymers  promises  to  provide  opportunities  to  

detect   and   repair   damage   in   polymeric   materials.   Incorporation   of   mechanophores   into  

polymers  results   in  materials  sensitive  to  mechanical  stimuli  and  leads  to  useful  molecular  

transformations   under   stress.   In   solution,   mechanically   induced   chain   scission   has   been  

used  to  release  reactive  end  groups  such  as  cyanoacrylates25  and  trifluorovinyl  ethers30  from  

mechanophore  precursors.  Furthermore,   transition  metal   complexes   located   in   the  center  

of  polymer  chains  have  been  dissociated  by  breaking  the  coordination  bonds  mechanically,  

as  an  alternative  to  thermal  activation  for  latent  catalysts.1,31,32      

An   inspiring  example  of  mechanically   induced  transformations  on  polymeric  materials  was  

reported  recently  for  the  mechanochromic  spiropyran  (SP)  to  merocyanine  (MC)  transition.  

Initially  by  ultrasound  and  later  by  the  tensile  experiments  on  elastomeric  materials,  it  was  

shown   that   the   colorless   SP   is   ring-­‐opened   to   form   the   red   MC   dye   upon   mechanical  

activation.8,33   Therefore,   linear   or   crosslinked   SP-­‐functionalized   PMA   matrix   changes   its  

color   before   failure   upon   stretching.   In   later   studies,   SP-­‐based   mechanochromic   probes  

were   embedded   in   polyurethanes   and   polycaprolactone   to   monitor   the   effect   of  

temperature  and  plasticizers.25,34,35      

Polymer Mechanochemistry

9

1

Page 17: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   7:   Examples   of   mechanoresponsive   materials   in   recent   literature:   A)   Mechanochromism:  

force-­‐induced   ring-­‐opening   of   spyropyran   to   merocyanine;33   B)   Mechanoluminescence:   polymers  

containing  bis(adamantyl)-­‐1,2-­‐dioxetane  mechanophores  emit  light  upon  stretching.36  

Another   elegant   example   as   a   potential   stress   reporter,   a  mechano-­‐luminescent  material  

based   on   a   polymer-­‐functionalized   1,2-­‐dioxetane  moiety,   has   been   developed   recently   in  

our  group.36  Dioxetanes  are  organic  peroxides  and  efficient  sources  of  electronically  excited  

products   upon   chemical   or   thermal   treatment.37   Opening   of   a   four-­‐membered   1,2-­‐

dioxetane  ring  with  two  adjacent  oxygen  atoms  yields  two  carbonyl  groups,  one  of  which  is  

in   an   electronically   excited   state   and   emits   a   photon   (at   420   nm)   on   relaxation.  

Bis(adamantyl)-­‐substituted   1,2-­‐dioxetane   is   thermally   stable   at   room   temperature   (up   to  

~200   C)   with   a   scission   barrier   of   37   kcal/mol,   36,38   and   was   successfully   activated  

mechanically   by   covalent   incorporation   in   poly(methylacrylate)   chains.   Experiments   have  

shown   that  dioxetanes  can  be  used  as  mechanical-­‐probe   for   spatiotemporal  mapping  and  

chain  scission  in  polymers.    

Mechanical  release  of  small  molecules  in  polymer  matrices  

The   proliferation   of   mechanophores   capable   of   producing   useful   reactive   intermediates  

within   the   polymer   main   chain   has   prompted   others   to   investigate   different   modes   of  

chemically   productive  mechanical   reactivity.   One   such  mode   is   the  mechanically   induced  

release   of   small   molecules,   a   recent   addition   to   the   mechanochemistry   field.  

Mechanophores  are   incorporated   into  elastomeric  networks  and  an  applied  force   leads  to  

conformational  changes  and  subsequent  scission  of  bonds  that  are  not  part  of  the  polymer  

backbone,  whilst  maintaining  the  overall  mechanical   integrity  of  the  polymer  matrix.39  The  

range   of   accessible   reactivities   remains   limited,   but   researchers   in   this   fledgling   area   are  

starting   to   make   use   of   the   small   molecule   products   for   further   reactions,   including  

polymerisation.   This   type   of   activation   displays   potential   for   applications   in  

mechanochemical  catalysis,  mapping  deformations  and  damages   in  polymer  networks  and  

self-­‐healing  or  self-­‐reinforcing  elastomers.40  

The  first  example  of  this  type  of  activation  was  the  mechanochemical  generation  of  an  acid,  

developed  by  Diesendruck  et  al.   Inspired  by  Craig’s  gDHC  system,   they   incorporated  gem-­‐  

dichlorotetrahydro   cyclopropanated   indene   into   polymethyl   acrylate   matrix   (PMA)   and  

showed  that  compression  resulted   in  ring  opening  of  cyclopropane  to  give  the  elimination  

product  2-­‐  chloronaphthalene,  with  the  release  of  HCl.41  Calorimetric  analysis  of  the  sample  

before  and  after  compression  demonstrated  that  up  to  20%  mechanophore  conversion  was  

achieved   at   a   load   of   352   MPa.   Control   polymer   in   which   the   mechanophore   was   not  

covalently   incorporated   into   the   PMA   matrix   showed   6%   conversion   under   the   same  

conditions.  Although  significantly  high  thermal  background  reaction  observed  with  control  

polymer   might   limit   its   practical   applications,   this   work   represented   an   important   step  

towards  the  realization  of  autonomous  self-­‐healing  materials.      

Reports  have  also  emerged  from  the  Boydston  group  describing  the  flex,  or  bond-­‐bending,  

activation   of   an   oxanorbornadiene,   a   Diels–Alder   adduct   of   furan   and   dimethyl  

acetylenedicarboxylate.42  They   incorporated   the  mechanophore   into  poly(methyl  acrylate)  

(PMA)   matrix   and   showed   that   the   furan   derivative   could   be   released   via   a   retro-­‐[4+2]  

cycloaddition   under   stress   applied   to   the   bulk   polymer.   The   mechanically   initiated  

cycloreversion  converts  main-­‐chain  alkene  moieties  into  alkynes  and  results  in  linearization  

Chapter 1

10

Page 18: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   7:   Examples   of   mechanoresponsive   materials   in   recent   literature:   A)   Mechanochromism:  

force-­‐induced   ring-­‐opening   of   spyropyran   to   merocyanine;33   B)   Mechanoluminescence:   polymers  

containing  bis(adamantyl)-­‐1,2-­‐dioxetane  mechanophores  emit  light  upon  stretching.36  

Another   elegant   example   as   a   potential   stress   reporter,   a  mechano-­‐luminescent  material  

based   on   a   polymer-­‐functionalized   1,2-­‐dioxetane  moiety,   has   been   developed   recently   in  

our  group.36  Dioxetanes  are  organic  peroxides  and  efficient  sources  of  electronically  excited  

products   upon   chemical   or   thermal   treatment.37   Opening   of   a   four-­‐membered   1,2-­‐

dioxetane  ring  with  two  adjacent  oxygen  atoms  yields  two  carbonyl  groups,  one  of  which  is  

in   an   electronically   excited   state   and   emits   a   photon   (at   420   nm)   on   relaxation.  

Bis(adamantyl)-­‐substituted   1,2-­‐dioxetane   is   thermally   stable   at   room   temperature   (up   to  

~200   C)   with   a   scission   barrier   of   37   kcal/mol,   36,38   and   was   successfully   activated  

mechanically   by   covalent   incorporation   in   poly(methylacrylate)   chains.   Experiments   have  

shown   that  dioxetanes  can  be  used  as  mechanical-­‐probe   for   spatiotemporal  mapping  and  

chain  scission  in  polymers.    

Mechanical  release  of  small  molecules  in  polymer  matrices  

The   proliferation   of   mechanophores   capable   of   producing   useful   reactive   intermediates  

within   the   polymer   main   chain   has   prompted   others   to   investigate   different   modes   of  

chemically   productive  mechanical   reactivity.   One   such  mode   is   the  mechanically   induced  

release   of   small   molecules,   a   recent   addition   to   the   mechanochemistry   field.  

Mechanophores  are   incorporated   into  elastomeric  networks  and  an  applied  force   leads  to  

conformational  changes  and  subsequent  scission  of  bonds  that  are  not  part  of  the  polymer  

backbone,  whilst  maintaining  the  overall  mechanical   integrity  of  the  polymer  matrix.39  The  

range   of   accessible   reactivities   remains   limited,   but   researchers   in   this   fledgling   area   are  

starting   to   make   use   of   the   small   molecule   products   for   further   reactions,   including  

polymerisation.   This   type   of   activation   displays   potential   for   applications   in  

mechanochemical  catalysis,  mapping  deformations  and  damages   in  polymer  networks  and  

self-­‐healing  or  self-­‐reinforcing  elastomers.40  

The  first  example  of  this  type  of  activation  was  the  mechanochemical  generation  of  an  acid,  

developed  by  Diesendruck  et  al.   Inspired  by  Craig’s  gDHC  system,   they   incorporated  gem-­‐  

dichlorotetrahydro   cyclopropanated   indene   into   polymethyl   acrylate   matrix   (PMA)   and  

showed  that  compression  resulted   in  ring  opening  of  cyclopropane  to  give  the  elimination  

product  2-­‐  chloronaphthalene,  with  the  release  of  HCl.41  Calorimetric  analysis  of  the  sample  

before  and  after  compression  demonstrated  that  up  to  20%  mechanophore  conversion  was  

achieved   at   a   load   of   352   MPa.   Control   polymer   in   which   the   mechanophore   was   not  

covalently   incorporated   into   the   PMA   matrix   showed   6%   conversion   under   the   same  

conditions.  Although  significantly  high  thermal  background  reaction  observed  with  control  

polymer   might   limit   its   practical   applications,   this   work   represented   an   important   step  

towards  the  realization  of  autonomous  self-­‐healing  materials.      

Reports  have  also  emerged  from  the  Boydston  group  describing  the  flex,  or  bond-­‐bending,  

activation   of   an   oxanorbornadiene,   a   Diels–Alder   adduct   of   furan   and   dimethyl  

acetylenedicarboxylate.42  They   incorporated   the  mechanophore   into  poly(methyl  acrylate)  

(PMA)   matrix   and   showed   that   the   furan   derivative   could   be   released   via   a   retro-­‐[4+2]  

cycloaddition   under   stress   applied   to   the   bulk   polymer.   The   mechanically   initiated  

cycloreversion  converts  main-­‐chain  alkene  moieties  into  alkynes  and  results  in  linearization  

Polymer Mechanochemistry

11

1

Page 19: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

that   occurs   during   polymer   elongation.   After   compression   the   polymer   was   soaked   in  

dichloromethane  and  the  small  molecules  released  mechanically  diffused  out  of  the  matrix;  

they  could  then  be  identified  by  and  monitored  with  GC-­‐MS  and  NMR.  However,  the  forces  

required   to   activate   the   mechanophore   caused   failure   in   the   PMA   matrix,   limiting   the  

number  of  loading  cycles.  In  a  second  publication,  the  authors  partly  addressed  this  issue  by  

incorporating   the   same   mechanophore   in   a   segmented   PU   matrix   which   required   lower  

stress   loadings   for   successive   mechanophore   activation.43   Nevertheless,   as   a   result   of  

random  scission  of  the  chemical  crosslinks  and  the  destruction  of  physical  crosslinks  within  

the   hard   domains   of   the   segmented   PU   matrix,   only   a   maximum   of   7%   mechanophore  

activation  could  be  reached  after  15  compression  cycles.    

 

Figure   8:   Mechanical   release   of   small   molecules   in   polymer   matrices:   A)   Potential   indole-­‐based  

mechanocatalyst   for   acid   generation.41   B)   Solid   state   mechanoactivation   of   oxanorbornadiene,  

producing  an  alkyne  in  the  polymer  backbone  and  releasing  a  small  molecule  furan.42  

 

 

A)

B)

Mechanochemical  catalysis  

Mechanocatalysts   are   catalysts   of   which   the   activity   or   specificity   is   modified   under   the  

influence   of  mechanical   force.   Two   distinct  modes   of   activation   can   be   envisioned:   steric  

modification  or  unblocking  of  active  sites.   In  steric  activation,  the  activity  of  the  catalyst   is  

modified   by   a   change   in   the   steric   environment   of   the   active   site,   e.g.   by   changing   the  

relative  position  of  catalytically  active  groups   that   form  the  active  site  or  by  changing   the  

configuration   of   the   binding   site.   Recently,   research   investigating   the   effect   of   a  

photochemical   switch   coupled   to   a   chelating   bisphosphine   on   the   activity   profile   of   the  

catalyst  was  published.  The  photochemical  switch,  a  biindane,  changes  the  bite  angle  of  the  

chelating   ligand,   and   influences   the   enantioselectivity   of   reactions   catalysed   by   Pd  

complexes   of   the   photoresponsive   ligand.   It   was   found   that   the   effect   of   switching   was  

largest  for  Heck  arylation  reactions.44  

The  second  and  the  more  common  approach  to  mechanocatalysis  in  synthetic  systems  is  to  

activate  a  catalyst  by  modifying  its  electronic  properties.  Catalysts  in  a  latent  state  because  

of   pairing   of   acidic   and   basic   sites   are   well   known   for   their   capability   to   be   activated  

thermally,  and  are  employed  in  a  number  of  different  reactions.  Some  of  the  most  striking  

examples  of  this  kind  of  catalysts  are  N-­‐heterocyclic  carbenes  (NHCs).45  These  Lewis  bases  

have  been  used  as  catalysts  in  various  organic  transformations,  including  condensation,  1,2-­‐  

and   1,4-­‐addition,   transesterification   and   ring-­‐opening   reactions.46,47   Due   to   their  

nucleophilicity   they   show   high   reactivity   towards   various   substrates   but   they   can   also   be  

masked   in   thermally   labile   precursors.48   NHC−metal   complexes   have   been   applied   as  

thermally   latent   catalysts   in   the   preparation   of   a   number   of   polymers,   such   as  

poly(urethane),  poly(methyl  methacrylate),  poly(caprolactone),  and  poly(amide).49–53  

By  their  nature,  latent  transition  metal  (TM)  catalysts54,55  with  strongly  bound  ligands  can  be  

adapted   to   form   mechanocatalysts   by   providing   the   ligands   with   ‘handles’   to   transfer  

mechanical   force.   These   handles   can   be   (linear)   polymer   chains,   which   provide   drag   in   a  

viscous   system   that   is   sheared   or   undergoes   elongational   strain,   or   the   ligands   can   be  

connected  to  a  polymer  network  in  an  elastic  system.          

Much  of  the  work  done  in  recent  years  on  polymer  mechanochemistry  has  made  use  of  the  

high   elongational   strain   rates   observed   around   collapsing   cavitation   bubbles   in   sonicated  

Chapter 1

12

Page 20: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

that   occurs   during   polymer   elongation.   After   compression   the   polymer   was   soaked   in  

dichloromethane  and  the  small  molecules  released  mechanically  diffused  out  of  the  matrix;  

they  could  then  be  identified  by  and  monitored  with  GC-­‐MS  and  NMR.  However,  the  forces  

required   to   activate   the   mechanophore   caused   failure   in   the   PMA   matrix,   limiting   the  

number  of  loading  cycles.  In  a  second  publication,  the  authors  partly  addressed  this  issue  by  

incorporating   the   same   mechanophore   in   a   segmented   PU   matrix   which   required   lower  

stress   loadings   for   successive   mechanophore   activation.43   Nevertheless,   as   a   result   of  

random  scission  of  the  chemical  crosslinks  and  the  destruction  of  physical  crosslinks  within  

the   hard   domains   of   the   segmented   PU   matrix,   only   a   maximum   of   7%   mechanophore  

activation  could  be  reached  after  15  compression  cycles.    

 

Figure   8:   Mechanical   release   of   small   molecules   in   polymer   matrices:   A)   Potential   indole-­‐based  

mechanocatalyst   for   acid   generation.41   B)   Solid   state   mechanoactivation   of   oxanorbornadiene,  

producing  an  alkyne  in  the  polymer  backbone  and  releasing  a  small  molecule  furan.42  

 

 

A)

B)

Mechanochemical  catalysis  

Mechanocatalysts   are   catalysts   of   which   the   activity   or   specificity   is   modified   under   the  

influence   of  mechanical   force.   Two   distinct  modes   of   activation   can   be   envisioned:   steric  

modification  or  unblocking  of  active  sites.   In  steric  activation,  the  activity  of  the  catalyst   is  

modified   by   a   change   in   the   steric   environment   of   the   active   site,   e.g.   by   changing   the  

relative  position  of  catalytically  active  groups   that   form  the  active  site  or  by  changing   the  

configuration   of   the   binding   site.   Recently,   research   investigating   the   effect   of   a  

photochemical   switch   coupled   to   a   chelating   bisphosphine   on   the   activity   profile   of   the  

catalyst  was  published.  The  photochemical  switch,  a  biindane,  changes  the  bite  angle  of  the  

chelating   ligand,   and   influences   the   enantioselectivity   of   reactions   catalysed   by   Pd  

complexes   of   the   photoresponsive   ligand.   It   was   found   that   the   effect   of   switching   was  

largest  for  Heck  arylation  reactions.44  

The  second  and  the  more  common  approach  to  mechanocatalysis  in  synthetic  systems  is  to  

activate  a  catalyst  by  modifying  its  electronic  properties.  Catalysts  in  a  latent  state  because  

of   pairing   of   acidic   and   basic   sites   are   well   known   for   their   capability   to   be   activated  

thermally,  and  are  employed  in  a  number  of  different  reactions.  Some  of  the  most  striking  

examples  of  this  kind  of  catalysts  are  N-­‐heterocyclic  carbenes  (NHCs).45  These  Lewis  bases  

have  been  used  as  catalysts  in  various  organic  transformations,  including  condensation,  1,2-­‐  

and   1,4-­‐addition,   transesterification   and   ring-­‐opening   reactions.46,47   Due   to   their  

nucleophilicity   they   show   high   reactivity   towards   various   substrates   but   they   can   also   be  

masked   in   thermally   labile   precursors.48   NHC−metal   complexes   have   been   applied   as  

thermally   latent   catalysts   in   the   preparation   of   a   number   of   polymers,   such   as  

poly(urethane),  poly(methyl  methacrylate),  poly(caprolactone),  and  poly(amide).49–53  

By  their  nature,  latent  transition  metal  (TM)  catalysts54,55  with  strongly  bound  ligands  can  be  

adapted   to   form   mechanocatalysts   by   providing   the   ligands   with   ‘handles’   to   transfer  

mechanical   force.   These   handles   can   be   (linear)   polymer   chains,   which   provide   drag   in   a  

viscous   system   that   is   sheared   or   undergoes   elongational   strain,   or   the   ligands   can   be  

connected  to  a  polymer  network  in  an  elastic  system.          

Much  of  the  work  done  in  recent  years  on  polymer  mechanochemistry  has  made  use  of  the  

high   elongational   strain   rates   observed   around   collapsing   cavitation   bubbles   in   sonicated  

Polymer Mechanochemistry

13

1

Page 21: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

solutions.13   In   addition   to   the   distinctive   features   of   sonochemically   induced   mechanical  

reactivity  described  above,  further  attention  needs  to  be  paid  to  the  sonication  conditions  in  

the  case  of  mechanochemical  catalysis,  because  catalyst  lifetime  and  turnover  number  may  

be  reduced  by  sonochemical  byproducts.  Therefore,  the  mechanochemical  catalysis  should  

be  performed  under  a  gas  that  increases  the  lifetime  of  the  active  catalyst  while  still  leading  

to  strong  cavitation  as  mentioned  above.  

The  concept  of  mechanochemical  activation  of  a  latent  catalyst  by  ultrasound  is  illustrated  

by   the   mechanochemical   scission   of   metal-­‐ligand   bonds   in   Ag(NHC)2   supramolecular  

polymer  complexes.1,56  External  force  selectively  breaks  Ag-­‐NHC  bonds  and  yields  free  NHC  

which  was  used  to  catalyze  the  transesterification  of  benzyl  alcohol  and  vinyl  acetate  under  

sonication.56,57  The  complex  form  of  the  carbene  displayed  no  activity  proving  the  latency  of  

the  catalyst.  Control  experiments  confirmed  that  the  catalyst  was  activated  mechanically.    

After   successful   application   of   the   concept   of  mechanocatalysis   its   generality   was   tested  

with   bis-­‐NHC   ruthenium–alkylidene   complex.31   Mechanistic   studies   revealed   that   ligand  

dissociation   is   a   crucial   step   in   catalyst   activation   for   Ru   mediated   olefin   metathesis  

reactions  to  form  coordinatively  unsaturated  reactive  Ru  species.58  Among  several  effective  

Ru  catalysts  bis-­‐NHC  ruthenium–alkylidene  complexes  were  shown  to  be  latent  at  ambient  

temperature   since   dissociation   of   strong   Ru-­‐NHC   bond   requires   elevated   temperatures.59  

Piermattei  showed  that  Ru  catalysts  with  pTHF  chains  attached  bis  NHC  ligands  resulted  in  a  

latent  metathesis  catalyst  that  can  be  activated  by  mechanical  force.56  Sonicating  a  solution  

of  diethyl  diallyl  malonate  (DEDAM)  in  the  presence  of  mechanically  responsive  Ru  catalysts  

(36  kg  mol-­‐1)  resulted  in  approximately  20%  conversion  after  1h.  Control  experiments  were  

conducted  to  prove  that  catalyst  activation  is  mechanical  rather  than  thermal  in  nature.    A  

lower  MW  analogue  of  the  catalyst  (18  kg  mol-­‐1)  showed  lower  activity  due  to  slower  chain  

scission   rate   that   decreased   the   amount   of   active   catalyst   formed   in   the   timespan   of  

sonication.   Replacing   polymer   actuators   by   butyl   chains   attached   to   NHC   resulted   in   a  

mechanically   silent   latent   catalyst,   showed   less   than   0.2  %   conversion   in   the   presence   of  

DEDAM  after  1h  of  sonication.    

 

Figure  9:  Mechanochemical  catalysis:  A)  Photo-­‐mechanoactivation  of  a  palladium  catalyst   for  Heck  

arylation   B)   Mechanically   activated   catalysts   and   the   corresponding   catalytic   reactions   C)  

Mechanoactivation  of  latent  ruthenium  catalyst  in  the  solid  state,  initiating  in  situ  polymerization  of  

norbornene  monomer  in  response  to  stress    

In   a   later   study,   bis-­‐NHC   ruthenium–alkylidene   complex  was   activated  under   compressive  

strain.60   In   order   to   initiate   Ru   mediated   polymerization   of   norbornene   in   solid   state,  

Chapter 1

14

Page 22: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

solutions.13   In   addition   to   the   distinctive   features   of   sonochemically   induced   mechanical  

reactivity  described  above,  further  attention  needs  to  be  paid  to  the  sonication  conditions  in  

the  case  of  mechanochemical  catalysis,  because  catalyst  lifetime  and  turnover  number  may  

be  reduced  by  sonochemical  byproducts.  Therefore,  the  mechanochemical  catalysis  should  

be  performed  under  a  gas  that  increases  the  lifetime  of  the  active  catalyst  while  still  leading  

to  strong  cavitation  as  mentioned  above.  

The  concept  of  mechanochemical  activation  of  a  latent  catalyst  by  ultrasound  is  illustrated  

by   the   mechanochemical   scission   of   metal-­‐ligand   bonds   in   Ag(NHC)2   supramolecular  

polymer  complexes.1,56  External  force  selectively  breaks  Ag-­‐NHC  bonds  and  yields  free  NHC  

which  was  used  to  catalyze  the  transesterification  of  benzyl  alcohol  and  vinyl  acetate  under  

sonication.56,57  The  complex  form  of  the  carbene  displayed  no  activity  proving  the  latency  of  

the  catalyst.  Control  experiments  confirmed  that  the  catalyst  was  activated  mechanically.    

After   successful   application   of   the   concept   of  mechanocatalysis   its   generality   was   tested  

with   bis-­‐NHC   ruthenium–alkylidene   complex.31   Mechanistic   studies   revealed   that   ligand  

dissociation   is   a   crucial   step   in   catalyst   activation   for   Ru   mediated   olefin   metathesis  

reactions  to  form  coordinatively  unsaturated  reactive  Ru  species.58  Among  several  effective  

Ru  catalysts  bis-­‐NHC  ruthenium–alkylidene  complexes  were  shown  to  be  latent  at  ambient  

temperature   since   dissociation   of   strong   Ru-­‐NHC   bond   requires   elevated   temperatures.59  

Piermattei  showed  that  Ru  catalysts  with  pTHF  chains  attached  bis  NHC  ligands  resulted  in  a  

latent  metathesis  catalyst  that  can  be  activated  by  mechanical  force.56  Sonicating  a  solution  

of  diethyl  diallyl  malonate  (DEDAM)  in  the  presence  of  mechanically  responsive  Ru  catalysts  

(36  kg  mol-­‐1)  resulted  in  approximately  20%  conversion  after  1h.  Control  experiments  were  

conducted  to  prove  that  catalyst  activation  is  mechanical  rather  than  thermal  in  nature.    A  

lower  MW  analogue  of  the  catalyst  (18  kg  mol-­‐1)  showed  lower  activity  due  to  slower  chain  

scission   rate   that   decreased   the   amount   of   active   catalyst   formed   in   the   timespan   of  

sonication.   Replacing   polymer   actuators   by   butyl   chains   attached   to   NHC   resulted   in   a  

mechanically   silent   latent   catalyst,   showed   less   than   0.2  %   conversion   in   the   presence   of  

DEDAM  after  1h  of  sonication.    

 

Figure  9:  Mechanochemical  catalysis:  A)  Photo-­‐mechanoactivation  of  a  palladium  catalyst   for  Heck  

arylation   B)   Mechanically   activated   catalysts   and   the   corresponding   catalytic   reactions   C)  

Mechanoactivation  of  latent  ruthenium  catalyst  in  the  solid  state,  initiating  in  situ  polymerization  of  

norbornene  monomer  in  response  to  stress    

In   a   later   study,   bis-­‐NHC   ruthenium–alkylidene   complex  was   activated  under   compressive  

strain.60   In   order   to   initiate   Ru   mediated   polymerization   of   norbornene   in   solid   state,  

Polymer Mechanochemistry

15

1

Page 23: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

polymer   catalyst   (34   kg   mol-­‐1)   and   a   norbornene   monomer   were   incorporated   in   a   high  

molecular   weight   Poly(tetrahydrofuran)   (pTHF)   matrix   (Mn   =   170   kDa,   PDI   =   1.3)   which  

provided   the   physical   cross-­‐linking   through   the   crystalline   domains   and   allowed  

macroscopic  forces  to  be  transferred  to  the  metal–ligand  bonds.  Consecutive  compressions  

showed  that  up  to  25%  of  norbornene  monomer  was  polymerized  after  five  loading  cycles.  

Aim  and  outline  of  this  thesis  

The  main  aim  of  this  thesis  is  to  gain  a  better  understanding  of  the  fundamental  processes  

and  mechanisms  underlying  mechanochemical  chain  scission   in  organometallic  complexes.  

Throughout  this  work,  we  have  used  Palladium  (II)  complexes  with  N-­‐heterocyclic  carbene  

(NHC)   or   imidazole   ligands.   These   coordination   complexes   are   embedded  within   polymer  

chains  and  coordination  bonds  are  broken  by  ultrasonication  or  by  straining  bulk  polymer  

samples.    

In  Chapter  2,  ultrasound  induced  chain  scission  in  coordination  complexes  of  Palladium  and  

Platinum   with   polytetrahydrofuran   functionalized   N-­‐heterocyclic   carbene   ligands   is  

reported.  Reversibility  of  chain  scission  and  molecular  weight  dependence  of   scission   rate  

were   determined.   Comparing   scission   of   Palladium   and   Platinum   containing   polymer  

showed   the   influence   of   ligand   dissociation   energy   on  mechanochemical   response   of   the  

coordination  polymers.    

Scission  of   the  Pd   -­‐  NHC  bond   releases   free  NHC.   In  Chapter   3,   this  basic  NHC,   is  used   to  

abstract   a   proton   from   a   2-­‐coumaranone   derivative,   which   decomposes   via   a  

chemiluminescent   pathway   once   deprotonated   in   the   presence   of   oxygen.   Rate   of  

ultrasound   induced   scission   and   molecular   weight   threshold   (Mlim)   for   mechanochemical  

chain   scission   were   determined   under   the   reaction   conditions   that   coumaranone  

decomposition  could  be  visually  monitored.    

In  Chapter  4,  thermal  ligand  exchange  between  imidazole  Palladium  (II)  complexes  has  been  

investigated  to  have  a  better  understanding  on  ligand  exchange  dynamics  and  to  determine  

the  rate  of  sonication  induced  ligand  exchange  in  Imidazole-­‐Pd  coordination  polymers.    

In   Chapter   5,   information   on   the   mechanism   of   mechanochemical   chain   scission   and  

mechanochemically   induced   ligand   exchange   of   Pd(II)   complexes,   obtained   in   previous  

chapters  was  used  to  direct  the  formation  of  heterocomplexes.  Symmetric  complexes  with  

high   and   low   molecular   weight   polymer-­‐attached   ligands   were   mixed   in   solution   and  

sonicated.  When  one  of   the   complexes  has  a  molecular  weight  higher   than   the   threshold  

(Mlim)  for  mechanochemical  chain  scission,  while  the  other  is  smaller,  sonication  leads  to  the  

directed  formation  of  a  heterocomplex  with  two  different  ligands.      

In  Chapter   6,   initial   attempts   and  preliminary   results   for   self-­‐healing  property  of   polymer  

films  of  poly(methyl  acrylate)-­‐vinyl  imidazole  copolymers,  which  were  crosslinked  by  ligand  

metal   coordination   were   reported.   Reversibility   of   mechanochemical   chain   scission   in  

coordination  complexes  was  utilized  in  the  solid  state  to  regain  the  mechanical  properties  of  

a  cross-­‐linked  polymer  film.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Chapter 1

16

Page 24: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

polymer   catalyst   (34   kg   mol-­‐1)   and   a   norbornene   monomer   were   incorporated   in   a   high  

molecular   weight   Poly(tetrahydrofuran)   (pTHF)   matrix   (Mn   =   170   kDa,   PDI   =   1.3)   which  

provided   the   physical   cross-­‐linking   through   the   crystalline   domains   and   allowed  

macroscopic  forces  to  be  transferred  to  the  metal–ligand  bonds.  Consecutive  compressions  

showed  that  up  to  25%  of  norbornene  monomer  was  polymerized  after  five  loading  cycles.  

Aim  and  outline  of  this  thesis  

The  main  aim  of  this  thesis  is  to  gain  a  better  understanding  of  the  fundamental  processes  

and  mechanisms  underlying  mechanochemical  chain  scission   in  organometallic  complexes.  

Throughout  this  work,  we  have  used  Palladium  (II)  complexes  with  N-­‐heterocyclic  carbene  

(NHC)   or   imidazole   ligands.   These   coordination   complexes   are   embedded  within   polymer  

chains  and  coordination  bonds  are  broken  by  ultrasonication  or  by  straining  bulk  polymer  

samples.    

In  Chapter  2,  ultrasound  induced  chain  scission  in  coordination  complexes  of  Palladium  and  

Platinum   with   polytetrahydrofuran   functionalized   N-­‐heterocyclic   carbene   ligands   is  

reported.  Reversibility  of  chain  scission  and  molecular  weight  dependence  of   scission   rate  

were   determined.   Comparing   scission   of   Palladium   and   Platinum   containing   polymer  

showed   the   influence   of   ligand   dissociation   energy   on  mechanochemical   response   of   the  

coordination  polymers.    

Scission  of   the  Pd   -­‐  NHC  bond   releases   free  NHC.   In  Chapter   3,   this  basic  NHC,   is  used   to  

abstract   a   proton   from   a   2-­‐coumaranone   derivative,   which   decomposes   via   a  

chemiluminescent   pathway   once   deprotonated   in   the   presence   of   oxygen.   Rate   of  

ultrasound   induced   scission   and   molecular   weight   threshold   (Mlim)   for   mechanochemical  

chain   scission   were   determined   under   the   reaction   conditions   that   coumaranone  

decomposition  could  be  visually  monitored.    

In  Chapter  4,  thermal  ligand  exchange  between  imidazole  Palladium  (II)  complexes  has  been  

investigated  to  have  a  better  understanding  on  ligand  exchange  dynamics  and  to  determine  

the  rate  of  sonication  induced  ligand  exchange  in  Imidazole-­‐Pd  coordination  polymers.    

In   Chapter   5,   information   on   the   mechanism   of   mechanochemical   chain   scission   and  

mechanochemically   induced   ligand   exchange   of   Pd(II)   complexes,   obtained   in   previous  

chapters  was  used  to  direct  the  formation  of  heterocomplexes.  Symmetric  complexes  with  

high   and   low   molecular   weight   polymer-­‐attached   ligands   were   mixed   in   solution   and  

sonicated.  When  one  of   the   complexes  has  a  molecular  weight  higher   than   the   threshold  

(Mlim)  for  mechanochemical  chain  scission,  while  the  other  is  smaller,  sonication  leads  to  the  

directed  formation  of  a  heterocomplex  with  two  different  ligands.      

In  Chapter   6,   initial   attempts   and  preliminary   results   for   self-­‐healing  property  of   polymer  

films  of  poly(methyl  acrylate)-­‐vinyl  imidazole  copolymers,  which  were  crosslinked  by  ligand  

metal   coordination   were   reported.   Reversibility   of   mechanochemical   chain   scission   in  

coordination  complexes  was  utilized  in  the  solid  state  to  regain  the  mechanical  properties  of  

a  cross-­‐linked  polymer  film.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Polymer Mechanochemistry

17

1

Page 25: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

References  

(1)     Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat  Chem  2010,  2  (6),  436–437.  

(2)     Kauzmann,  W.;  Eyring,  H.  J.  Am.  Chem.  Soc.  1940,  62  (11),  3113–3125.  

(3)     Beyer,  M.  K.  J.  Chem.  Phys.  2000,  112  (17),  7307–7312.  

(4)     Groote,  R.;  Szyja,  B.  M.;  Pidko,  E.  A.;  Hensen,  E.  J.  M.;  Sijbesma,  R.  P.  Macromolecules  2011,  

44  (23),  9187–9195.  

(5)     Staudinger,  H.;  Heuer,  W.  Berichte  Dtsch.  Chem.  Ges.  B  Ser.  1934,  67  (7),  1159–1164.  

(6)     Hickenboth,  C.  R.;  Moore,  J.  S.;  White,  S.  R.;  Sottos,  N.  R.;  Baudry,  J.;  Wilson,  S.  R.  Nature  

2007,  446  (7134),  423–427.  

(7)     Ong,  M.  T.;  Leiding,  J.;  Tao,  H.;  Virshup,  A.  M.;  Martínez,  T.  J.  J.  Am.  Chem.  Soc.  2009,  131  (18),  

6377–6379.  

(8)     Potisek,  S.  L.;  Davis,  D.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2007,  129  

(45),  13808–13809.  

(9)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  J.  Am.  Chem.  Soc.  2010,  132  (30),  10609–10614.  

(10)     Black,  A.  L.;  Orlicki,  J.  A.;  Craig,  S.  L.  J.  Mater.  Chem.  2011,  21  (23),  8460–8465.  

(11)     Zocchi,  G.  Annu.  Rev.  Biophys.  2009,  38  (1),  75–88.  

(12)     Klibanov,  A.  M.;  Samokhin,  G.  P.;  Martinek,  K.;  Berezin,  I.  V.  Biochim.  Biophys.  Acta  BBA  -­‐  

Enzymol.  1976,  438  (1),  1–12.  

(13)     Caruso,  M.  M.;  Davis,  D.  A.;  Shen,  Q.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Moore,  J.  S.  

Chem.  Rev.  2009,  109  (11),  5755–5798.  

(14)     Kuijpers,  M.  W.  A.;  Iedema,  P.  D.;  Kemmere,  M.  F.;  Keurentjes,  J.  T.  F.  Polymer  2004,  45  (19),  

6461–6467.  

(15)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

(16)     Rooze,  J.;  Groote,  R.;  Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.;  van  Iersel,  M.  M.;  Rebrov,  E.  V.;  

Schouten,  J.  C.;  Keurentjes,  J.  T.  F.  J.  Phys.  Chem.  B  2011,  115  (38),  11038–11043.  

(17)     Groote,  R.;  Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  ACS  Macro  Lett.  2012,  1  (8),  1012–1015.  

(18)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(19)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  Chem.  –  Eur.  J.  2009,  15  (48),  13331–13335.  

(20)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(21)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(22)     Encina,  M.  V.;  Lissi,  E.;  Sarasúa,  M.;  Gargallo,  L.;  Radic,  D.  J.  Polym.  Sci.  Polym.  Lett.  Ed.  1980,  

18  (12),  757–760.  

(23)     Berkowski,  K.  L.;  Potisek,  S.  L.;  Hickenboth,  C.  R.;  Moore,  J.  S.  Macromolecules  2005,  38  (22),  

8975–8978.  

(24)     Karthikeyan,  S.;  Potisek,  S.  L.;  Piermattei,  A.;  Sijbesma,  R.  P.  J.  Am.  Chem.  Soc.  2008,  130  (45),  

14968–14969.  

(25)     Kryger,  M.  J.;  Ong,  M.  T.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Martinez,  T.  J.;  Moore,  J.  S.  J.  

Am.  Chem.  Soc.  2010,  132  (13),  4558–4559.  

(26)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2008,  No.  37,  4416–4418.  

(27)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.  Int.  Ed.  2004,  43  (34),  4460–4462.  

(28)     Paulusse,  J.  M.  J.;  Huijbers,  J.  P.  J.;  Sijbesma,  R.  P.  Macromolecules  2005,  38  (15),  6290–6298.  

(29)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(30)     Klukovich,  H.  M.;  Kean,  Z.  S.;  Iacono,  S.  T.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2011,  133  (44),  17882–

17888.  

(31)     Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  Organometallics  2012,  31  (6),  2476–2481.  

(32)     Powell,  A.  B.;  Bielawski,  C.  W.;  Cowley,  A.  H.  J.  Am.  Chem.  Soc.  2010,  132  (29),  10184–10194.  

(33)     Davis,  D.  A.;  Hamilton,  A.;  Yang,  J.;  Cremar,  L.  D.;  Van  Gough,  D.;  Potisek,  S.  L.;  Ong,  M.  T.;  

Braun,  P.  V.;  Martínez,  T.  J.;  White,  S.  R.;  Moore,  J.  S.;  Sottos,  N.  R.  Nature  2009,  459  (7243),  

68–72.  

(34)     Kingsbury,  C.  M.;  May,  P.  A.;  Davis,  D.  A.;  White,  S.  R.;  Moore,  J.  S.;  Sottos,  N.  R.  J.  Mater.  

Chem.  2011,  21  (23),  8381–8388.  

(35)     O’Bryan,  G.;  Wong,  B.  M.;  McElhanon,  J.  R.  ACS  Appl.  Mater.  Interfaces  2010,  2  (6),  1594–

1600.  

(36)     Chen,  Y.;  Spiering,  A.  J.  H.;  Karthikeyan,  S.;  Peters,  G.  W.  M.;  Meijer,  E.  W.;  Sijbesma,  R.  P.  Nat.  

Chem.  2012,  4  (7),  559–562.  

(37)     Turro,  N.  J.;  Lechtken,  P.;  Schore,  N.  E.;  Schuster,  G.;  Steinmetzer,  H.  C.;  Yekta,  A.  Acc.  Chem.  

Res.  1974,  7  (4),  97–105.  

(38)     Schuster,  G.  B.;  Turro,  N.  J.;  Steinmetzer,  H.  C.;  Schaap,  A.  P.;  Faler,  G.;  Adam,  W.;  Liu,  J.  C.  J.  

Am.  Chem.  Soc.  1975,  97  (24),  7110–7118.  

(39)     Sottos,  N.  R.  Nat.  Chem.  2014,  6  (5),  381–383.  

(40)     Wiggins,  K.  M.;  Brantley,  J.  N.;  Bielawski,  C.  W.  Chem.  Soc.  Rev.  2013,  42  (17),  7130–7147.  

(41)     Diesendruck,  C.  E.;  Steinberg,  B.  D.;  Sugai,  N.;  Silberstein,  M.  N.;  Sottos,  N.  R.;  White,  S.  R.;  

Braun,  P.  V.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2012,  134  (30),  12446–12449.  

(42)     Larsen,  M.  B.;  Boydston,  A.  J.  J.  Am.  Chem.  Soc.  2013,  135  (22),  8189–8192.  

(43)     Larsen,  M.  B.;  Boydston,  A.  J.  J.  Am.  Chem.  Soc.  2014,  136  (4),  1276–1279.  

Chapter 1

18

Page 26: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

References  

(1)     Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat  Chem  2010,  2  (6),  436–437.  

(2)     Kauzmann,  W.;  Eyring,  H.  J.  Am.  Chem.  Soc.  1940,  62  (11),  3113–3125.  

(3)     Beyer,  M.  K.  J.  Chem.  Phys.  2000,  112  (17),  7307–7312.  

(4)     Groote,  R.;  Szyja,  B.  M.;  Pidko,  E.  A.;  Hensen,  E.  J.  M.;  Sijbesma,  R.  P.  Macromolecules  2011,  

44  (23),  9187–9195.  

(5)     Staudinger,  H.;  Heuer,  W.  Berichte  Dtsch.  Chem.  Ges.  B  Ser.  1934,  67  (7),  1159–1164.  

(6)     Hickenboth,  C.  R.;  Moore,  J.  S.;  White,  S.  R.;  Sottos,  N.  R.;  Baudry,  J.;  Wilson,  S.  R.  Nature  

2007,  446  (7134),  423–427.  

(7)     Ong,  M.  T.;  Leiding,  J.;  Tao,  H.;  Virshup,  A.  M.;  Martínez,  T.  J.  J.  Am.  Chem.  Soc.  2009,  131  (18),  

6377–6379.  

(8)     Potisek,  S.  L.;  Davis,  D.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2007,  129  

(45),  13808–13809.  

(9)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  J.  Am.  Chem.  Soc.  2010,  132  (30),  10609–10614.  

(10)     Black,  A.  L.;  Orlicki,  J.  A.;  Craig,  S.  L.  J.  Mater.  Chem.  2011,  21  (23),  8460–8465.  

(11)     Zocchi,  G.  Annu.  Rev.  Biophys.  2009,  38  (1),  75–88.  

(12)     Klibanov,  A.  M.;  Samokhin,  G.  P.;  Martinek,  K.;  Berezin,  I.  V.  Biochim.  Biophys.  Acta  BBA  -­‐  

Enzymol.  1976,  438  (1),  1–12.  

(13)     Caruso,  M.  M.;  Davis,  D.  A.;  Shen,  Q.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Moore,  J.  S.  

Chem.  Rev.  2009,  109  (11),  5755–5798.  

(14)     Kuijpers,  M.  W.  A.;  Iedema,  P.  D.;  Kemmere,  M.  F.;  Keurentjes,  J.  T.  F.  Polymer  2004,  45  (19),  

6461–6467.  

(15)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

(16)     Rooze,  J.;  Groote,  R.;  Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.;  van  Iersel,  M.  M.;  Rebrov,  E.  V.;  

Schouten,  J.  C.;  Keurentjes,  J.  T.  F.  J.  Phys.  Chem.  B  2011,  115  (38),  11038–11043.  

(17)     Groote,  R.;  Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  ACS  Macro  Lett.  2012,  1  (8),  1012–1015.  

(18)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(19)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  Chem.  –  Eur.  J.  2009,  15  (48),  13331–13335.  

(20)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(21)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(22)     Encina,  M.  V.;  Lissi,  E.;  Sarasúa,  M.;  Gargallo,  L.;  Radic,  D.  J.  Polym.  Sci.  Polym.  Lett.  Ed.  1980,  

18  (12),  757–760.  

(23)     Berkowski,  K.  L.;  Potisek,  S.  L.;  Hickenboth,  C.  R.;  Moore,  J.  S.  Macromolecules  2005,  38  (22),  

8975–8978.  

(24)     Karthikeyan,  S.;  Potisek,  S.  L.;  Piermattei,  A.;  Sijbesma,  R.  P.  J.  Am.  Chem.  Soc.  2008,  130  (45),  

14968–14969.  

(25)     Kryger,  M.  J.;  Ong,  M.  T.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Martinez,  T.  J.;  Moore,  J.  S.  J.  

Am.  Chem.  Soc.  2010,  132  (13),  4558–4559.  

(26)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2008,  No.  37,  4416–4418.  

(27)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.  Int.  Ed.  2004,  43  (34),  4460–4462.  

(28)     Paulusse,  J.  M.  J.;  Huijbers,  J.  P.  J.;  Sijbesma,  R.  P.  Macromolecules  2005,  38  (15),  6290–6298.  

(29)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(30)     Klukovich,  H.  M.;  Kean,  Z.  S.;  Iacono,  S.  T.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2011,  133  (44),  17882–

17888.  

(31)     Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  Organometallics  2012,  31  (6),  2476–2481.  

(32)     Powell,  A.  B.;  Bielawski,  C.  W.;  Cowley,  A.  H.  J.  Am.  Chem.  Soc.  2010,  132  (29),  10184–10194.  

(33)     Davis,  D.  A.;  Hamilton,  A.;  Yang,  J.;  Cremar,  L.  D.;  Van  Gough,  D.;  Potisek,  S.  L.;  Ong,  M.  T.;  

Braun,  P.  V.;  Martínez,  T.  J.;  White,  S.  R.;  Moore,  J.  S.;  Sottos,  N.  R.  Nature  2009,  459  (7243),  

68–72.  

(34)     Kingsbury,  C.  M.;  May,  P.  A.;  Davis,  D.  A.;  White,  S.  R.;  Moore,  J.  S.;  Sottos,  N.  R.  J.  Mater.  

Chem.  2011,  21  (23),  8381–8388.  

(35)     O’Bryan,  G.;  Wong,  B.  M.;  McElhanon,  J.  R.  ACS  Appl.  Mater.  Interfaces  2010,  2  (6),  1594–

1600.  

(36)     Chen,  Y.;  Spiering,  A.  J.  H.;  Karthikeyan,  S.;  Peters,  G.  W.  M.;  Meijer,  E.  W.;  Sijbesma,  R.  P.  Nat.  

Chem.  2012,  4  (7),  559–562.  

(37)     Turro,  N.  J.;  Lechtken,  P.;  Schore,  N.  E.;  Schuster,  G.;  Steinmetzer,  H.  C.;  Yekta,  A.  Acc.  Chem.  

Res.  1974,  7  (4),  97–105.  

(38)     Schuster,  G.  B.;  Turro,  N.  J.;  Steinmetzer,  H.  C.;  Schaap,  A.  P.;  Faler,  G.;  Adam,  W.;  Liu,  J.  C.  J.  

Am.  Chem.  Soc.  1975,  97  (24),  7110–7118.  

(39)     Sottos,  N.  R.  Nat.  Chem.  2014,  6  (5),  381–383.  

(40)     Wiggins,  K.  M.;  Brantley,  J.  N.;  Bielawski,  C.  W.  Chem.  Soc.  Rev.  2013,  42  (17),  7130–7147.  

(41)     Diesendruck,  C.  E.;  Steinberg,  B.  D.;  Sugai,  N.;  Silberstein,  M.  N.;  Sottos,  N.  R.;  White,  S.  R.;  

Braun,  P.  V.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2012,  134  (30),  12446–12449.  

(42)     Larsen,  M.  B.;  Boydston,  A.  J.  J.  Am.  Chem.  Soc.  2013,  135  (22),  8189–8192.  

(43)     Larsen,  M.  B.;  Boydston,  A.  J.  J.  Am.  Chem.  Soc.  2014,  136  (4),  1276–1279.  

Polymer Mechanochemistry

19

1

Page 27: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

(44)     Kean,  Z.  S.;  Akbulatov,  S.;  Tian,  Y.;  Widenhoefer,  R.  A.;  Boulatov,  R.;  Craig,  S.  L.  Angew.  Chem.  

Int.  Ed.  2014,  53  (52),  14508–14511.  

(45)     Fèvre,  M.;  Pinaud,  J.;  Gnanou,  Y.;  Vignolle,  J.;  Taton,  D.  Chem.  Soc.  Rev.  2013,  42  (5),  2142–

2172.  

(46)     Marion,  N.;  Díez-­‐González,  S.;  Nolan,  S.  P.  Angew.  Chem.  Int.  Ed.  2007,  46  (17),  2988–3000.  

(47)     Hopkinson,  M.  N.;  Richter,  C.;  Schedler,  M.;  Glorius,  F.  Nature  2014,  510  (7506),  485–496.  

(48)     Moore,  J.  L.;  Rovis,  T.  Top.  Curr.  Chem.  2010,  291.  

(49)     Bantu,  B.;  Pawar,  G.  M.;  Decker,  U.;  Wurst,  K.;  Schmidt,  A.  M.;  Buchmeiser,  M.  R.  Chem.  –  Eur.  

J.  2009,  15  (13),  3103–3109.  

(50)     Naumann,  S.;  Schmidt,  F.  G.;  Schowner,  R.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4  

(9),  2731–2740.  

(51)     Naumann,  S.;  Schmidt,  F.  G.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4  (15),  4172–

4181.  

(52)     Naumann,  S.;  Schmidt,  F.  G.;  Speiser,  M.;  Böhl,  M.;  Epple,  S.;  Bonten,  C.;  Buchmeiser,  M.  R.  

Macromolecules  2013,  46  (21),  8426–8433.  

(53)     Naumann,  S.;  Speiser,  M.;  Schowner,  R.;  Giebel,  E.;  Buchmeiser,  M.  R.  Macromolecules  2014,  

47  (14),  4548–4556.  

(54)     Kantchev,  E.  A.  B.;  O’Brien,  C.  J.;  Organ,  M.  G.  Angew.  Chem.  Int.  Ed.  2007,  46  (16),  2768–

2813.  

(55)     Monsaert,  S.;  Lozano  Vila,  A.;  Drozdzak,  R.;  Van  Der  Voort,  P.;  Verpoort,  F.  Chem.  Soc.  Rev.  

2009.  

(56)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(57)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(58)     Dias,  E.  L.;  Nguyen,  S.  T.;  Grubbs,  R.  H.  J.  Am.  Chem.  Soc.  1997,  119  (17),  3887–3897.  

(59)     Van  der  Schaaf,  P.  A.;  Kolly,  R.;  Kirner,  H.-­‐J.;  Rime,  F.;  Mühlebach,  A.;  Hafner,  A.  J.  Organomet.  

Chem.  2000,  606  (1),  65–74.  

(60)     Jakobs,  R.  T.  M.;  Ma,  S.;  Sijbesma,  R.  P.  ACS  Macro  Lett.  2013,  2  (7),  613–616.  

Chapter 1

20

Page 28: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter 2Mechanochemical chain scission in NHC-Pd centered coordination polymers

Ultrasound induced chain scission in coordination complexes of Palladium and Platinum

with polytetrahydrofuran functionalized N-heterocyclic carbene ligand is reported. Scission

is reversible when the polymer complex is sonicated in toluene under methane. The lowest

molecular weight complex that breaks was determined to be 20 kDa. Above this size the rate

of scission increases linearly with molecular weight of the polymer complex. Constrained

geometry simulations of external force (COGEF) calculations with DFT method provided

insight into the response of the NHC-Pd center to applied external force. Comparing scission

rates of Palladium and Platinum coordination polymers showed the influence of the force

needed for ligand dissociation.

(44)     Kean,  Z.  S.;  Akbulatov,  S.;  Tian,  Y.;  Widenhoefer,  R.  A.;  Boulatov,  R.;  Craig,  S.  L.  Angew.  Chem.  

Int.  Ed.  2014,  53  (52),  14508–14511.  

(45)     Fèvre,  M.;  Pinaud,  J.;  Gnanou,  Y.;  Vignolle,  J.;  Taton,  D.  Chem.  Soc.  Rev.  2013,  42  (5),  2142–

2172.  

(46)     Marion,  N.;  Díez-­‐González,  S.;  Nolan,  S.  P.  Angew.  Chem.  Int.  Ed.  2007,  46  (17),  2988–3000.  

(47)     Hopkinson,  M.  N.;  Richter,  C.;  Schedler,  M.;  Glorius,  F.  Nature  2014,  510  (7506),  485–496.  

(48)     Moore,  J.  L.;  Rovis,  T.  Top.  Curr.  Chem.  2010,  291.  

(49)     Bantu,  B.;  Pawar,  G.  M.;  Decker,  U.;  Wurst,  K.;  Schmidt,  A.  M.;  Buchmeiser,  M.  R.  Chem.  –  Eur.  

J.  2009,  15  (13),  3103–3109.  

(50)     Naumann,  S.;  Schmidt,  F.  G.;  Schowner,  R.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4  

(9),  2731–2740.  

(51)     Naumann,  S.;  Schmidt,  F.  G.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4  (15),  4172–

4181.  

(52)     Naumann,  S.;  Schmidt,  F.  G.;  Speiser,  M.;  Böhl,  M.;  Epple,  S.;  Bonten,  C.;  Buchmeiser,  M.  R.  

Macromolecules  2013,  46  (21),  8426–8433.  

(53)     Naumann,  S.;  Speiser,  M.;  Schowner,  R.;  Giebel,  E.;  Buchmeiser,  M.  R.  Macromolecules  2014,  

47  (14),  4548–4556.  

(54)     Kantchev,  E.  A.  B.;  O’Brien,  C.  J.;  Organ,  M.  G.  Angew.  Chem.  Int.  Ed.  2007,  46  (16),  2768–

2813.  

(55)     Monsaert,  S.;  Lozano  Vila,  A.;  Drozdzak,  R.;  Van  Der  Voort,  P.;  Verpoort,  F.  Chem.  Soc.  Rev.  

2009.  

(56)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(57)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(58)     Dias,  E.  L.;  Nguyen,  S.  T.;  Grubbs,  R.  H.  J.  Am.  Chem.  Soc.  1997,  119  (17),  3887–3897.  

(59)     Van  der  Schaaf,  P.  A.;  Kolly,  R.;  Kirner,  H.-­‐J.;  Rime,  F.;  Mühlebach,  A.;  Hafner,  A.  J.  Organomet.  

Chem.  2000,  606  (1),  65–74.  

(60)     Jakobs,  R.  T.  M.;  Ma,  S.;  Sijbesma,  R.  P.  ACS  Macro  Lett.  2013,  2  (7),  613–616.  

Page 29: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Mechanical   activation  of   chemical  bonds  promises   to  provide  opportunities   to  detect  and  

repair   damage   in   polymeric   materials.   Incorporation   of   mechanophores   into   polymers  

results   in   materials   sensitive   to   mechanical   stimuli   and   leads   to   useful   molecular  

transformations   under   stress.   In   solution,   mechanically   induced   chain   scission   has   been  

used  to  release  reactive  end  groups  such  as  cyanoacrylates,1  trifluorovinyl  ethers2  or  azides3  

from  mechanophore   precursors.   Furthermore,   transition   metal   complexes   located   in   the  

center   of   polymer   chains   have   been   dissociated   by   breaking   the   coordination   bonds  

mechanically,  as  an  alternative  to  thermal  activation  for  latent  catalysts.4–6      

One   of   the   most   efficient   ways   to   exert   force   on   a   polymer   in   solution   is   the   use   of  

sonication.7   Upon   sonication   of   solution,   cavitation   leads   to   strong   elongational   stresses  

around  collapsing  bubbles.8  The  part  of  the  polymer  chain  closest  to  the  collapsing  bubble  

wall  is  pulled  at  a  higher  velocity  than  the  far  end,  and  this  velocity  gradient  creates  stress  

along  the  backbone.  This  force  reaches  a  maximum  value  at  the  center  of  the  chain  because  

the  flow  field  is  centrosymmetric  with  respect  to  the  molecule  and  as  a  consequence  chain  

scission   occurs   preferentially   around   the   chain   midpoint9   for   polymers   above   a   critical  

molecular  weight.    

In   our   group,   high-­‐molecular-­‐weight   linear   coordination   polymers   of   diphenylphosphine  

telechelic  polytetrahydrofuran  with  palladium(II)  dichloride  have  been  studied.10  Molecular  

weights   of   these   polymers   could   be   altered   by   ultrasound   and   it   has   been   shown   that  

polytetrahydrofuran   chains   remain   intact   during   sonication.11   This   implies   that   only   the  

reversible   palladium–phosphorus   bonds   are   broken   and   coordinatively   unsaturated  

palladium  complexes   are   transiently  produced  by   the  application  of  mechanical   forces  on  

these   coordination   polymers.12   Furthermore,   polymers   which   include   both   PdII   and   PtII,  

were   sonicated   and   showed   that   mechanical   force   selectively   breaks   the   weaker   Pd-­‐

Phosphine  bonds  which  are  randomly  distributed  along  the  polymer  backbone.  Scission  rate  

is  influenced  by  small  changes  in  metal  ligand  coordination  bond  strength.13  

 

Phosphine   ligands   have   been   intensively   researched   in   Pd-­‐mediated   C-­‐C   and   C-­‐N   bond  

forming  reactions,  which  are  among  the  most  versatile  and  powerful  synthetic  methods.14  

However,   in  the   last  two  decades  the  popularity  of  phosphines  has  been  challenged  by  N-­‐

heterocyclic   carbenes   (NHCs).15   Compared   to   tertiary   phosphines,   NHC   ligands   show  

superior  thermal  stability  and  their  synthetic  availability  provides  the  possibility  to  extensive  

tuning   of   steric   and   electronic   properties.16–18   Palladium-­‐NHC   complexes   exhibit   high  

stability  both   in   solid   state  and   in   solution  due   to   strong  metal   ligand  bonds,  allowing   for  

easy   storage   and   handling.19   Organ   et.   al.   prepared   well-­‐defined   Pd-­‐PEPPSI-­‐(NHC)  

complexes  with  NHC  and  Pyridine  (Py)  ligands  coordinated  to  Pd(II).18,20  They  have  proposed  

that  Py  ligand  serves  as  a  ‘throw  away’  ligand  that  dissociates  from  the  Pd  center  in  catalyst  

activation  process.  Although   the  activation  mechanism  of   these   latent   catalysts   is   not   yet  

fully  understood,  it  has  been  shown  that  the  Pd:ligand  ratio,  optimally  1:1,21,22   is  crucial  to  

obtain  high  reactivity.23,24      

This   chapter   presents   the   ultrasound   induced   chain   scission   in   Pd(NHC)2   and   Pt(NHC)2  

complexes,   functionalized  with  poly(tetrahydrofuran)  as   force  accumulating  arms.  Scission  

rates  were  calculated  from  the  change  in  molecular  weight  distribution  of  polymers  during  

sonication.  Molecular  weight  dependence  of  the  scission  rate  was  established  and  the  effect  

of   bond   strength   on   scission   rates   were   determined   with   Constrained   geometry  

optimization   (COGEF)   calculations.   These   calculations   also   provided   a   more   in   depth  

understanding  on  the  structural  change  of  the  complex  under  the   influence  of  an  external  

force  that  leads  to  chain  scission.    

 

 

 

 

 

 

 

Chapter 2

22

Page 30: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Mechanical   activation  of   chemical  bonds  promises   to  provide  opportunities   to  detect  and  

repair   damage   in   polymeric   materials.   Incorporation   of   mechanophores   into   polymers  

results   in   materials   sensitive   to   mechanical   stimuli   and   leads   to   useful   molecular  

transformations   under   stress.   In   solution,   mechanically   induced   chain   scission   has   been  

used  to  release  reactive  end  groups  such  as  cyanoacrylates,1  trifluorovinyl  ethers2  or  azides3  

from  mechanophore   precursors.   Furthermore,   transition   metal   complexes   located   in   the  

center   of   polymer   chains   have   been   dissociated   by   breaking   the   coordination   bonds  

mechanically,  as  an  alternative  to  thermal  activation  for  latent  catalysts.4–6      

One   of   the   most   efficient   ways   to   exert   force   on   a   polymer   in   solution   is   the   use   of  

sonication.7   Upon   sonication   of   solution,   cavitation   leads   to   strong   elongational   stresses  

around  collapsing  bubbles.8  The  part  of  the  polymer  chain  closest  to  the  collapsing  bubble  

wall  is  pulled  at  a  higher  velocity  than  the  far  end,  and  this  velocity  gradient  creates  stress  

along  the  backbone.  This  force  reaches  a  maximum  value  at  the  center  of  the  chain  because  

the  flow  field  is  centrosymmetric  with  respect  to  the  molecule  and  as  a  consequence  chain  

scission   occurs   preferentially   around   the   chain   midpoint9   for   polymers   above   a   critical  

molecular  weight.    

In   our   group,   high-­‐molecular-­‐weight   linear   coordination   polymers   of   diphenylphosphine  

telechelic  polytetrahydrofuran  with  palladium(II)  dichloride  have  been  studied.10  Molecular  

weights   of   these   polymers   could   be   altered   by   ultrasound   and   it   has   been   shown   that  

polytetrahydrofuran   chains   remain   intact   during   sonication.11   This   implies   that   only   the  

reversible   palladium–phosphorus   bonds   are   broken   and   coordinatively   unsaturated  

palladium  complexes   are   transiently  produced  by   the  application  of  mechanical   forces  on  

these   coordination   polymers.12   Furthermore,   polymers   which   include   both   PdII   and   PtII,  

were   sonicated   and   showed   that   mechanical   force   selectively   breaks   the   weaker   Pd-­‐

Phosphine  bonds  which  are  randomly  distributed  along  the  polymer  backbone.  Scission  rate  

is  influenced  by  small  changes  in  metal  ligand  coordination  bond  strength.13  

 

Phosphine   ligands   have   been   intensively   researched   in   Pd-­‐mediated   C-­‐C   and   C-­‐N   bond  

forming  reactions,  which  are  among  the  most  versatile  and  powerful  synthetic  methods.14  

However,   in  the   last  two  decades  the  popularity  of  phosphines  has  been  challenged  by  N-­‐

heterocyclic   carbenes   (NHCs).15   Compared   to   tertiary   phosphines,   NHC   ligands   show  

superior  thermal  stability  and  their  synthetic  availability  provides  the  possibility  to  extensive  

tuning   of   steric   and   electronic   properties.16–18   Palladium-­‐NHC   complexes   exhibit   high  

stability  both   in   solid   state  and   in   solution  due   to   strong  metal   ligand  bonds,  allowing   for  

easy   storage   and   handling.19   Organ   et.   al.   prepared   well-­‐defined   Pd-­‐PEPPSI-­‐(NHC)  

complexes  with  NHC  and  Pyridine  (Py)  ligands  coordinated  to  Pd(II).18,20  They  have  proposed  

that  Py  ligand  serves  as  a  ‘throw  away’  ligand  that  dissociates  from  the  Pd  center  in  catalyst  

activation  process.  Although   the  activation  mechanism  of   these   latent   catalysts   is   not   yet  

fully  understood,  it  has  been  shown  that  the  Pd:ligand  ratio,  optimally  1:1,21,22   is  crucial  to  

obtain  high  reactivity.23,24      

This   chapter   presents   the   ultrasound   induced   chain   scission   in   Pd(NHC)2   and   Pt(NHC)2  

complexes,   functionalized  with  poly(tetrahydrofuran)  as   force  accumulating  arms.  Scission  

rates  were  calculated  from  the  change  in  molecular  weight  distribution  of  polymers  during  

sonication.  Molecular  weight  dependence  of  the  scission  rate  was  established  and  the  effect  

of   bond   strength   on   scission   rates   were   determined   with   Constrained   geometry  

optimization   (COGEF)   calculations.   These   calculations   also   provided   a   more   in   depth  

understanding  on  the  structural  change  of  the  complex  under  the   influence  of  an  external  

force  that  leads  to  chain  scission.    

 

 

 

 

 

 

 

Mechanochemical chain scission in NHC-Pd

23

2

Page 31: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  Discussions  

Synthesis  

 

Scheme   1:   Synthetic   route   to  mechanically   responsive   NHC-­‐Metal   complexes   (i)   DTBP   (ii)   1-­‐Ethyl  

imidazole,  20  min  (iii)  ion  exchange  resin,  MeOH    (iv)  NaOtBu,  THF  –  4Å  MS,  1h  (v)  M(PhCN)2Cl2  (M:  

Pd  or  Pt),  8h,  (vi)  C4H9I,  THF,  (vii)  ion  exchange  resin,  MeOH,  (viii)  PdCl2,  Cs2CO3,  dioxane,  5h.  

Polymeric  ligands  were  prepared  from  imidazolium  terminated  precursor  salts  (Scheme  1),  

which  were  obtained   via   a   cationic   ring  opening  polymerization  of   tetrahydrofuran   (THF).  

Polymerizations  were   initiated   by  methyl   triflate   and   subsequently   terminated   by  N-­‐ethyl  

imidazole  to  yield  (EtIm-­‐pTHF)Cl.25  Molecular  weights  (MW)  of  resulting  polymers  (Table  1)  

depend  on  the  polymerization  time  at   room  temperature,  and  were  analyzed  by   1H  NMR,  

MALDI-­‐TOF   and   gel   permeation   chromatography   (GPC).   The   triflate   counterions   were  

exchanged  by  chloride  over  ion  exchange  resin  in  methanol.    

 

 

 

OS CF3

OOH3C

O

i, ii, iii

O O N N

O O N NCl

M

OONN

ClCl

+

iv, v

(EtIm-pTHF)Cl

M: Pd Pd(NHC-pTHF)2Cl2M: Pt Pt(NHC-pTHF)2Cl2

(EtIm-pTHF)Cl

N N

Pd

NN

ClClN NCl

vi, vii viiiN N

(ButEtIm)ClEtImPd(NHC-ButEt)2Cl2

Table  1:  Molecular  weights  of  EtIm-­‐pTHF  determined  by  GPC,  NMR  and  MALDI-­‐TOF  

 

Polymer  

 

Time  (h)a  

 

Molecular  Weight  (kDa)  

GPCb   NMRc   MALDId  

(EtIm-­‐pTHF8k)Cl   1   19   9.4   8  

(EtIm-­‐pTHF12k)Cl   1.5   28   16   12  

(EtIm-­‐pTHF18k)Cl   2   48   21   18  

(EtIm-­‐pTHF25k)Cl   3   63   29   25  

(EtIm-­‐pTHF33k)Cl   4   72   38   33  

(EtIm-­‐pTHF40k)Cl   5   86   53   40  

(EtIm-­‐pTHF50k)Cl   6   108   63   50  

a)   polymerizations   were   initated   by   CH3CF3SO3   and   performed   at   20oC   for   given   time,   b)   2mg/ml   polymer  

solutions  in  THF  were  submitted  to  GPC  at  20oC,  1  ml/min  flow  rate,  PS  standards  were  used  for  calibration  c)  1H   NMR   spectra  were   taken   in   CD2Cl2   and  Molecular   weights  were   determined   by   end   group   analysis   d)   1  

mg/ml  polymer  solutions   in  CHCl3  were  submitted   to  MALDI-­‐TOF  and  peak   tops  were   reported  as  molecular  

weight.    

(EtIm-­‐pTHF)Cl   salts   were   deprotonated   on   the   imidazolium   C2   position   to   create   a   N-­‐

heterocyclic   carbene   (NHC-­‐pTHF)   which   subsequently   was   coordinated   to   Palladium   or  

Platinum   via   exchange   of   the   benzonitrile   (PhCN)   ligand   in   M(PhCN)2Cl2.   The   desired  

complexes  were  obtained   in  high  yield  since  PhCN  coordination   is  weak  and  allows   ligand  

exchange  at  room  temperature.  Coordination  of  two  NHC-­‐pTHF  ligands  resulted  in  doubling  

of  molecular  weight  compared  to  the  bare  ligands  as  monitored  by  GPC.  After  8h,  molecular  

weight   distributions   did   not   change   and   reactions   were   considered   as   completed.   MW  

distribution   of   polymers   reveals   bimodal   GPC   traces   due   to   small   amounts   of   pTHF   not  

terminated   by   EtIm.   In   order   to   determine   initial   weight   fraction   of   M(NHC-­‐pTHF)2Cl2  

complexes   bimodal   distributions   obtained   in   GPC   were   simulated   by   double   Gaussian  

function.  Areas  under   the  peaks  were  determined  and  used   to   calculate  weight   fractions.  

Ligand   coordination   was   also   evident   from   1H   NMR   which   showed   that   signals   for   H(1)  

proton   of   imidazole   heterocycle   had   disappeared   completely   and   H(4,5)   protons   shifted  

upfield  upon  coordination  to  metals.    

 

Chapter 2

24

Page 32: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  Discussions  

Synthesis  

 

Scheme   1:   Synthetic   route   to  mechanically   responsive   NHC-­‐Metal   complexes   (i)   DTBP   (ii)   1-­‐Ethyl  

imidazole,  20  min  (iii)  ion  exchange  resin,  MeOH    (iv)  NaOtBu,  THF  –  4Å  MS,  1h  (v)  M(PhCN)2Cl2  (M:  

Pd  or  Pt),  8h,  (vi)  C4H9I,  THF,  (vii)  ion  exchange  resin,  MeOH,  (viii)  PdCl2,  Cs2CO3,  dioxane,  5h.  

Polymeric  ligands  were  prepared  from  imidazolium  terminated  precursor  salts  (Scheme  1),  

which  were  obtained   via   a   cationic   ring  opening  polymerization  of   tetrahydrofuran   (THF).  

Polymerizations  were   initiated   by  methyl   triflate   and   subsequently   terminated   by  N-­‐ethyl  

imidazole  to  yield  (EtIm-­‐pTHF)Cl.25  Molecular  weights  (MW)  of  resulting  polymers  (Table  1)  

depend  on  the  polymerization  time  at   room  temperature,  and  were  analyzed  by   1H  NMR,  

MALDI-­‐TOF   and   gel   permeation   chromatography   (GPC).   The   triflate   counterions   were  

exchanged  by  chloride  over  ion  exchange  resin  in  methanol.    

 

 

 

OS CF3

OOH3C

O

i, ii, iii

O O N N

O O N NCl

M

OONN

ClCl

+

iv, v

(EtIm-pTHF)Cl

M: Pd Pd(NHC-pTHF)2Cl2M: Pt Pt(NHC-pTHF)2Cl2

(EtIm-pTHF)Cl

N N

Pd

NN

ClClN NCl

vi, vii viiiN N

(ButEtIm)ClEtImPd(NHC-ButEt)2Cl2

Table  1:  Molecular  weights  of  EtIm-­‐pTHF  determined  by  GPC,  NMR  and  MALDI-­‐TOF  

 

Polymer  

 

Time  (h)a  

 

Molecular  Weight  (kDa)  

GPCb   NMRc   MALDId  

(EtIm-­‐pTHF8k)Cl   1   19   9.4   8  

(EtIm-­‐pTHF12k)Cl   1.5   28   16   12  

(EtIm-­‐pTHF18k)Cl   2   48   21   18  

(EtIm-­‐pTHF25k)Cl   3   63   29   25  

(EtIm-­‐pTHF33k)Cl   4   72   38   33  

(EtIm-­‐pTHF40k)Cl   5   86   53   40  

(EtIm-­‐pTHF50k)Cl   6   108   63   50  

a)   polymerizations   were   initated   by   CH3CF3SO3   and   performed   at   20oC   for   given   time,   b)   2mg/ml   polymer  

solutions  in  THF  were  submitted  to  GPC  at  20oC,  1  ml/min  flow  rate,  PS  standards  were  used  for  calibration  c)  1H   NMR   spectra  were   taken   in   CD2Cl2   and  Molecular   weights  were   determined   by   end   group   analysis   d)   1  

mg/ml  polymer  solutions   in  CHCl3  were  submitted   to  MALDI-­‐TOF  and  peak   tops  were   reported  as  molecular  

weight.    

(EtIm-­‐pTHF)Cl   salts   were   deprotonated   on   the   imidazolium   C2   position   to   create   a   N-­‐

heterocyclic   carbene   (NHC-­‐pTHF)   which   subsequently   was   coordinated   to   Palladium   or  

Platinum   via   exchange   of   the   benzonitrile   (PhCN)   ligand   in   M(PhCN)2Cl2.   The   desired  

complexes  were  obtained   in  high  yield  since  PhCN  coordination   is  weak  and  allows   ligand  

exchange  at  room  temperature.  Coordination  of  two  NHC-­‐pTHF  ligands  resulted  in  doubling  

of  molecular  weight  compared  to  the  bare  ligands  as  monitored  by  GPC.  After  8h,  molecular  

weight   distributions   did   not   change   and   reactions   were   considered   as   completed.   MW  

distribution   of   polymers   reveals   bimodal   GPC   traces   due   to   small   amounts   of   pTHF   not  

terminated   by   EtIm.   In   order   to   determine   initial   weight   fraction   of   M(NHC-­‐pTHF)2Cl2  

complexes   bimodal   distributions   obtained   in   GPC   were   simulated   by   double   Gaussian  

function.  Areas  under   the  peaks  were  determined  and  used   to   calculate  weight   fractions.  

Ligand   coordination   was   also   evident   from   1H   NMR   which   showed   that   signals   for   H(1)  

proton   of   imidazole   heterocycle   had   disappeared   completely   and   H(4,5)   protons   shifted  

upfield  upon  coordination  to  metals.    

 

Mechanochemical chain scission in NHC-Pd

25

2

Page 33: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chain  Scission  in  Pd(NHC-­‐pTHF)2Cl2      

Chain   scission   in   polymers   Pd(NHC-­‐pTHF)2Cl2   was   investigated   in   toluene   solutions   (10  

mg/ml)  which  were  sonicated  for  60  min  using  a  sonication  probe.  During  sonication  Ar  or  

CH4  was  bubbled  through  the  solution  starting  15  minutes  prior  to  sonication.  The  double-­‐

jacketed  sonication  vessel  was  cooled  down  to  2oC  by  water  circulation  from  a  thermostat-­‐  

controlled  bath.  Temperature  inside  the  sonication  vessel  was  checked  by  a  thermocouple  

and  was  constant  at  25  oC  after  thermal  equilibration  (~  5  min).      

 

Figure   1:   Molecular   weight   change   during   sonication   of   Pd(NHC-­‐pTHF)2Cl2   in   the   presence   of  

trapping   agents   AcOH   and   MeCN   to   block   coordination,   monitored   by   GPC.   Polymer   solutions  

(10mg/ml)   in   dry   toluene   were   subjected   to   continuous   sonication   for   1h,   while   the   internal  

temperature   was   constant   at   25oC.   200   µL   aliquots   were   collected   at   given   times,   solvent   was  

evaporated   under   reduced   pressure;   residue  was   dissolved   in   THF   and   submitted   immediately   to  

GPC.  

Bond   rupture   on  mechano-­‐responsive   organometallic   unit   at   the   center   of   polymer   chain  

yields   free   NHC   and   coordinatively   unsaturated   Pd.   Reversibility   of   the   ligand-­‐metal  

interactions   might   then   result   in   re-­‐coordination   of   sonication   products.   Therefore,  

sonication   was   performed   in   the   presence   of   acetic   acid   (AcOH)   (2.4   %,   w/w)   and  

acetonitrile  (MeCN)  (1.8  %,  w/w)  as  trapping  agents,  to  stabilize  the  scission  products  and  

prevent   the   re-­‐coordination.   The   stability   of   the   Pd-­‐NHC   coordination   complexes   in   the  

presence  of  these  trapping  agents  was  established  by  stirring  in  toluene  in  the  presence  of  

AcOH  and  MeCN  at  room  temperature.  In  the  course  of  1h,  no  change  in  MW  was  observed  

with  GPC.    

 

Figure   2:   Reversibility   test   of   Pd(NHC-­‐pTHF18k)2Cl2:   a)   under   CH4   b)   under   Ar.   c)   Comparison   of  

change   in  molecular   weight   distribution   of   Pd(NHC-­‐pTHF18k)2Cl2   as   a   function   of   sonication   time  

with  and  without  trapping  agents  under  CH4  and  Ar.  d)  Ultrasound  induced  reversible  chain  scission  

of   the   Pd-­‐NHC   bond.   Samples   with   a   concentration   of   10   mg/ml   were   subjected   to   continuous  

sonication  in  dry  toluene  for  1h.  In  the  absence  of  trapping  agents  molecular  weight  distribution  did  

not   change   under   CH4   in   contrast   to   sonication   under   Ar.   Solid   and   dashed   lines   represent   GPC  

traces  for  Pd(NHC-­‐pTHF18k)2Cl2  before  and  after  sonication  respectively.  

After  sonication  in  the  presence  of  the  trapping  agents,  GPC  traces  showed  the  formation  of  

species  with  half  of  the  initial  MW,  confirming  that  scission  occurs  selectively  at  the  central  

bonds  (Figure  1).  In  the  presence  of  trapping  agents,  1h  of  continuous  sonication  resulted  in  

approximately  55%  chain  scission  for  Pd(NHC-­‐pTHF18k)2Cl2.  However,  without  any  trapping  

agent  under  CH4,  no  significant  change   in  molecular  weight  was  observed  after  1h   (Figure  

Chapter 2

26

Page 34: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chain  Scission  in  Pd(NHC-­‐pTHF)2Cl2      

Chain   scission   in   polymers   Pd(NHC-­‐pTHF)2Cl2   was   investigated   in   toluene   solutions   (10  

mg/ml)  which  were  sonicated  for  60  min  using  a  sonication  probe.  During  sonication  Ar  or  

CH4  was  bubbled  through  the  solution  starting  15  minutes  prior  to  sonication.  The  double-­‐

jacketed  sonication  vessel  was  cooled  down  to  2oC  by  water  circulation  from  a  thermostat-­‐  

controlled  bath.  Temperature  inside  the  sonication  vessel  was  checked  by  a  thermocouple  

and  was  constant  at  25  oC  after  thermal  equilibration  (~  5  min).      

 

Figure   1:   Molecular   weight   change   during   sonication   of   Pd(NHC-­‐pTHF)2Cl2   in   the   presence   of  

trapping   agents   AcOH   and   MeCN   to   block   coordination,   monitored   by   GPC.   Polymer   solutions  

(10mg/ml)   in   dry   toluene   were   subjected   to   continuous   sonication   for   1h,   while   the   internal  

temperature   was   constant   at   25oC.   200   µL   aliquots   were   collected   at   given   times,   solvent   was  

evaporated   under   reduced   pressure;   residue  was   dissolved   in   THF   and   submitted   immediately   to  

GPC.  

Bond   rupture   on  mechano-­‐responsive   organometallic   unit   at   the   center   of   polymer   chain  

yields   free   NHC   and   coordinatively   unsaturated   Pd.   Reversibility   of   the   ligand-­‐metal  

interactions   might   then   result   in   re-­‐coordination   of   sonication   products.   Therefore,  

sonication   was   performed   in   the   presence   of   acetic   acid   (AcOH)   (2.4   %,   w/w)   and  

acetonitrile  (MeCN)  (1.8  %,  w/w)  as  trapping  agents,  to  stabilize  the  scission  products  and  

prevent   the   re-­‐coordination.   The   stability   of   the   Pd-­‐NHC   coordination   complexes   in   the  

presence  of  these  trapping  agents  was  established  by  stirring  in  toluene  in  the  presence  of  

AcOH  and  MeCN  at  room  temperature.  In  the  course  of  1h,  no  change  in  MW  was  observed  

with  GPC.    

 

Figure   2:   Reversibility   test   of   Pd(NHC-­‐pTHF18k)2Cl2:   a)   under   CH4   b)   under   Ar.   c)   Comparison   of  

change   in  molecular   weight   distribution   of   Pd(NHC-­‐pTHF18k)2Cl2   as   a   function   of   sonication   time  

with  and  without  trapping  agents  under  CH4  and  Ar.  d)  Ultrasound  induced  reversible  chain  scission  

of   the   Pd-­‐NHC   bond.   Samples   with   a   concentration   of   10   mg/ml   were   subjected   to   continuous  

sonication  in  dry  toluene  for  1h.  In  the  absence  of  trapping  agents  molecular  weight  distribution  did  

not   change   under   CH4   in   contrast   to   sonication   under   Ar.   Solid   and   dashed   lines   represent   GPC  

traces  for  Pd(NHC-­‐pTHF18k)2Cl2  before  and  after  sonication  respectively.  

After  sonication  in  the  presence  of  the  trapping  agents,  GPC  traces  showed  the  formation  of  

species  with  half  of  the  initial  MW,  confirming  that  scission  occurs  selectively  at  the  central  

bonds  (Figure  1).  In  the  presence  of  trapping  agents,  1h  of  continuous  sonication  resulted  in  

approximately  55%  chain  scission  for  Pd(NHC-­‐pTHF18k)2Cl2.  However,  without  any  trapping  

agent  under  CH4,  no  significant  change   in  molecular  weight  was  observed  after  1h   (Figure  

Mechanochemical chain scission in NHC-Pd

27

2

Page 35: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

2).   This   indicates   that   chain   scission   is   selective   to   NHC-­‐metal   coordination   bond   and  

without   trapping   agents   scission   is   completely   reversible.   Control   experiments   were  

performed  with   low  molecular  weight   complex   Pd(NHC-­‐But)2Cl2    to   investigate   the   role   of  

thermal  effects  in  scission.  Ultrasound  did  not  cause  any  changes  when  Pd(NHC-­‐But)2Cl2  was  

sonicated  with  trapping  agents   for  1h  (under  CH4  or  Ar).  The  amount  of   total   imidazolium  

after  1h  of  sonication  of  the  36k  polymer  is  the  same  regardless  of  saturation  gas;  however,  

under   Ar   sonication   yields   up   to   30%   irreversible   chain   scission.   Further   reaction   of   the  

scission  products   (free  NHC  and   coordinatively  unsaturated  Pd   center)  with   sonochemical  

impurities   which   are   produced   due   to   possible   solvent   pyrolysis   in   hot   spots,26   prevents  

reversible  recoordination  under  argon.  The  use  of  CH4  instead  of  Ar  as  the  saturation  gas  led  

to   suppression   of   thermal   effects   during   sonication   and   reversible   scission   on   NHC-­‐Pd  

coordination  bond  with  increased  lifetime  of  mechanically  released  free  NHC  in  solution.    

1H   NMR   spectra,   taken   for   Pd(NHC-­‐pTHF25k)2Cl2   samples   before   and   after   sonication,  

showed   that   the   sonication   revealed   two   new   sets   of   peaks   (Figure   3).   These   peaks   are  

assumed   to   be   due   to   free   NHC   that   is   taken   off   and   subsequently   protonated   by   the  

trapping   agent,   and   Pd   center   with  monoNHC.   Since   the   NHC   taken   off   by   ultrasound   is  

prevented  from  re-­‐coordination,  the  remaining  monoNHC-­‐Pd  part  should  be  coordinated  by  

another   ligand  presents   in   the   solution.  MeCN  was  used   to   serve  as   the   stabilizing   ligand  

during   the   course   of   sonication.   However,   NMR   spectra   showed   that   MeCN   was   not  

coordinated.  It  is  known  for  hetero  NHC  complexes  of  Pd  that  a  stronger  donating  co-­‐ligand  

leads   to   downfield   shift   in   NMR   spectrum   of   the   NHC   ligand.27   Since  MeCN   is   a   weaker  

donor  than  NHC,  shift  in  opposite  direction  would  have  been  expected  upon  coordination  of  

MeCN.  Thus,   it   is  not  trivial  at   this  point  to  determine  the  exact  chemical  structure  of   the  

scission   product   that   contains   Pd,   since   coordinatively   unsaturated   Pd   center   could  

coordinate  to  any  other  ligand  (i.e.  acetate  anion)  or  dimerize28  to  stabilize  its  structure.    

 

Figure  3:  Stacked  1H  NMR  spectra  for  (a)  EtIm-­‐pTHF25k    (b)  Pd(NHC-­‐pTHF25k)2Cl2    and  (c)  Pd(NHC-­‐

pTHF25k)2Cl2  after  sonication.  

Determination  of  limiting  molecular  weight  Mlim  

It   is   well   established   that   mechanochemical   scission   of   polymers   only   occurs   above   a  

molecular   weight   threshold   (Mlim).29   Below   Mlim,   no   scission   takes   place   since   polymer  

chains   are   too   short   to   accumulate   the   force   required   to   break   chemical   bonds.30   Mlim  

lowers   significantly   if   a   weak   bond   is   incorporated   into   a   polymer   chain.   Coordination  

bonds,  weaker   than  covalent  bonds  on  polymer  backbone  break  more  easily  and  result   in  

lower   Mlim   compared   to   their   covalent   counterparts.13   In   addition   to   bond   strength,  

experimental   conditions   such   as   ultrasound   intensity,   solvent   composition,   temperature,  

and   gas   molecules   dissolved   in   solution   also   influence   Mlim.31   In   order   to   compare  

mechanochemical  properties  of  different  mechanophores,  experimental   conditions   should  

be   the   same.   Therefore   we   decided   to   investigate   Mlim   and   the   molecular   weight  

dependence   of   scission   rate   for   a   specific   set   of   sonication   conditions,   and   to   correlate  

scissability  with  bond  dissociation  energies  for  the  Metal-­‐NHC  (Pd,  Pt)  complexes.      

Chapter 2

28

Page 36: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

2).   This   indicates   that   chain   scission   is   selective   to   NHC-­‐metal   coordination   bond   and  

without   trapping   agents   scission   is   completely   reversible.   Control   experiments   were  

performed  with   low  molecular  weight   complex   Pd(NHC-­‐But)2Cl2    to   investigate   the   role   of  

thermal  effects  in  scission.  Ultrasound  did  not  cause  any  changes  when  Pd(NHC-­‐But)2Cl2  was  

sonicated  with  trapping  agents   for  1h  (under  CH4  or  Ar).  The  amount  of   total   imidazolium  

after  1h  of  sonication  of  the  36k  polymer  is  the  same  regardless  of  saturation  gas;  however,  

under   Ar   sonication   yields   up   to   30%   irreversible   chain   scission.   Further   reaction   of   the  

scission  products   (free  NHC  and   coordinatively  unsaturated  Pd   center)  with   sonochemical  

impurities   which   are   produced   due   to   possible   solvent   pyrolysis   in   hot   spots,26   prevents  

reversible  recoordination  under  argon.  The  use  of  CH4  instead  of  Ar  as  the  saturation  gas  led  

to   suppression   of   thermal   effects   during   sonication   and   reversible   scission   on   NHC-­‐Pd  

coordination  bond  with  increased  lifetime  of  mechanically  released  free  NHC  in  solution.    

1H   NMR   spectra,   taken   for   Pd(NHC-­‐pTHF25k)2Cl2   samples   before   and   after   sonication,  

showed   that   the   sonication   revealed   two   new   sets   of   peaks   (Figure   3).   These   peaks   are  

assumed   to   be   due   to   free   NHC   that   is   taken   off   and   subsequently   protonated   by   the  

trapping   agent,   and   Pd   center   with  monoNHC.   Since   the   NHC   taken   off   by   ultrasound   is  

prevented  from  re-­‐coordination,  the  remaining  monoNHC-­‐Pd  part  should  be  coordinated  by  

another   ligand  presents   in   the   solution.  MeCN  was  used   to   serve  as   the   stabilizing   ligand  

during   the   course   of   sonication.   However,   NMR   spectra   showed   that   MeCN   was   not  

coordinated.  It  is  known  for  hetero  NHC  complexes  of  Pd  that  a  stronger  donating  co-­‐ligand  

leads   to   downfield   shift   in   NMR   spectrum   of   the   NHC   ligand.27   Since  MeCN   is   a   weaker  

donor  than  NHC,  shift  in  opposite  direction  would  have  been  expected  upon  coordination  of  

MeCN.  Thus,   it   is  not  trivial  at   this  point  to  determine  the  exact  chemical  structure  of   the  

scission   product   that   contains   Pd,   since   coordinatively   unsaturated   Pd   center   could  

coordinate  to  any  other  ligand  (i.e.  acetate  anion)  or  dimerize28  to  stabilize  its  structure.    

 

Figure  3:  Stacked  1H  NMR  spectra  for  (a)  EtIm-­‐pTHF25k    (b)  Pd(NHC-­‐pTHF25k)2Cl2    and  (c)  Pd(NHC-­‐

pTHF25k)2Cl2  after  sonication.  

Determination  of  limiting  molecular  weight  Mlim  

It   is   well   established   that   mechanochemical   scission   of   polymers   only   occurs   above   a  

molecular   weight   threshold   (Mlim).29   Below   Mlim,   no   scission   takes   place   since   polymer  

chains   are   too   short   to   accumulate   the   force   required   to   break   chemical   bonds.30   Mlim  

lowers   significantly   if   a   weak   bond   is   incorporated   into   a   polymer   chain.   Coordination  

bonds,  weaker   than  covalent  bonds  on  polymer  backbone  break  more  easily  and  result   in  

lower   Mlim   compared   to   their   covalent   counterparts.13   In   addition   to   bond   strength,  

experimental   conditions   such   as   ultrasound   intensity,   solvent   composition,   temperature,  

and   gas   molecules   dissolved   in   solution   also   influence   Mlim.31   In   order   to   compare  

mechanochemical  properties  of  different  mechanophores,  experimental   conditions   should  

be   the   same.   Therefore   we   decided   to   investigate   Mlim   and   the   molecular   weight  

dependence   of   scission   rate   for   a   specific   set   of   sonication   conditions,   and   to   correlate  

scissability  with  bond  dissociation  energies  for  the  Metal-­‐NHC  (Pd,  Pt)  complexes.      

Mechanochemical chain scission in NHC-Pd

29

2

Page 37: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

In   order   to   prevent   complication   of   the   analysis   from   scission   of   covalent   bonds,   the  

polymeric   ligand  must   have   a  molecular  weight   that   is   lower   than  Mlim  for   covalent   bond  

scission.   Therefore,   the   limiting  molecular   weight   of   covalent   pTHF   was   determined   in   a  

separate   set  of  experiments.   EtIm-­‐pTHF   samples  with  Mw  ranging  between  8  and  50  kDa  

were   subjected   to   ultrasound   at   20   kHz   with   a   power   density   of   15.4  W/cm2   in   toluene  

solutions   (10   mg/ml)   under   CH4.   Internal   temperature   of   the   double-­‐jacketed   sonication  

vessel  was  constant  at  25oC  (±2  oC)  throughout  sonication.    

GPC   traces   for   polymer   samples   before   and   after   sonication   showed   that   the   molecular  

weight  distribution  for  EtIm-­‐pTHF  below  25  kDa  did  not  change  significantly  by  sonication.  

However,  polymer  with  a  MW  higher  than  30  kDa  degraded  irreversibly  as  shown  by  both  

GPC   (Figure   4)   and   MALDI-­‐TOF   (Figure   5).   This   sets   a   limit   (Mlim)   to   the   MW   of   force-­‐

accumulating  ligands  (EtIm-­‐pTHF)  below  which  they  are  not  destroyed  by  ultrasound.  Thus,  

only   ligands   with   a   molecular   weight   lower   than   Mlim   for   pTHF   (<30kDa)   were   used   to  

synthesize  mechano-­‐responsive  polymers  with  a  central  Pd-­‐NHC  mechanophore.    

 

Figure  4:  GPC  traces  for  polymer  samples  of  EtIm-­‐pTHF  taken  before  and  after  sonication.  10  mg/ml  

samples  were   sonicated   (continuous)   in   toluene  under  CH4.    Aliquots  were   taken  and   solvent  was  

evaporated,   residue   was   dissolved   in   THF   and   directly   submitted   to   GPC.   Black   and   red   lines  

represented  traces  for  aliquots  taken  before  and  after  sonication  samples.  Curves  were  normalized  

to  peak  areas.    

 

Figure  5:  MALDI-­‐TOF  spectra  for  EtIm-­‐pTHF50k  before  (left)  and  after  (right)  1h  of  sonication.  

The   GPC   trace   of   Pd(NHC-­‐pTHF8k)2Cl2   did   not   change   after   1   h   of   continuous   sonication  

under   CH4   in   the   presence   of   acetonitrile   and   acetic   acid   as   trapping   agents   (Figure   6).  

However,   higher   molecular   weight   complex   Pd(NHC-­‐pTHF25k)2Cl2   showed   around   %80  

chain   scission   after   1h   of   continuous   sonication.   Mechanochemical   fragmentation   of  

Pd(NHC-­‐pTHF25k)2Cl2  revealed  a  product  with  a  sharp  peak  corresponding  to  MW  of  25  kDa,  

that   is   consistent   with   the   previously   reported   center   specific   chain   scission   for  

mechanophore   incorporated   polymers.32   This   observation   confirms   that  Mlim   for   Pd(NHC-­‐

pTHF)Cl2   complex   is   higher   than   16   kDa   under   above   mentioned   conditions   and   central  

coordination  bond  remains  intact  when  polymer  chain  is  not  sufficiently  long.  

Chapter 2

30

Page 38: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

In   order   to   prevent   complication   of   the   analysis   from   scission   of   covalent   bonds,   the  

polymeric   ligand  must   have   a  molecular  weight   that   is   lower   than  Mlim  for   covalent   bond  

scission.   Therefore,   the   limiting  molecular   weight   of   covalent   pTHF   was   determined   in   a  

separate   set  of  experiments.   EtIm-­‐pTHF   samples  with  Mw  ranging  between  8  and  50  kDa  

were   subjected   to   ultrasound   at   20   kHz   with   a   power   density   of   15.4  W/cm2   in   toluene  

solutions   (10   mg/ml)   under   CH4.   Internal   temperature   of   the   double-­‐jacketed   sonication  

vessel  was  constant  at  25oC  (±2  oC)  throughout  sonication.    

GPC   traces   for   polymer   samples   before   and   after   sonication   showed   that   the   molecular  

weight  distribution  for  EtIm-­‐pTHF  below  25  kDa  did  not  change  significantly  by  sonication.  

However,  polymer  with  a  MW  higher  than  30  kDa  degraded  irreversibly  as  shown  by  both  

GPC   (Figure   4)   and   MALDI-­‐TOF   (Figure   5).   This   sets   a   limit   (Mlim)   to   the   MW   of   force-­‐

accumulating  ligands  (EtIm-­‐pTHF)  below  which  they  are  not  destroyed  by  ultrasound.  Thus,  

only   ligands   with   a   molecular   weight   lower   than   Mlim   for   pTHF   (<30kDa)   were   used   to  

synthesize  mechano-­‐responsive  polymers  with  a  central  Pd-­‐NHC  mechanophore.    

 

Figure  4:  GPC  traces  for  polymer  samples  of  EtIm-­‐pTHF  taken  before  and  after  sonication.  10  mg/ml  

samples  were   sonicated   (continuous)   in   toluene  under  CH4.    Aliquots  were   taken  and   solvent  was  

evaporated,   residue   was   dissolved   in   THF   and   directly   submitted   to   GPC.   Black   and   red   lines  

represented  traces  for  aliquots  taken  before  and  after  sonication  samples.  Curves  were  normalized  

to  peak  areas.    

 

Figure  5:  MALDI-­‐TOF  spectra  for  EtIm-­‐pTHF50k  before  (left)  and  after  (right)  1h  of  sonication.  

The   GPC   trace   of   Pd(NHC-­‐pTHF8k)2Cl2   did   not   change   after   1   h   of   continuous   sonication  

under   CH4   in   the   presence   of   acetonitrile   and   acetic   acid   as   trapping   agents   (Figure   6).  

However,   higher   molecular   weight   complex   Pd(NHC-­‐pTHF25k)2Cl2   showed   around   %80  

chain   scission   after   1h   of   continuous   sonication.   Mechanochemical   fragmentation   of  

Pd(NHC-­‐pTHF25k)2Cl2  revealed  a  product  with  a  sharp  peak  corresponding  to  MW  of  25  kDa,  

that   is   consistent   with   the   previously   reported   center   specific   chain   scission   for  

mechanophore   incorporated   polymers.32   This   observation   confirms   that  Mlim   for   Pd(NHC-­‐

pTHF)Cl2   complex   is   higher   than   16   kDa   under   above   mentioned   conditions   and   central  

coordination  bond  remains  intact  when  polymer  chain  is  not  sufficiently  long.  

Mechanochemical chain scission in NHC-Pd

31

2

Page 39: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   6:   GPC   traces   for   polymer   samples   taken   before   and   after   sonication.   a)   EtIm-­‐pTHF18k,   b)  

EtIm-­‐pTHF50k,  c)  Pd(NHC-­‐pTHF8k)2Cl2    and  d)  Pd(NHC-­‐pTHF25k)2Cl2.  Palladium  containing  polymers,  

Pd(NHC-­‐pTHF)2Cl2,  were  sonicated  in  the  presence  of  trapping  agents.  

Scission  Rates  for  M(NHC-­‐pTHF)2Cl2      

The   rate   of   mechanically   induced   scission   in   covalent   polymers   has   been   shown   to   be  

proportional  to  [MW  -­‐  Mlim]  when  the  initial  molecular  weight  is  greater  than  Mlim32,33  When  

such  a  molecular  weight  dependence  is  observed,  thermal  decomposition  can  be  excluded,  

because  it  proves  that  bond  rupture  process  is  mechano-­‐chemical  and  not  the  result  of  local  

heating   or   reaction   with   solvent   decomposition   products.34   The   increase   of   scission   rate  

with  increasing  molecular  weight  is  caused  by  the  larger  accumulated  elongational  stress  at  

the  center  of   long  polymer  chains,   in  combination  with  their   longer  relaxation  time,  which  

reduces  the  rate  of  stress  relaxation  by  chain  motion.35  

 

Figure  7:  Real  data  for  GPC  traces  and  deconvolution  curves  represented  with  black  and  gray   lines  

respectively  for  initial  MW  distribution  of  (a)  Pd(NHC-­‐pTHF25k)2Cl2    and  (b)  Pt(NHC-­‐pTHF25k)2Cl2.  

Molecular  weight  dependence  of   the  mechanochemical   scission   rate   for  M(NHC-­‐pTHF)2Cl2    

complexes   were   determined   by   sonicating   samples   separately.   Molecular   weight  

distributions   for   collected  aliquots   throughout   sonication  were  determined  by  GPC   in  THF  

and  peak  areas  were  calculated  by  deconvolution  of  traces  with  a  bimodal  molecular  weight  

distribution  (Figure  7).  The  concentration  change  of  polymer  complex  during  sonication  was  

estimated   under   the   assumption   that   peak   area   is   proportional   to   weight   fraction   in   RI  

detection.   The   scission   rates   of   the   Pd   and   Pt   complexes   were   fitted   with   equation   3,  

assuming  first  order  reaction  kinetics  since  chain  scission  is  selective  to  chain  mid-­‐point.  

  𝑃𝑃 𝑥𝑥 → 2𝑃𝑃 𝑥𝑥 2     (1)  

 

   − ! ! !!"

= 𝑘𝑘!"[𝑃𝑃 𝑥𝑥 ]     (2)  

where   P(x)   is   initial   polymer   and   P(x/2)   is   the   fragmentation   product.   Thus,   degradation  follows  first-­‐order  kinetics  and  weight  fraction  of  starting  material  C1  decays  exponentially  with  the  equation;    

  𝐶𝐶! 𝑡𝑡 = 𝐶𝐶! 0 ×𝑒𝑒 !!!"!   (3)  

where,  t    is  sonication  time,  C1(0)  is  initial  weight  fraction  of  starting  material,  C1(t)  is  weight  fraction   of   starting   material   at   any   time   during   sonication   and   ksc   is   mechanochemical  scission  rate  coefficient.  

Chapter 2

32

Page 40: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   6:   GPC   traces   for   polymer   samples   taken   before   and   after   sonication.   a)   EtIm-­‐pTHF18k,   b)  

EtIm-­‐pTHF50k,  c)  Pd(NHC-­‐pTHF8k)2Cl2    and  d)  Pd(NHC-­‐pTHF25k)2Cl2.  Palladium  containing  polymers,  

Pd(NHC-­‐pTHF)2Cl2,  were  sonicated  in  the  presence  of  trapping  agents.  

Scission  Rates  for  M(NHC-­‐pTHF)2Cl2      

The   rate   of   mechanically   induced   scission   in   covalent   polymers   has   been   shown   to   be  

proportional  to  [MW  -­‐  Mlim]  when  the  initial  molecular  weight  is  greater  than  Mlim32,33  When  

such  a  molecular  weight  dependence  is  observed,  thermal  decomposition  can  be  excluded,  

because  it  proves  that  bond  rupture  process  is  mechano-­‐chemical  and  not  the  result  of  local  

heating   or   reaction   with   solvent   decomposition   products.34   The   increase   of   scission   rate  

with  increasing  molecular  weight  is  caused  by  the  larger  accumulated  elongational  stress  at  

the  center  of   long  polymer  chains,   in  combination  with  their   longer  relaxation  time,  which  

reduces  the  rate  of  stress  relaxation  by  chain  motion.35  

 

Figure  7:  Real  data  for  GPC  traces  and  deconvolution  curves  represented  with  black  and  gray   lines  

respectively  for  initial  MW  distribution  of  (a)  Pd(NHC-­‐pTHF25k)2Cl2    and  (b)  Pt(NHC-­‐pTHF25k)2Cl2.  

Molecular  weight  dependence  of   the  mechanochemical   scission   rate   for  M(NHC-­‐pTHF)2Cl2    

complexes   were   determined   by   sonicating   samples   separately.   Molecular   weight  

distributions   for   collected  aliquots   throughout   sonication  were  determined  by  GPC   in  THF  

and  peak  areas  were  calculated  by  deconvolution  of  traces  with  a  bimodal  molecular  weight  

distribution  (Figure  7).  The  concentration  change  of  polymer  complex  during  sonication  was  

estimated   under   the   assumption   that   peak   area   is   proportional   to   weight   fraction   in   RI  

detection.   The   scission   rates   of   the   Pd   and   Pt   complexes   were   fitted   with   equation   3,  

assuming  first  order  reaction  kinetics  since  chain  scission  is  selective  to  chain  mid-­‐point.  

  𝑃𝑃 𝑥𝑥 → 2𝑃𝑃 𝑥𝑥 2     (1)  

 

   − ! ! !!"

= 𝑘𝑘!"[𝑃𝑃 𝑥𝑥 ]     (2)  

where   P(x)   is   initial   polymer   and   P(x/2)   is   the   fragmentation   product.   Thus,   degradation  follows  first-­‐order  kinetics  and  weight  fraction  of  starting  material  C1  decays  exponentially  with  the  equation;    

  𝐶𝐶! 𝑡𝑡 = 𝐶𝐶! 0 ×𝑒𝑒 !!!"!   (3)  

where,  t    is  sonication  time,  C1(0)  is  initial  weight  fraction  of  starting  material,  C1(t)  is  weight  fraction   of   starting   material   at   any   time   during   sonication   and   ksc   is   mechanochemical  scission  rate  coefficient.  

Mechanochemical chain scission in NHC-Pd

33

2

Page 41: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   8:   Change   in   initial   polymer   concentration  vs.   sonication   time   for  M(NHC-­‐pTHF18k)2Cl2   (M:  

Pd,  and  Pt).    

M(NHC-­‐pTHF)2Cl2  samples  were  sonicated  separately  and  their  scission  rate  constants  (ksc)  

were   determined  by   exponential   fitting.   The  ksc   vs  MW  plot   in   Figure   8   shows   that   chain  

scission   in   Pt   containing   polymers   is   slower   than   in   Pd   containing   counterparts   for   each  

initial  MW.  Although  the  calculated  ksc’s  are  linearly  dependent  on  MW  for  both  metals,  the  

slopes  of  the  lines  are  different.  Madras  et  al.  used  the  following  equation  to  express  ksc  as  a  

function  of  MW:35    

  𝑘𝑘!"  =  𝑘𝑘!  (𝑀𝑀𝑀𝑀   −  𝑀𝑀!"#)ƛ   (4)  

In   which  𝑘𝑘!  is   an   empirical   proportionality   constant.   In   the   Pd   and   Pt   complexes,   the  

molecular  weight  dependent  scission  rates  fit  well  to  equation  4  with  exponent  λ  equal  to  1.  

Mlim  values  for  both  mechanophores  were  estimated  from  linear  extrapolation  of  the  plots  

of   ksc   vs   MW   (Figure   10)   as   20   kDa   and   22   kDa   for   Pd(NHC-­‐pTHF)2Cl2   and   for   Pt(NHC-­‐

pTHF)2Cl2  (Table  2)  respectively.    

 

 

 

 

Table  2:  kd  and  Mlim  values  for  Pd(NHC-­‐pTHF)2Cl2,  Pt(NHC-­‐pTHF)2Cl2  and  EtIm-­‐pTHF  

Polymer   kd  (mol  x  g-­‐1  x  min-­‐1)   Mlim  (kDa)  

Pd(NHC-­‐pTHF)2Cl2   7.42×10!! ±3.63×10!!   20  

Pt(NHC-­‐pTHF)2Cl2   5.27×10!! ±2.23×10!!   22  

EtIm-­‐pTHF   3.12×10!! ±5.54×10!!   29  

 kd  is  the  slope  and  Mlim  is  the  intercept  in  the  plot  of  ksc  vs.  MW    

The  different  𝑘𝑘!  values   for   the  Pd   and  Pt   complexes   are   in   contrast   to   the  observed  MW  

dependencies   for   covalent   bond   scission   of   isomers   of   cyclobutane   mechanophores,  

reported  by  Moore  et  al  (Figure  9).36  They  observed  that  calculated  Fmax  (the  maximum  force  

imposed   on   the   model   mechanophores   for   bond   rupture   to   occur)   and   experimentally  

determined  Mlim  values  differ  significantly  for  different  mechanophores.  However,  slopes  in  

a  plot  of  ksc  vs  MW  are  identical  for  all  regardless  of  mechanophore’s  chemical  structure.  

 

Figure  9:  Structures  of  cis  and  trans  dicyano-­‐substituted  cyclobutanes  (DCC  and  DCT),  cis  and  trans  

monocyano-­‐substituted  cyclobutanes  (MCC  and  MCT),  and  cis  and  trans  cyclobutanes  without  cyano  

substituents   (NCC   and  NCT).   R   represents   a   poly(methylacrylate)   chain.   The   table   summarizes   the  

Fmax  and  Mlim  values  for  related  mechanophores.  Adapted  from  ref  36.    

Chapter 2

34

Page 42: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   8:   Change   in   initial   polymer   concentration  vs.   sonication   time   for  M(NHC-­‐pTHF18k)2Cl2   (M:  

Pd,  and  Pt).    

M(NHC-­‐pTHF)2Cl2  samples  were  sonicated  separately  and  their  scission  rate  constants  (ksc)  

were   determined  by   exponential   fitting.   The  ksc   vs  MW  plot   in   Figure   8   shows   that   chain  

scission   in   Pt   containing   polymers   is   slower   than   in   Pd   containing   counterparts   for   each  

initial  MW.  Although  the  calculated  ksc’s  are  linearly  dependent  on  MW  for  both  metals,  the  

slopes  of  the  lines  are  different.  Madras  et  al.  used  the  following  equation  to  express  ksc  as  a  

function  of  MW:35    

  𝑘𝑘!"  =  𝑘𝑘!  (𝑀𝑀𝑀𝑀   −  𝑀𝑀!"#)ƛ   (4)  

In   which  𝑘𝑘!  is   an   empirical   proportionality   constant.   In   the   Pd   and   Pt   complexes,   the  

molecular  weight  dependent  scission  rates  fit  well  to  equation  4  with  exponent  λ  equal  to  1.  

Mlim  values  for  both  mechanophores  were  estimated  from  linear  extrapolation  of  the  plots  

of   ksc   vs   MW   (Figure   10)   as   20   kDa   and   22   kDa   for   Pd(NHC-­‐pTHF)2Cl2   and   for   Pt(NHC-­‐

pTHF)2Cl2  (Table  2)  respectively.    

 

 

 

 

Table  2:  kd  and  Mlim  values  for  Pd(NHC-­‐pTHF)2Cl2,  Pt(NHC-­‐pTHF)2Cl2  and  EtIm-­‐pTHF  

Polymer   kd  (mol  x  g-­‐1  x  min-­‐1)   Mlim  (kDa)  

Pd(NHC-­‐pTHF)2Cl2   7.42×10!! ±3.63×10!!   20  

Pt(NHC-­‐pTHF)2Cl2   5.27×10!! ±2.23×10!!   22  

EtIm-­‐pTHF   3.12×10!! ±5.54×10!!   29  

 kd  is  the  slope  and  Mlim  is  the  intercept  in  the  plot  of  ksc  vs.  MW    

The  different  𝑘𝑘!  values   for   the  Pd   and  Pt   complexes   are   in   contrast   to   the  observed  MW  

dependencies   for   covalent   bond   scission   of   isomers   of   cyclobutane   mechanophores,  

reported  by  Moore  et  al  (Figure  9).36  They  observed  that  calculated  Fmax  (the  maximum  force  

imposed   on   the   model   mechanophores   for   bond   rupture   to   occur)   and   experimentally  

determined  Mlim  values  differ  significantly  for  different  mechanophores.  However,  slopes  in  

a  plot  of  ksc  vs  MW  are  identical  for  all  regardless  of  mechanophore’s  chemical  structure.  

 

Figure  9:  Structures  of  cis  and  trans  dicyano-­‐substituted  cyclobutanes  (DCC  and  DCT),  cis  and  trans  

monocyano-­‐substituted  cyclobutanes  (MCC  and  MCT),  and  cis  and  trans  cyclobutanes  without  cyano  

substituents   (NCC   and  NCT).   R   represents   a   poly(methylacrylate)   chain.   The   table   summarizes   the  

Fmax  and  Mlim  values  for  related  mechanophores.  Adapted  from  ref  36.    

Mechanochemical chain scission in NHC-Pd

35

2

Page 43: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   10:  Molecular  weight   dependence  of   scission   rates   (ksc)   for   coordination  polymers  M(NHC-­‐

pTHF)2Cl2    (M:Pd  and  Pt),  and  covalent  polymer  (EtIm-­‐pTHF)Cl.  

In   order   to   verify   the   observed   difference   in   kd   and  Mlim   between   Pd   and   Pt   containing  

polymers,  we  have  first  calculated  the  bond  dissociation  energy  (BDE)  for  M-­‐NHC  bond  for  

Pd  and  Pt  using  DFT  method.  The  PBE0  hybrid  functional  that  accounts  for  25%  of  Hartree-­‐

Fock  exchange  has  been  employed  together  with  the  TZVP  basis  set  reported  by  Ahlrichs  et.  

al.  37,38  Data  represents  the  reaction  ML2  →  LM  +  L    where  ligand  (L)  was  chosen  as  EtMetIm  

for   structural   simplicity.   BDEs   for   M-­‐ligand   coordination   bonds   were   calculated   as   195  

kJ/mol  and  245  kJ/mol  for  Pd-­‐NHC  and  Pt-­‐NHC  respectively.  Energies  mentioned  here  is  the  

energy   difference   between   coordinated   and   dissociated   states.   This   indicates   that   the  

energy   needed   to   remove   one   NHC   ligand   coordinated   to   metal   in   M(NHC-­‐pTHF)2Cl2  

complexes  is  significantly  lower  than  C-­‐C  or  C-­‐O  bond  dissociation  energies  (>350  kJ/mol).39  

Therefore,  the  coordination  bond  in  M-­‐NHC  incorporated  polymers  is  the  weakest  bond  on  

the  chain  backbone  and  breaks  selectively  during  sonication.    

Mlim   for   Pd   and   Pt   containing   M(NHC-­‐pTHF)2Cl2   complexes   are   higher   than   the   one    

previously  reported  for  Ag(NHC-­‐pTHF)2PF6  complex    (<13  kDa33).  That  is  also  consistent  with  

the  BDE  for  Ag-­‐NHC  bond,  which  was  calculated  as  141  kJ/mol  using  the  same  method  as  for  

Pd  and  Pt.    

Determining  of  BDEs  for  the  complexes  in  concern  showed  that  the  higher  BDE  corresponds  

to   a   higher   Mlim.   However,   it   does   not   explain   the   difference   in   the   molecular   weight  

dependence  of  scission  rates,  i.e.  the  difference  in  slopes  (𝑘𝑘!)  in  Figure  10.    

An  explanation  of  the  difference  in  𝑘𝑘!  values  for  coordinative  bonds  versus  covalent  bonds,  

and  between  different  coordinative  bonds  must  be  sought  in  the  relation  between  force  and  

scission  rate,  which  is  expressed  in  its  most  simple  form  as  the  Bell-­‐Evans  equation:40,41  

𝑘𝑘 𝐹𝐹 = 𝑘𝑘!𝑒𝑒!∆!‡/!!!  

 This  equation  expresses  the  fact  that  the  force  dependence  of  scission  rate  k(F)  can  be  very  

different  for  even  for  bonds  with  equal  k0  when  there  is  a  difference  between  the  values  of  

Δx‡,   the   change   in   bond   length   between   the   force   free   equilibrium  bond   length,   and   the  

transition  state  bond  length  under  force.  Given  the  differences  in  equilibrium  bond  lengths  

and   the   shape   of   the   potential   well   for   coordinative   bonds   and   covalent   (C-­‐C)   bonds,  

differences   in  the  slope  of  the  plots   in  Figure  10  are  not  surprising.  Similar  trend   in   force-­‐

rate  relationship  was  reported  previously  in  literature  as  shown  in  Figure  11.42  

 

Figure  11:  Measured  (points)  and  calculated  (lines)  rate–force  correlations  of  three  reactions.42  

In   order   to   determine   the   Fmax   dependence   of   Mlim   for   M(NHC-­‐pTHF)2Cl2   complexes   the  

pulling   force  was   implemented  by   the  COGEF  method,  where   constrained  geometries   are  

used   to   simulate   the   presence   of   a   pulling   force.43   Distance   between   carbene   carbons   of  

different  NHCs   coordinated   to   the   same  Pd   center   (C1   and  C2)  was   increased   sequentially  

and  the  energy  of  the  system  was  calculated.  BDE,  in  this  case,  was  calculated  as  the  relative  

energy  at  very  high-­‐constrained  distance  (Figure  12a).  Fmax  values  were  calculated  from  the  

derivative  of  relative  energy  vs  constrained  distance  curve  (Figure  12b).    

Chapter 2

36

Page 44: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   10:  Molecular  weight   dependence  of   scission   rates   (ksc)   for   coordination  polymers  M(NHC-­‐

pTHF)2Cl2    (M:Pd  and  Pt),  and  covalent  polymer  (EtIm-­‐pTHF)Cl.  

In   order   to   verify   the   observed   difference   in   kd   and  Mlim   between   Pd   and   Pt   containing  

polymers,  we  have  first  calculated  the  bond  dissociation  energy  (BDE)  for  M-­‐NHC  bond  for  

Pd  and  Pt  using  DFT  method.  The  PBE0  hybrid  functional  that  accounts  for  25%  of  Hartree-­‐

Fock  exchange  has  been  employed  together  with  the  TZVP  basis  set  reported  by  Ahlrichs  et.  

al.  37,38  Data  represents  the  reaction  ML2  →  LM  +  L    where  ligand  (L)  was  chosen  as  EtMetIm  

for   structural   simplicity.   BDEs   for   M-­‐ligand   coordination   bonds   were   calculated   as   195  

kJ/mol  and  245  kJ/mol  for  Pd-­‐NHC  and  Pt-­‐NHC  respectively.  Energies  mentioned  here  is  the  

energy   difference   between   coordinated   and   dissociated   states.   This   indicates   that   the  

energy   needed   to   remove   one   NHC   ligand   coordinated   to   metal   in   M(NHC-­‐pTHF)2Cl2  

complexes  is  significantly  lower  than  C-­‐C  or  C-­‐O  bond  dissociation  energies  (>350  kJ/mol).39  

Therefore,  the  coordination  bond  in  M-­‐NHC  incorporated  polymers  is  the  weakest  bond  on  

the  chain  backbone  and  breaks  selectively  during  sonication.    

Mlim   for   Pd   and   Pt   containing   M(NHC-­‐pTHF)2Cl2   complexes   are   higher   than   the   one    

previously  reported  for  Ag(NHC-­‐pTHF)2PF6  complex    (<13  kDa33).  That  is  also  consistent  with  

the  BDE  for  Ag-­‐NHC  bond,  which  was  calculated  as  141  kJ/mol  using  the  same  method  as  for  

Pd  and  Pt.    

Determining  of  BDEs  for  the  complexes  in  concern  showed  that  the  higher  BDE  corresponds  

to   a   higher   Mlim.   However,   it   does   not   explain   the   difference   in   the   molecular   weight  

dependence  of  scission  rates,  i.e.  the  difference  in  slopes  (𝑘𝑘!)  in  Figure  10.    

An  explanation  of  the  difference  in  𝑘𝑘!  values  for  coordinative  bonds  versus  covalent  bonds,  

and  between  different  coordinative  bonds  must  be  sought  in  the  relation  between  force  and  

scission  rate,  which  is  expressed  in  its  most  simple  form  as  the  Bell-­‐Evans  equation:40,41  

𝑘𝑘 𝐹𝐹 = 𝑘𝑘!𝑒𝑒!∆!‡/!!!  

 This  equation  expresses  the  fact  that  the  force  dependence  of  scission  rate  k(F)  can  be  very  

different  for  even  for  bonds  with  equal  k0  when  there  is  a  difference  between  the  values  of  

Δx‡,   the   change   in   bond   length   between   the   force   free   equilibrium  bond   length,   and   the  

transition  state  bond  length  under  force.  Given  the  differences  in  equilibrium  bond  lengths  

and   the   shape   of   the   potential   well   for   coordinative   bonds   and   covalent   (C-­‐C)   bonds,  

differences   in  the  slope  of  the  plots   in  Figure  10  are  not  surprising.  Similar  trend   in   force-­‐

rate  relationship  was  reported  previously  in  literature  as  shown  in  Figure  11.42  

 

Figure  11:  Measured  (points)  and  calculated  (lines)  rate–force  correlations  of  three  reactions.42  

In   order   to   determine   the   Fmax   dependence   of   Mlim   for   M(NHC-­‐pTHF)2Cl2   complexes   the  

pulling   force  was   implemented  by   the  COGEF  method,  where   constrained  geometries   are  

used   to   simulate   the   presence   of   a   pulling   force.43   Distance   between   carbene   carbons   of  

different  NHCs   coordinated   to   the   same  Pd   center   (C1   and  C2)  was   increased   sequentially  

and  the  energy  of  the  system  was  calculated.  BDE,  in  this  case,  was  calculated  as  the  relative  

energy  at  very  high-­‐constrained  distance  (Figure  12a).  Fmax  values  were  calculated  from  the  

derivative  of  relative  energy  vs  constrained  distance  curve  (Figure  12b).    

Mechanochemical chain scission in NHC-Pd

37

2

Page 45: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Table  3:  BDE,  Mlim,  Fmax  for  coordination  polymers  with  Ag,  Pd,  Pt  and  covalent  polymer  pTHF.    

Polymer   Bond   ∆E  (kJ/mol)   Mlim  (kDa)   Fmax  (nN)  

Ag(NHC-­‐pTHF)2PF6   Ag-­‐Ca,b   141   >13   0.5  

Pd(NHC-­‐pTHF)2Cl2   Pd-­‐Ca   195   20   3.0  

Pt(NHC-­‐pTHF)2Cl2   Pt-­‐Ca   245   22   4.2  

pTHF  C-­‐Cc  

C-­‐Oc  

370  

344  29d  

6.9  

7.6  a  C  represents  the  carbene  carbon  of  NHC  that  is  coordinated  to  the  metal  center  bref  25  cref  41  dthis  work  

As  summarized  in  Table  3,  application  of  an  external  force  of  2.9  nN  and  4.2  nN  would  be  

sufficient  to  achieve  scission  of  the  Pd–C  and  Pt-­‐C  bonds  in  M-­‐NHC  complexes  respectively  

at  room  temperature.  Since  these  values  are  significantly   lower  than  the  force  required  to  

break  bonds  on  pTHF  chain,43  M-­‐ligand  bonds  are  the  most  susceptible  bonds  on  polymer  

backbone  to  mechanical  rupture.  Mlim  values,  which  were  found  experimentally  for  Pd(NHC-­‐

pTHF)2Cl2,  Pt(NHC-­‐pTHF)2Cl2  and  pTHF  are   in  good  agreement  with  the  trend   in  calculated  

BDE  and  Fmax  values.    

 

 

Figure   12:   Chemical   structure   of   Pd(NHC)2Cl2  that  was   used   in   calculations   and   change   in   relative  

energy   (a)   and   force   (b)   as   a   function  of  distance  between  NHC-­‐Pd.  Arrows  on   chemical   structure  

indicate  the  pulling  direction.  

Chain  scission  mechanism  

In  the  previous  section  COGEF  calculations  were  done  with  the  constraint  C1-­‐Pd-­‐C2  distance  

as   constraint.   In   order   to   understand   the   chain   scission   mechanism,   however,   a   second  

series   of   simulations   was   carried   out.   In   mechanochemical   experiments,   tensile   force   is  

transduced  through  the  polymer  chain  to  the  central  mechanophore,  and  the  applied  force  

is   orthogonal   to   C-­‐Pd   bond   (Figure   14c).   Therefore,   in   the   calculations,   pulling   was   also  

simulated  as  orthogonal   to  NHC-­‐Pd  coordination  bond.  Ethyl  groups  were  used   instead  of  

long   pTHF   chains   to   reduce   computational   effort.   The   distance   between   two   terminal  

carbon   atoms   on   different   NHCs   coordinated   to   the   same   Pd   center   was   increased  

sequentially  from  8  to  12 Å.  (Figure  13).    

Chapter 2

38

Page 46: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Table  3:  BDE,  Mlim,  Fmax  for  coordination  polymers  with  Ag,  Pd,  Pt  and  covalent  polymer  pTHF.    

Polymer   Bond   ∆E  (kJ/mol)   Mlim  (kDa)   Fmax  (nN)  

Ag(NHC-­‐pTHF)2PF6   Ag-­‐Ca,b   141   >13   0.5  

Pd(NHC-­‐pTHF)2Cl2   Pd-­‐Ca   195   20   3.0  

Pt(NHC-­‐pTHF)2Cl2   Pt-­‐Ca   245   22   4.2  

pTHF  C-­‐Cc  

C-­‐Oc  

370  

344  29d  

6.9  

7.6  a  C  represents  the  carbene  carbon  of  NHC  that  is  coordinated  to  the  metal  center  bref  25  cref  41  dthis  work  

As  summarized  in  Table  3,  application  of  an  external  force  of  2.9  nN  and  4.2  nN  would  be  

sufficient  to  achieve  scission  of  the  Pd–C  and  Pt-­‐C  bonds  in  M-­‐NHC  complexes  respectively  

at  room  temperature.  Since  these  values  are  significantly   lower  than  the  force  required  to  

break  bonds  on  pTHF  chain,43  M-­‐ligand  bonds  are  the  most  susceptible  bonds  on  polymer  

backbone  to  mechanical  rupture.  Mlim  values,  which  were  found  experimentally  for  Pd(NHC-­‐

pTHF)2Cl2,  Pt(NHC-­‐pTHF)2Cl2  and  pTHF  are   in  good  agreement  with  the  trend   in  calculated  

BDE  and  Fmax  values.    

 

 

Figure   12:   Chemical   structure   of   Pd(NHC)2Cl2  that  was   used   in   calculations   and   change   in   relative  

energy   (a)   and   force   (b)   as   a   function  of  distance  between  NHC-­‐Pd.  Arrows  on   chemical   structure  

indicate  the  pulling  direction.  

Chain  scission  mechanism  

In  the  previous  section  COGEF  calculations  were  done  with  the  constraint  C1-­‐Pd-­‐C2  distance  

as   constraint.   In   order   to   understand   the   chain   scission   mechanism,   however,   a   second  

series   of   simulations   was   carried   out.   In   mechanochemical   experiments,   tensile   force   is  

transduced  through  the  polymer  chain  to  the  central  mechanophore,  and  the  applied  force  

is   orthogonal   to   C-­‐Pd   bond   (Figure   14c).   Therefore,   in   the   calculations,   pulling   was   also  

simulated  as  orthogonal   to  NHC-­‐Pd  coordination  bond.  Ethyl  groups  were  used   instead  of  

long   pTHF   chains   to   reduce   computational   effort.   The   distance   between   two   terminal  

carbon   atoms   on   different   NHCs   coordinated   to   the   same   Pd   center   was   increased  

sequentially  from  8  to  12 Å.  (Figure  13).    

Mechanochemical chain scission in NHC-Pd

39

2

Page 47: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   13:   Energy   optimized   structures   for   Pd(NHC)2Cl2   while   increasing   the   constrained   distance  

between  the  atoms  indicated  in  black.    

On   the  calculated  curve  of   the  energy  vs.  distance   (Figure  14)   three  characteristic   regions  

can  be  seen.   i)  Small-­‐constrained  distance  between  the  pulling  ends.  There   is  only  a  small  

increase   of   the   energy   due   to   the   relatively   small   torsional   angle   constraints.   Here,   the  

distance  between  the  central  Pd  and  the  carbon  atoms  does  not  increase.  

ii)  The  strain   in   the  system   is   sufficient   to   stretch   the  Pd-­‐C  bonds.  This   is  observed  as   the  

steep   increase   of   the   energy   until   the   cleavage   point.   At   this   stage,   the   NHC   and   ethyl  

groups   lay   in   the   same   plane,   and   all   strain   imposed   on   the   system   is   converted   to  

stretching  of  the  Pd-­‐C  bond.  The  complex   is  still  symmetric  at  this  point  and  the  distances  

C1-­‐Pd  and  C2-­‐Pd  are  equal.  

iii)   At   a   constrained   distance   between   terminal   atoms   of   the   complex   (11.6   Å),   the  

symmetry  of  the  system  is  broken  and  Pd  settles  much  closer  to  the  C2  atom.    

 

Figure  14:  a)  Change  in  relative  energy  and  b)  distance  between  NHC-­‐Pd  as  a  function  of  constrained  

distance.   c)   Chemical   structure  of   Pd(NHC)2Cl2  that  was  used   in   calculations.   d)   Idealized  potential  

energy  surface  (PES)  of  the  investigated  system  along  the  Pd-­‐C1  and  Pd-­‐C2  bond  length  coordinates.  

The  geometry  of  the  system  has  been  frozen  during  the  simulations,  and  the  eventual  distortion  of  

the  imidazole  rings   is  not   included  in  the  results.  Arrows  on  chemical  structure  indicate  the  pulling  

direction.  

The  C1-­‐Pd-­‐C2  bridge  stretching  proceeds  via  the  path  that  is  shown  in  Figure  14d.  Both  C-­‐Pd  

bonds  stretch  at  the  same  rate  and  the  energy  rises  along  the  steepest,  diagonal  path.  Once  

the  energy  rises  above  a  threshold  value,  it  is  more  energetically  favorable  to  split  along  one  

of  the  two  directions  to  give  C1-­‐Pd  or  C2-­‐Pd  bond  cleavage.    

Comparison  with  calculations  for  silver  carbene  complex  shows  that  scission  of  the  Pd-­‐NHC  

has   a  higher   activation  barrier   than   scission  of  Ag-­‐NHC  bond.   This   explains   the   significant  

difference   in  Mlims   to   break   Ag-­‐NHC   and   Pd-­‐NHC   bonds.   Additionally,   when   the   distance  

between   two   terminal   atoms   increases,   the   most   labile   bond   (Pd-­‐C   coordination   bond)  

breaks  eventually  on  the  chain.  

Chapter 2

40

Page 48: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   13:   Energy   optimized   structures   for   Pd(NHC)2Cl2   while   increasing   the   constrained   distance  

between  the  atoms  indicated  in  black.    

On   the  calculated  curve  of   the  energy  vs.  distance   (Figure  14)   three  characteristic   regions  

can  be  seen.   i)  Small-­‐constrained  distance  between  the  pulling  ends.  There   is  only  a  small  

increase   of   the   energy   due   to   the   relatively   small   torsional   angle   constraints.   Here,   the  

distance  between  the  central  Pd  and  the  carbon  atoms  does  not  increase.  

ii)  The  strain   in   the  system   is   sufficient   to   stretch   the  Pd-­‐C  bonds.  This   is  observed  as   the  

steep   increase   of   the   energy   until   the   cleavage   point.   At   this   stage,   the   NHC   and   ethyl  

groups   lay   in   the   same   plane,   and   all   strain   imposed   on   the   system   is   converted   to  

stretching  of  the  Pd-­‐C  bond.  The  complex   is  still  symmetric  at  this  point  and  the  distances  

C1-­‐Pd  and  C2-­‐Pd  are  equal.  

iii)   At   a   constrained   distance   between   terminal   atoms   of   the   complex   (11.6   Å),   the  

symmetry  of  the  system  is  broken  and  Pd  settles  much  closer  to  the  C2  atom.    

 

Figure  14:  a)  Change  in  relative  energy  and  b)  distance  between  NHC-­‐Pd  as  a  function  of  constrained  

distance.   c)   Chemical   structure  of   Pd(NHC)2Cl2  that  was  used   in   calculations.   d)   Idealized  potential  

energy  surface  (PES)  of  the  investigated  system  along  the  Pd-­‐C1  and  Pd-­‐C2  bond  length  coordinates.  

The  geometry  of  the  system  has  been  frozen  during  the  simulations,  and  the  eventual  distortion  of  

the  imidazole  rings   is  not   included  in  the  results.  Arrows  on  chemical  structure  indicate  the  pulling  

direction.  

The  C1-­‐Pd-­‐C2  bridge  stretching  proceeds  via  the  path  that  is  shown  in  Figure  14d.  Both  C-­‐Pd  

bonds  stretch  at  the  same  rate  and  the  energy  rises  along  the  steepest,  diagonal  path.  Once  

the  energy  rises  above  a  threshold  value,  it  is  more  energetically  favorable  to  split  along  one  

of  the  two  directions  to  give  C1-­‐Pd  or  C2-­‐Pd  bond  cleavage.    

Comparison  with  calculations  for  silver  carbene  complex  shows  that  scission  of  the  Pd-­‐NHC  

has   a  higher   activation  barrier   than   scission  of  Ag-­‐NHC  bond.   This   explains   the   significant  

difference   in  Mlims   to   break   Ag-­‐NHC   and   Pd-­‐NHC   bonds.   Additionally,   when   the   distance  

between   two   terminal   atoms   increases,   the   most   labile   bond   (Pd-­‐C   coordination   bond)  

breaks  eventually  on  the  chain.  

Mechanochemical chain scission in NHC-Pd

41

2

Page 49: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Fmax  was  calculated  for  pulling  on  the  terminal  carbon  atoms  of  the  ethyl  groups  as  2.2  nN  

that  is  significantly  lower  than  that  for  pulling  directly  from  carbene  carbons  (Figure  15).  The  

lower  number  indicates  that  although  the  trend  in  Table  3  is  realistic,  however,  Fmax  values  

obtained  from  pulling  on  C1  and  C2  are  upper  limits  to  break  NHC-­‐Pd  bond  mechanically  and  

overestimate  the  actual  value  of  the  force  needed  to  break  this  bond.    

 

Figure  15:  Force  as  a  function  of  distance  between  terminal  atoms  on  the  ethyl  groups.    Arrows  on  

chemical  structure  indicate  the  pulling  direction.  M:Pd.  

 

 

 

 

 

 

 

 

 

 

Conclusions    

Ultrasound   induced   chain   scission   in   Pd(NHC-­‐pTHF)2Cl2   complexes   were   investigated   in  

toluene.  Reversibility  of  chain  scission  depends  on  the  nature  of  gas  molecules  dissolved  in  

toluene  and  scission  is  reversible  when  the  polymer  complex  is  sonicated  under  CH4.  Mlim  for  

Pd(NHC-­‐pTHF)2Cl2  was   calculated   from   the   graph   of   ksc   vs  Mw  graph   as   approximately   20  

kDa.  Critical  molecular  weight  for  the  complex  (Mc)  determined  as  60  kDa  that   is  twice  as  

high  as  Mlim  for  EtIm-­‐pTHF.  Above  Mc  ultrasound  induced  degradation  of  ligands  due  to  C-­‐C  

bond  scission  would  prevent  reversibility.  16  kDa  complex  Pd(NHC-­‐pTHF18k)2Cl2  and  small  

molecule  model   complex  Pd(NHC-­‐ButEtIm)2Cl2  were  not   affected  by   sonication   confirming  

the  stability  of  complexes  when  the  force  is  lower  than  the  threshold  value  to  break  Pd-­‐NHC  

bond.   Scission   rate   is   dependent   on   initial   molecular   weight   of   the   polymer   and   this  

indicates   that   chain   scission   has   a   true   mechanical   nature.   Constrained   geometry  

optimization  (COGEF)  calculations  showed  that  distance  between  NHC  and  Pd  increases  for  

both  NHCs  coordinated  to  metal  center  and  one  of  them  breaks  upon  stepwise  increase  in  

constrained   distance.   The   results   of   the   calculations   also   showed   the   importance   of  

choosing  attachment  points  when  determining  the  value  of  Fmax.  This   is  also   in  agreement  

with  the  findings  on  attachment  point-­‐reactivity  relation  reported  by  Craig  et  al.44–46    

Results   confirmed   that   ligands   can   be   released   from   thermally   stable   Pd-­‐NHC   complexes  

mechanically  to  produce  free  NHC  and  coordinatively  unsaturated  metal  center,  and  further  

use  of  these  reactive  mechano-­‐chemical  products  is  of  interest.    

 

 

 

 

 

 

 

 

 

Chapter 2

42

Page 50: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Fmax  was  calculated  for  pulling  on  the  terminal  carbon  atoms  of  the  ethyl  groups  as  2.2  nN  

that  is  significantly  lower  than  that  for  pulling  directly  from  carbene  carbons  (Figure  15).  The  

lower  number  indicates  that  although  the  trend  in  Table  3  is  realistic,  however,  Fmax  values  

obtained  from  pulling  on  C1  and  C2  are  upper  limits  to  break  NHC-­‐Pd  bond  mechanically  and  

overestimate  the  actual  value  of  the  force  needed  to  break  this  bond.    

 

Figure  15:  Force  as  a  function  of  distance  between  terminal  atoms  on  the  ethyl  groups.    Arrows  on  

chemical  structure  indicate  the  pulling  direction.  M:Pd.  

 

 

 

 

 

 

 

 

 

 

Conclusions    

Ultrasound   induced   chain   scission   in   Pd(NHC-­‐pTHF)2Cl2   complexes   were   investigated   in  

toluene.  Reversibility  of  chain  scission  depends  on  the  nature  of  gas  molecules  dissolved  in  

toluene  and  scission  is  reversible  when  the  polymer  complex  is  sonicated  under  CH4.  Mlim  for  

Pd(NHC-­‐pTHF)2Cl2  was   calculated   from   the   graph   of   ksc   vs  Mw  graph   as   approximately   20  

kDa.  Critical  molecular  weight  for  the  complex  (Mc)  determined  as  60  kDa  that   is  twice  as  

high  as  Mlim  for  EtIm-­‐pTHF.  Above  Mc  ultrasound  induced  degradation  of  ligands  due  to  C-­‐C  

bond  scission  would  prevent  reversibility.  16  kDa  complex  Pd(NHC-­‐pTHF18k)2Cl2  and  small  

molecule  model   complex  Pd(NHC-­‐ButEtIm)2Cl2  were  not   affected  by   sonication   confirming  

the  stability  of  complexes  when  the  force  is  lower  than  the  threshold  value  to  break  Pd-­‐NHC  

bond.   Scission   rate   is   dependent   on   initial   molecular   weight   of   the   polymer   and   this  

indicates   that   chain   scission   has   a   true   mechanical   nature.   Constrained   geometry  

optimization  (COGEF)  calculations  showed  that  distance  between  NHC  and  Pd  increases  for  

both  NHCs  coordinated  to  metal  center  and  one  of  them  breaks  upon  stepwise  increase  in  

constrained   distance.   The   results   of   the   calculations   also   showed   the   importance   of  

choosing  attachment  points  when  determining  the  value  of  Fmax.  This   is  also   in  agreement  

with  the  findings  on  attachment  point-­‐reactivity  relation  reported  by  Craig  et  al.44–46    

Results   confirmed   that   ligands   can   be   released   from   thermally   stable   Pd-­‐NHC   complexes  

mechanically  to  produce  free  NHC  and  coordinatively  unsaturated  metal  center,  and  further  

use  of  these  reactive  mechano-­‐chemical  products  is  of  interest.    

 

 

 

 

 

 

 

 

 

Mechanochemical chain scission in NHC-Pd

43

2

Page 51: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Experimental  

General  

All   chemicals   were   purchased   from   commercial   sources   used   without   further   purification,   unless  

specified  otherwise.  Dry  tetrahydrofuran  (THF,  HPLC  grade)  was  degassed  with  argon  and  purified  by  

passage  through  activated  alumina  solvent  column  prior  to  use.  A  Varian  400MR  or  a  Varian  Mercury  

400  spectrometer  was  used  to  record  1H  NMR  (400  MHz)  Chemical  shifts  are  reported  in  ppm  and  

referenced   to   tetramethylsilane  or   solvent.  Gel  permeation  chromatography   (GPC)  was  performed  

on   a   Shimadzu   LC10-­‐AD,   using   Polymer   Laboratories   PL   Gel   5μm  MIXED-­‐C   and  MIXEDD   columns    

(linear   range  of  MW:  200–2000000  g/mol),   a  Shimadzu  SPD-­‐M10A  UV-­‐vis  detector  at  254  nm  and  

RID-­‐10A  refractive  index  detector,  and  THF  as  eluent  at  a  flow  rate  of  1  mL/min  (20  °C).  Polystyrene  

standards  were  used  for  calibration.      

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   reactor   with   a   volume   of   10  mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   The   temperature   in   the   reactor   was  maintained  with   a   Lauda   E300   cooling  

bath  and  measured  using  a  0.5  mm  diameter  thermocouple.  Solutions  were  sonicated  continuously,  

temperature   of   the   solution  was   checked   by   thermocouple   and   it  was   constant   at   20oC   after   the  

thermal  equilibrium  was  achieved  in  the  first  3-­‐5  mins.  During  sonication  saturation  a  gas  (Ar  or  CH4)  

was   bubbled   through   solution   via   teflon   tubing.   Aliquots   of   100   µL   were   taken   at   different   time  

intervals.   Toluene  was   removed   under   reduced   pressure   and   residues  were   dissolved   in   THF   and  

submitted   to   GPC.     GPC   results   were   analyzed   and   scission   kinetics   was   determined   by   double  

Gaussian  de-­‐convolution  method.    

 Synthesis  of  α-­‐(N-­‐ethylimidazolium)-­‐ω-­‐methoxy  poly(tetrahydrofuran)    

Polymer   salts   precursor   to   ligand   NHC-­‐pTHF   were   synthesized   via   cationic   ring-­‐opening  

polymerization  of  tetrahydrofuran  (THF).47  THF  (100  mL)  and  DTBP  (200  µL,  0.92  mmol)  were  added  

methyl   triflate   (100   µL,   0.91  mmol)   inside   a     Schlenk   round-­‐bottom   flask   under   Ar   to   initiate   the  

polymerization.  After  stirring  for  defined  time  (Figure  S1),  the  polymerization  was  terminated  by  N-­‐

ethylimidazole   (200   µL,   ca.   2.1   mmol).   After   20  mins,   solution   was   diluted   to   app   ¼   of   its   initial  

volume  and  precipitated   in  water  (400  mL)  overnight  at  ambient  temperature.  White  polymer  was  

washed  with  water,  dissolved  in  diethyl  ether  (200  mL),  dried  over  MgSO4  and  precipitated  overnight  

at  –30  °C,  white  powder  was  filtered  washed  with  cold  Et2O  and  yielded  ligands  as  white  powder.  Ion  

exchange  of   the  anion   to   chloride  was   carried  out  by   stirring   the  polymer  with  Dowex®  exchange  

resin   in  methanol   for  2–3  hours.  Then,   the   resin  was   removed  by   filtration  and   the  methanol  was  

evaporated   in   vacuo   and   the   residue  was   precipitated   in   Et2O   at   -­‐30oC   again.   In   order   to   remove  

traces  of  solvents,  ligands  were  left  under  vacuum  at  ambient  temperature  overnight  prior  to  use.  1H  

NMR  [EtIm-­‐pTHF12kCl  CD2Cl2,  400  MHz]:  11  ppm  (s,  NHC),  7.2  ppm  (d,  NHC),  4.4  ppm  (t,  N-­‐CH2)  3.0-­‐

3.6  ppm  (br  O-­‐CH2-­‐),  1.3-­‐2.2  ppm  (br,  OCH2-­‐CH2).  

Synthesis  of  Pd(II)–NHC  polymer  complexes  Pd(NHC-­‐pTHF)2Cl2.  

EtIm-­‐pTHF  (400  mg)  was  dissolved  in  THF  (10  ml)  and  stirred  over  4A  molecular  sieves  for  30  mins  

under  Ar.  NatOBu   (2.5   eq.)  was   added   in   one  portion   and   solution   stirred   for   another   30  Å  mins.  

Then,   Pd(PhCN)2Cl2   (0.5   eq.)   was   added   and   ligand   exchange   yielded   desired   polymer   attached  

mechanophores.  Crude  products  were  filtered  over  alumina  (eluent:  THF),  and  concentrated  under  

reduced   pressure.   Then,   the   complex   was   dissolved   in   DCM   washed   with   water   and   dried   over  

MgSO4.   Solution   containing   Pd(NHC-­‐pTHF)2Cl2   in   DCM   passed   through   Celite.   Solvent   was  

evaporated  under   reduced  pressure  and   light  yellow  polymer  was   left  under  vacuum  overnight   to  

remove  all  residual  solvents.  1H  NMR  [Pd(NHC-­‐pTHF12k)2Cl2  CD2Cl2,  400  MHz]:  6.9  ppm  (s,  NHC),  4.5  

ppm  (t,  N-­‐CH2),    3.0-­‐3.5  ppm  (br,  O-­‐CH2-­‐),  1.1-­‐1.9  ppm  (br,  OCH2-­‐CH2-­‐).  

Synthesis  of  model  complex  Pd(NHC-­‐EtBut)2Cl2  

Ethyl  imidazole  (480  mg,  5  mmol)  and  iodobutane  (1840  mg,  10  mmol)  were  refluxed  in  dry  THF  (25  

ml)  overnight.  Then,  two-­‐phase  liquid  reaction  mixture  was  cooled  down  to  room  temperature  and  

upper   transparent  part  was  removed  with  glass  pipette.  Yellow  oily  product  was  washed  with  THF  

(10   ml   x   3)   and   then   residue   was   dissolved   in   MeOH   (10   ml).   Ion   exchange   resin   was   added   to  

exchange  iodine  counterions  with  chlorine.  After  stirring  2h  at  RT  mixture  was  filtered  over  a  filter  

paper,  solvent  was  evaporated  under  reduced  pressure  and   light  yellow  oil  was   left  under  vacuum  

overnight.    

EtButIm   (100  mg,   0.53  mmol),   PdCl2   (46  mg,   0.26  mmol)   and   Cs2CO3   (814.5  mg,   2.5  mmol)  were  

dissolved   in  Dioxane   (4.4  mL)  and   the   reaction  mixture  was  heated   to  80   °C.  After  5  hours,   it  was  

allowed   to   cool   down   to   room   temperature.   Volatiles   were   removed   under   reduced   pressure.  

Purification  by  column  chromatography  (CH2Cl2)  afforded  Pd(NHC-­‐EtBut)2Cl2  as  a  light  yellow  solid.  

Chapter 2

44

Page 52: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Experimental  

General  

All   chemicals   were   purchased   from   commercial   sources   used   without   further   purification,   unless  

specified  otherwise.  Dry  tetrahydrofuran  (THF,  HPLC  grade)  was  degassed  with  argon  and  purified  by  

passage  through  activated  alumina  solvent  column  prior  to  use.  A  Varian  400MR  or  a  Varian  Mercury  

400  spectrometer  was  used  to  record  1H  NMR  (400  MHz)  Chemical  shifts  are  reported  in  ppm  and  

referenced   to   tetramethylsilane  or   solvent.  Gel  permeation  chromatography   (GPC)  was  performed  

on   a   Shimadzu   LC10-­‐AD,   using   Polymer   Laboratories   PL   Gel   5μm  MIXED-­‐C   and  MIXEDD   columns    

(linear   range  of  MW:  200–2000000  g/mol),   a  Shimadzu  SPD-­‐M10A  UV-­‐vis  detector  at  254  nm  and  

RID-­‐10A  refractive  index  detector,  and  THF  as  eluent  at  a  flow  rate  of  1  mL/min  (20  °C).  Polystyrene  

standards  were  used  for  calibration.      

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   reactor   with   a   volume   of   10  mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   The   temperature   in   the   reactor   was  maintained  with   a   Lauda   E300   cooling  

bath  and  measured  using  a  0.5  mm  diameter  thermocouple.  Solutions  were  sonicated  continuously,  

temperature   of   the   solution  was   checked   by   thermocouple   and   it  was   constant   at   20oC   after   the  

thermal  equilibrium  was  achieved  in  the  first  3-­‐5  mins.  During  sonication  saturation  a  gas  (Ar  or  CH4)  

was   bubbled   through   solution   via   teflon   tubing.   Aliquots   of   100   µL   were   taken   at   different   time  

intervals.   Toluene  was   removed   under   reduced   pressure   and   residues  were   dissolved   in   THF   and  

submitted   to   GPC.     GPC   results   were   analyzed   and   scission   kinetics   was   determined   by   double  

Gaussian  de-­‐convolution  method.    

 Synthesis  of  α-­‐(N-­‐ethylimidazolium)-­‐ω-­‐methoxy  poly(tetrahydrofuran)    

Polymer   salts   precursor   to   ligand   NHC-­‐pTHF   were   synthesized   via   cationic   ring-­‐opening  

polymerization  of  tetrahydrofuran  (THF).47  THF  (100  mL)  and  DTBP  (200  µL,  0.92  mmol)  were  added  

methyl   triflate   (100   µL,   0.91  mmol)   inside   a     Schlenk   round-­‐bottom   flask   under   Ar   to   initiate   the  

polymerization.  After  stirring  for  defined  time  (Figure  S1),  the  polymerization  was  terminated  by  N-­‐

ethylimidazole   (200   µL,   ca.   2.1   mmol).   After   20  mins,   solution   was   diluted   to   app   ¼   of   its   initial  

volume  and  precipitated   in  water  (400  mL)  overnight  at  ambient  temperature.  White  polymer  was  

washed  with  water,  dissolved  in  diethyl  ether  (200  mL),  dried  over  MgSO4  and  precipitated  overnight  

at  –30  °C,  white  powder  was  filtered  washed  with  cold  Et2O  and  yielded  ligands  as  white  powder.  Ion  

exchange  of   the  anion   to   chloride  was   carried  out  by   stirring   the  polymer  with  Dowex®  exchange  

resin   in  methanol   for  2–3  hours.  Then,   the   resin  was   removed  by   filtration  and   the  methanol  was  

evaporated   in   vacuo   and   the   residue  was   precipitated   in   Et2O   at   -­‐30oC   again.   In   order   to   remove  

traces  of  solvents,  ligands  were  left  under  vacuum  at  ambient  temperature  overnight  prior  to  use.  1H  

NMR  [EtIm-­‐pTHF12kCl  CD2Cl2,  400  MHz]:  11  ppm  (s,  NHC),  7.2  ppm  (d,  NHC),  4.4  ppm  (t,  N-­‐CH2)  3.0-­‐

3.6  ppm  (br  O-­‐CH2-­‐),  1.3-­‐2.2  ppm  (br,  OCH2-­‐CH2).  

Synthesis  of  Pd(II)–NHC  polymer  complexes  Pd(NHC-­‐pTHF)2Cl2.  

EtIm-­‐pTHF  (400  mg)  was  dissolved  in  THF  (10  ml)  and  stirred  over  4A  molecular  sieves  for  30  mins  

under  Ar.  NatOBu   (2.5   eq.)  was   added   in   one  portion   and   solution   stirred   for   another   30  Å  mins.  

Then,   Pd(PhCN)2Cl2   (0.5   eq.)   was   added   and   ligand   exchange   yielded   desired   polymer   attached  

mechanophores.  Crude  products  were  filtered  over  alumina  (eluent:  THF),  and  concentrated  under  

reduced   pressure.   Then,   the   complex   was   dissolved   in   DCM   washed   with   water   and   dried   over  

MgSO4.   Solution   containing   Pd(NHC-­‐pTHF)2Cl2   in   DCM   passed   through   Celite.   Solvent   was  

evaporated  under   reduced  pressure  and   light  yellow  polymer  was   left  under  vacuum  overnight   to  

remove  all  residual  solvents.  1H  NMR  [Pd(NHC-­‐pTHF12k)2Cl2  CD2Cl2,  400  MHz]:  6.9  ppm  (s,  NHC),  4.5  

ppm  (t,  N-­‐CH2),    3.0-­‐3.5  ppm  (br,  O-­‐CH2-­‐),  1.1-­‐1.9  ppm  (br,  OCH2-­‐CH2-­‐).  

Synthesis  of  model  complex  Pd(NHC-­‐EtBut)2Cl2  

Ethyl  imidazole  (480  mg,  5  mmol)  and  iodobutane  (1840  mg,  10  mmol)  were  refluxed  in  dry  THF  (25  

ml)  overnight.  Then,  two-­‐phase  liquid  reaction  mixture  was  cooled  down  to  room  temperature  and  

upper   transparent  part  was  removed  with  glass  pipette.  Yellow  oily  product  was  washed  with  THF  

(10   ml   x   3)   and   then   residue   was   dissolved   in   MeOH   (10   ml).   Ion   exchange   resin   was   added   to  

exchange  iodine  counterions  with  chlorine.  After  stirring  2h  at  RT  mixture  was  filtered  over  a  filter  

paper,  solvent  was  evaporated  under  reduced  pressure  and   light  yellow  oil  was   left  under  vacuum  

overnight.    

EtButIm   (100  mg,   0.53  mmol),   PdCl2   (46  mg,   0.26  mmol)   and   Cs2CO3   (814.5  mg,   2.5  mmol)  were  

dissolved   in  Dioxane   (4.4  mL)  and   the   reaction  mixture  was  heated   to  80   °C.  After  5  hours,   it  was  

allowed   to   cool   down   to   room   temperature.   Volatiles   were   removed   under   reduced   pressure.  

Purification  by  column  chromatography  (CH2Cl2)  afforded  Pd(NHC-­‐EtBut)2Cl2  as  a  light  yellow  solid.  

Mechanochemical chain scission in NHC-Pd

45

2

Page 53: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

1H  NMR  spectra  for  EtButIm  and  Pd(NHC-­‐EtBut)2Cl2.  

References  

(1)     Kryger,  M.  J.;  Ong,  M.  T.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Martinez,  T.  J.;  Moore,  J.  S.  J.  

Am.  Chem.  Soc.  2010,  132  (13),  4558–4559.  

(2)     Klukovich,  H.  M.;  Kean,  Z.  S.;  Iacono,  S.  T.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2011,  133  (44),  17882–

17888.  

(3)     Brantley,  J.  N.;  Wiggins,  K.  M.;  Bielawski,  C.  W.  Science  2011,  333  (6049),  1606–1609.  

(4)     Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat  Chem  2010,  2  (6),  436–437.  

(5)     Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  Organometallics  2012,  31  (6),  2476–2481.  

(6)     Powell,  A.  B.;  Bielawski,  C.  W.;  Cowley,  A.  H.  J.  Am.  Chem.  Soc.  2010,  132  (29),  10184–10194.  

(7)     Caruso,  M.  M.;  Davis,  D.  A.;  Shen,  Q.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Moore,  J.  S.  

Chem.  Rev.  2009,  109  (11),  5755–5798.  

(8)     Kuijpers,  M.  W.  A.;  Iedema,  P.  D.;  Kemmere,  M.  F.;  Keurentjes,  J.  T.  F.  Polymer  2004,  45  (19),  

6461–6467.  

(9)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

(10)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.  Int.  Ed.  2004,  43  (34),  4460–4462.  

(11)     Paulusse,  J.  M.  J.;  Huijbers,  J.  P.  J.;  Sijbesma,  R.  P.  Macromolecules  2005,  38  (15),  6290–6298.  

(12)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(13)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2008,  No.  37,  4416–4418.  

(14)     Johansson  Seechurn,  C.  C.  C.;  Kitching,  M.  O.;  Colacot,  T.  J.;  Snieckus,  V.  Angew.  Chem.  Int.  Ed.  

2012,  51  (21),  5062–5085.  

(15)     Hopkinson,  M.  N.;  Richter,  C.;  Schedler,  M.;  Glorius,  F.  Nature  2014,  510  (7506),  485–496.  

(16)     Valente,  C.;  Çalimsiz,  S.;  Hoi,  K.  H.;  Mallik,  D.;  Sayah,  M.;  Organ,  M.  G.  Angew.  Chem.  Int.  Ed.  

2012,  51  (14),  3314–3332.  

(17)     Herrmann,  W.  A.  Angew.  Chem.  Int.  Ed.  2002,  41  (8),  1290–1309.  

(18)     O’Brien,  C.  J.;  Kantchev,  E.  A.  B.;  Valente,  C.;  Hadei,  N.;  Chass,  G.  A.;  Lough,  A.;  Hopkinson,  A.  

C.;  Organ,  M.  G.  Chem.  –  Eur.  J.  2006,  12  (18),  4743–4748.  

(19)     Kantchev,  E.  A.  B.;  O’Brien,  C.  J.;  Organ,  M.  G.  Angew.  Chem.  Int.  Ed.  2007,  46  (16),  2768–

2813.  

(20)     Organ,  M.  G.;  Avola,  S.;  Dubovyk,  I.;  Hadei,  N.;  Kantchev,  E.  A.  B.;  O’Brien,  C.  J.;  Valente,  C.  

Chem.  –  Eur.  J.  2006,  12  (18),  4749–4755.  

(21)     O’Brien,  C.  J.;  Kantchev,  E.  A.  B.;  Chass,  G.  A.;  Hadei,  N.;  Hopkinson,  A.  C.;  Organ,  M.  G.;  

Setiadi,  D.  H.;  Tang,  T.-­‐H.;  Fang,  D.-­‐C.  Tetrahedron  2005,  61  (41),  9723–9735.  

(22)     Christmann,  U.;  Vilar,  R.  Angew.  Chem.  Int.  Ed.  2005,  44  (3),  366–374.  

(23)     Phan,  N.  T.  S.;  Van  Der  Sluys,  M.;  Jones,  C.  W.  Adv.  Synth.  Catal.  2006,  348  (6),  609–679.  

(24)     Marion,  N.;  Nolan,  S.  P.  Acc.  Chem.  Res.  2008,  41  (11),  1440–1449.  

(25)     Karthikeyan,  S.;  Potisek,  S.  L.;  Piermattei,  A.;  Sijbesma,  R.  P.  J.  Am.  Chem.  Soc.  2008,  130  (45),  

14968–14969.  

(26)     Szwarc,  M.  J.  Chem.  Phys.  1948,  16  (2),  128–136.  

(27)     Huynh,  H.  V.;  Han,  Y.;  Jothibasu,  R.;  Yang,  J.  A.  Organometallics  2009,  28  (18),  5395–5404.  

(28)     Liu,  S.-­‐T.;  Lee,  C.-­‐I.;  Fu,  C.-­‐F.;  Chen,  C.-­‐H.;  Liu,  Y.-­‐H.;  Elsevier,  C.  J.;  Peng,  S.-­‐M.;  Chen,  J.-­‐T.  

Organometallics  2009,  28  (24),  6957–6962.  

(29)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(30)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  Chem.  –  Eur.  J.  2009,  15  (48),  13331–13335.  

(31)     Wiggins,  K.  M.;  Brantley,  J.  N.;  Bielawski,  C.  W.  Chem.  Soc.  Rev.  2013,  42  (17),  7130–7147.  

(32)     Groote,  R.;  Szyja,  B.  M.;  Pidko,  E.  A.;  Hensen,  E.  J.  M.;  Sijbesma,  R.  P.  Macromolecules  2011,  

44  (23),  9187–9195.  

(33)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(34)     Madras,  G.;  Chung,  G.  Y.;  Smith,  J.  M.;  McCoy,  B.  J.  Ind.  Eng.  Chem.  Res.  1997,  36  (6),  2019–

2024.  

(35)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(36)     Kryger,  M.  J.;  Munaretto,  A.  M.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2011,  133  (46),  18992–18998.  

(37)     Adamo,  C.;  Barone,  V.  J.  Chem.  Phys.  1999,  110  (13),  6158–6170.  

(38)     Schäfer,  A.;  Huber,  C.;  Ahlrichs,  R.  J.  Chem.  Phys.  1994,  100  (8),  5829–5835.  

(39)     Blanksby,  S.  J.;  Ellison,  G.  B.  Acc.  Chem.  Res.  2003,  36  (4),  255–263.  

Chapter 2

46

Page 54: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

1H  NMR  spectra  for  EtButIm  and  Pd(NHC-­‐EtBut)2Cl2.  

References  

(1)     Kryger,  M.  J.;  Ong,  M.  T.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Martinez,  T.  J.;  Moore,  J.  S.  J.  

Am.  Chem.  Soc.  2010,  132  (13),  4558–4559.  

(2)     Klukovich,  H.  M.;  Kean,  Z.  S.;  Iacono,  S.  T.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2011,  133  (44),  17882–

17888.  

(3)     Brantley,  J.  N.;  Wiggins,  K.  M.;  Bielawski,  C.  W.  Science  2011,  333  (6049),  1606–1609.  

(4)     Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat  Chem  2010,  2  (6),  436–437.  

(5)     Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  Organometallics  2012,  31  (6),  2476–2481.  

(6)     Powell,  A.  B.;  Bielawski,  C.  W.;  Cowley,  A.  H.  J.  Am.  Chem.  Soc.  2010,  132  (29),  10184–10194.  

(7)     Caruso,  M.  M.;  Davis,  D.  A.;  Shen,  Q.;  Odom,  S.  A.;  Sottos,  N.  R.;  White,  S.  R.;  Moore,  J.  S.  

Chem.  Rev.  2009,  109  (11),  5755–5798.  

(8)     Kuijpers,  M.  W.  A.;  Iedema,  P.  D.;  Kemmere,  M.  F.;  Keurentjes,  J.  T.  F.  Polymer  2004,  45  (19),  

6461–6467.  

(9)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

(10)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.  Int.  Ed.  2004,  43  (34),  4460–4462.  

(11)     Paulusse,  J.  M.  J.;  Huijbers,  J.  P.  J.;  Sijbesma,  R.  P.  Macromolecules  2005,  38  (15),  6290–6298.  

(12)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(13)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2008,  No.  37,  4416–4418.  

(14)     Johansson  Seechurn,  C.  C.  C.;  Kitching,  M.  O.;  Colacot,  T.  J.;  Snieckus,  V.  Angew.  Chem.  Int.  Ed.  

2012,  51  (21),  5062–5085.  

(15)     Hopkinson,  M.  N.;  Richter,  C.;  Schedler,  M.;  Glorius,  F.  Nature  2014,  510  (7506),  485–496.  

(16)     Valente,  C.;  Çalimsiz,  S.;  Hoi,  K.  H.;  Mallik,  D.;  Sayah,  M.;  Organ,  M.  G.  Angew.  Chem.  Int.  Ed.  

2012,  51  (14),  3314–3332.  

(17)     Herrmann,  W.  A.  Angew.  Chem.  Int.  Ed.  2002,  41  (8),  1290–1309.  

(18)     O’Brien,  C.  J.;  Kantchev,  E.  A.  B.;  Valente,  C.;  Hadei,  N.;  Chass,  G.  A.;  Lough,  A.;  Hopkinson,  A.  

C.;  Organ,  M.  G.  Chem.  –  Eur.  J.  2006,  12  (18),  4743–4748.  

(19)     Kantchev,  E.  A.  B.;  O’Brien,  C.  J.;  Organ,  M.  G.  Angew.  Chem.  Int.  Ed.  2007,  46  (16),  2768–

2813.  

(20)     Organ,  M.  G.;  Avola,  S.;  Dubovyk,  I.;  Hadei,  N.;  Kantchev,  E.  A.  B.;  O’Brien,  C.  J.;  Valente,  C.  

Chem.  –  Eur.  J.  2006,  12  (18),  4749–4755.  

(21)     O’Brien,  C.  J.;  Kantchev,  E.  A.  B.;  Chass,  G.  A.;  Hadei,  N.;  Hopkinson,  A.  C.;  Organ,  M.  G.;  

Setiadi,  D.  H.;  Tang,  T.-­‐H.;  Fang,  D.-­‐C.  Tetrahedron  2005,  61  (41),  9723–9735.  

(22)     Christmann,  U.;  Vilar,  R.  Angew.  Chem.  Int.  Ed.  2005,  44  (3),  366–374.  

(23)     Phan,  N.  T.  S.;  Van  Der  Sluys,  M.;  Jones,  C.  W.  Adv.  Synth.  Catal.  2006,  348  (6),  609–679.  

(24)     Marion,  N.;  Nolan,  S.  P.  Acc.  Chem.  Res.  2008,  41  (11),  1440–1449.  

(25)     Karthikeyan,  S.;  Potisek,  S.  L.;  Piermattei,  A.;  Sijbesma,  R.  P.  J.  Am.  Chem.  Soc.  2008,  130  (45),  

14968–14969.  

(26)     Szwarc,  M.  J.  Chem.  Phys.  1948,  16  (2),  128–136.  

(27)     Huynh,  H.  V.;  Han,  Y.;  Jothibasu,  R.;  Yang,  J.  A.  Organometallics  2009,  28  (18),  5395–5404.  

(28)     Liu,  S.-­‐T.;  Lee,  C.-­‐I.;  Fu,  C.-­‐F.;  Chen,  C.-­‐H.;  Liu,  Y.-­‐H.;  Elsevier,  C.  J.;  Peng,  S.-­‐M.;  Chen,  J.-­‐T.  

Organometallics  2009,  28  (24),  6957–6962.  

(29)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(30)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  Chem.  –  Eur.  J.  2009,  15  (48),  13331–13335.  

(31)     Wiggins,  K.  M.;  Brantley,  J.  N.;  Bielawski,  C.  W.  Chem.  Soc.  Rev.  2013,  42  (17),  7130–7147.  

(32)     Groote,  R.;  Szyja,  B.  M.;  Pidko,  E.  A.;  Hensen,  E.  J.  M.;  Sijbesma,  R.  P.  Macromolecules  2011,  

44  (23),  9187–9195.  

(33)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(34)     Madras,  G.;  Chung,  G.  Y.;  Smith,  J.  M.;  McCoy,  B.  J.  Ind.  Eng.  Chem.  Res.  1997,  36  (6),  2019–

2024.  

(35)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(36)     Kryger,  M.  J.;  Munaretto,  A.  M.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2011,  133  (46),  18992–18998.  

(37)     Adamo,  C.;  Barone,  V.  J.  Chem.  Phys.  1999,  110  (13),  6158–6170.  

(38)     Schäfer,  A.;  Huber,  C.;  Ahlrichs,  R.  J.  Chem.  Phys.  1994,  100  (8),  5829–5835.  

(39)     Blanksby,  S.  J.;  Ellison,  G.  B.  Acc.  Chem.  Res.  2003,  36  (4),  255–263.  

Mechanochemical chain scission in NHC-Pd

47

2

Page 55: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

(40)     Bell,  G.  I.  Science  1978,  200  (4342),  618–627.  

(41)     Evans,  E.;  Ritchie,  K.  Biophys.  J.  1997,  72  (4),  1541–1555.  

(42)     Kucharski,  T.  J.;  Boulatov,  R.  J.  Mater.  Chem.  2011,  21  (23),  8237–8255.  

(43)     Beyer,  M.  K.  J.  Chem.  Phys.  2000,  112  (17),  7307–7312.  

(44)     Gossweiler,  G.  R.;  Kouznetsova,  T.  B.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2015,  137  (19),  6148–6151.  

(45)     Wang,  J.;  Kouznetsova,  T.  B.;  Kean,  Z.  S.;  Fan,  L.;  Mar,  B.  D.;  Martínez,  T.  J.;  Craig,  S.  L.  J.  Am.  

Chem.  Soc.  2014,  136  (43),  15162–15165.  

(46)     Brown,  C.  L.;  Craig,  S.  L.  Chem.  Sci.  2015,  6  (4),  2158–2165.  

(47)     Dubreuil,  M.  F.;  Farcy,  N.  G.;  Goethals,  E.  J.  Macromol.  Rapid  Commun.  1999,  20  (7),  383–386.  

 

 

 

 

 

Chapter 2

48

Page 56: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter 3Mechanical scission of Pd(NHC)2Cl2 complexes probed with chemiluminescence

Polytetrahydrofuran (pTHF) functionalized N-heterocyclic carbene (NHC) ligand has been

dissociated mechanically from Pd(NHC-pTHF)2Cl2 complex. A 2-coumaranone derivative,

which decomposes via a chemiluminescent pathway once deprotonated, was used to moni-

tor scission event. Chemiluminescence of coumaranone involves the formation of thermally

labile 1,2-dioxetanone that requires molecular oxygen to form. Therefore, ultrasound was

applied in air-saturated toluene. Influence of saturation gas on the scission rate and molecu-

lar weight threshold (Mlim) for mechanochemical chain scission was determined.

(40)     Bell,  G.  I.  Science  1978,  200  (4342),  618–627.  

(41)     Evans,  E.;  Ritchie,  K.  Biophys.  J.  1997,  72  (4),  1541–1555.  

(42)     Kucharski,  T.  J.;  Boulatov,  R.  J.  Mater.  Chem.  2011,  21  (23),  8237–8255.  

(43)     Beyer,  M.  K.  J.  Chem.  Phys.  2000,  112  (17),  7307–7312.  

(44)     Gossweiler,  G.  R.;  Kouznetsova,  T.  B.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2015,  137  (19),  6148–6151.  

(45)     Wang,  J.;  Kouznetsova,  T.  B.;  Kean,  Z.  S.;  Fan,  L.;  Mar,  B.  D.;  Martínez,  T.  J.;  Craig,  S.  L.  J.  Am.  

Chem.  Soc.  2014,  136  (43),  15162–15165.  

(46)     Brown,  C.  L.;  Craig,  S.  L.  Chem.  Sci.  2015,  6  (4),  2158–2165.  

(47)     Dubreuil,  M.  F.;  Farcy,  N.  G.;  Goethals,  E.  J.  Macromol.  Rapid  Commun.  1999,  20  (7),  383–386.  

 

 

 

 

 

Page 57: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

N-­‐heterocyclic   carbenes   (NHCs)1   are   nucleophilic   and   highly   basic   molecules,   which   have  

been   used   in   various   organic   transformations,   including   condensation,   transesterification  

and  ring-­‐opening  reactions.2,3  Although  they  are  highly  reactive  towards  various  substrates  

in  their  free  form,  their  reactivity  can  be  masked  in  thermally  labile  precursors.4  NHC−metal  

complexes  have  been  applied  as   thermally   latent   catalysts   in   a  number  of  polymerization  

reactions.5–9    

Mechanical   activation   of   chemical   bonds   offers   an   alternative   to   conventional   processes  

such  as  thermal  or  photochemical  activation.12  Mechanical  work  done  by  an  external  force  

lowers  the  energy  barrier  for  bond  dissociation  to  such  an  extent  that  thermal  fluctuations  

can  exceed  this  barrier  at  room  temperature.13  For  efficient  transduction  of  applied  external  

force  onto  mechanically  labile  bonds  of  molecules  in  solution,  attachment  of  polymer  chains  

is   required.   Incorporation   of   functional   groups  with  mechanically   responsive   labile   bonds  

(“mechanophores”)   into  polymers  has  provided  materials   that  are  sensitive   to  mechanical  

stimuli   and   led   to   useful   strain   induced   molecular   transformations.   Mechanoresponsive  

materials  have  gained  more  interest  over  the  last  decade;  e.g.  materials  that  change  color  

upon  strain  or  heal  microcracks  autonomously  by  use  of  proper  mechanophores  have  been  

developed.14,15  

Polymer  attached  NHC−metal  complexes,  for  instance  Ag(NHC-­‐pTHF)2PF6,  have  been  utilized  

in   our   group   as   latent   mechanochemical   catalysts.10,11   Coordination   bonds   in   these  

complexes   were   broken   mechanically   to   release   free   NHC,   which   then   catalyzed  

transesterification   reactions.   In   the   previous   chapter   of   this   thesis,   it   was   shown   that  

Pd(NHC-­‐pTHF)2Cl2   complexes   can   also   be   used   as   thermally   stable   mechanoresponsive  

coordination  polymers  to  release  NHC  upon  sonication.  

Recently,   our   group   reported   the  development  of   a  mechano-­‐luminescent  material   based  

on  a  polymer-­‐functionalized  1,2-­‐dioxetane  moiety.16  Dioxetanes  are  organic  peroxides  and  

efficient  sources  of  electronically  excited  products  upon  chemical  or  thermal  treatments.17  

Opening  of  a  four-­‐membered  1,2-­‐dioxetane  ring  with  two  adjacent  oxygen  atoms  yields  two  

carbonyl   groups,  one  of  which   is   in  an  electronically  excited   state  and  emits  a  photon   (at  

420  nm)  on  relaxation.  Bis(adamantyl)-­‐substituted  1,2-­‐dioxetane  is  thermally  stable  at  room  

temperature  and  has  been  successfully  activated  mechanically  by  straining  polymer  samples  

that   covalently   incorporate   the   luminescent   unit   in   poly(methylacrylate)   chains.  

Experiments   have   shown   that   dioxetanes   can   be   used   as   mechanochemical-­‐probe   for  

spatiotemporal  mapping  and  chain  scission  in  polymers.  A  limitation  of  using  the  dioxetane  

unit   as   a  mechanophore   is   that   for   a   given  dioxetane,   the  barrier   to   scission   is   fixed.   The  

currently   employed   bisadamantyl   dioxetane   has   a   scission   barrier   of   155   kJ/mol,   and   is  

thermally  stable  up  to  ~200  oC.16,18  Lowering  the  barrier  for  decomposition  by  reducing  the  

steric  bulk  of  the  dioxetane  will  also  reduce  its  thermal  stability.  Therefore,  it  is  of  interest  

to   develop   an   alternative   scheme   for   mechanoluminescene   in   which   luminescence   is  

induced  chemically  by  deprotonation  of  a  chemiluminescent  probe  by  a  mechanochemically  

released  base,  such  as  an  NHC-­‐group.  Such  a  scheme  was  first  explored  in  collaboration  with  

Jessica   Clough   in   our   group,   by   studying   the   release   of   NHC   from   polymeric   Ag-­‐carbene  

complexes,   to   initiate   base   induced   chemiluminescence   of   2-­‐Coumaranone.   Although  

luminescence   was   indeed   observed,   the   thermal   lability   of   the   Ag   complexes   prevented  

detailed   analysis   of   the   mechanochemical   effects.     In   the   current   Chapter,   the   use   of  

Pd(NHC)2Cl2   complexes   for   mechanochemical   release   of   NHC   for   the   base   induced  

chemiluminescence  of  coumaranone  is  investigated.  

2-­‐Coumaranones  

Lofthouse   et   al.   reported   the   synthesis   and   chemiluminescence   of   2-­‐coumaranone  

derivatives   in   1979.19   They   reported   that   dimethylformamide   solutions   of   coumaranone  

derivatives   emit   violet   light   in   the  presence  of   triethylamine,   at   room   temperature  under  

air.    Later,  Schramm  et  al.  reported  a  systematic  study  of  the  mechanism  of  this  reaction.  By  

isolating   the   decomposition   products   they   established   that   the   chemiluminescent   step  

involves   the   formation  and  decomposition  of  an  unstable  1,2-­‐dioxetanone   intermediate.20  

Because   the  proposed  1,2-­‐dioxetanone   intermediate  decomposes   rapidly,   it   has  not  been  

isolated  so  far,  but  indirect  proof  for  its  existence  comes  from  the  study  of  charge  transfer  

complexes  with  perylene.  21,22    

Investigation   of   decomposition   mechanisms   for   1,2-­‐dioxetanone   under   sonication  

conditions  is  a  subject  of  a  separate  study  in  our  laboratories  and  beyond  the  scope  of  this  

thesis.    

Chapter 3

50

Page 58: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

N-­‐heterocyclic   carbenes   (NHCs)1   are   nucleophilic   and   highly   basic   molecules,   which   have  

been   used   in   various   organic   transformations,   including   condensation,   transesterification  

and  ring-­‐opening  reactions.2,3  Although  they  are  highly  reactive  towards  various  substrates  

in  their  free  form,  their  reactivity  can  be  masked  in  thermally  labile  precursors.4  NHC−metal  

complexes  have  been  applied  as   thermally   latent   catalysts   in   a  number  of  polymerization  

reactions.5–9    

Mechanical   activation   of   chemical   bonds   offers   an   alternative   to   conventional   processes  

such  as  thermal  or  photochemical  activation.12  Mechanical  work  done  by  an  external  force  

lowers  the  energy  barrier  for  bond  dissociation  to  such  an  extent  that  thermal  fluctuations  

can  exceed  this  barrier  at  room  temperature.13  For  efficient  transduction  of  applied  external  

force  onto  mechanically  labile  bonds  of  molecules  in  solution,  attachment  of  polymer  chains  

is   required.   Incorporation   of   functional   groups  with  mechanically   responsive   labile   bonds  

(“mechanophores”)   into  polymers  has  provided  materials   that  are  sensitive   to  mechanical  

stimuli   and   led   to   useful   strain   induced   molecular   transformations.   Mechanoresponsive  

materials  have  gained  more  interest  over  the  last  decade;  e.g.  materials  that  change  color  

upon  strain  or  heal  microcracks  autonomously  by  use  of  proper  mechanophores  have  been  

developed.14,15  

Polymer  attached  NHC−metal  complexes,  for  instance  Ag(NHC-­‐pTHF)2PF6,  have  been  utilized  

in   our   group   as   latent   mechanochemical   catalysts.10,11   Coordination   bonds   in   these  

complexes   were   broken   mechanically   to   release   free   NHC,   which   then   catalyzed  

transesterification   reactions.   In   the   previous   chapter   of   this   thesis,   it   was   shown   that  

Pd(NHC-­‐pTHF)2Cl2   complexes   can   also   be   used   as   thermally   stable   mechanoresponsive  

coordination  polymers  to  release  NHC  upon  sonication.  

Recently,   our   group   reported   the  development  of   a  mechano-­‐luminescent  material   based  

on  a  polymer-­‐functionalized  1,2-­‐dioxetane  moiety.16  Dioxetanes  are  organic  peroxides  and  

efficient  sources  of  electronically  excited  products  upon  chemical  or  thermal  treatments.17  

Opening  of  a  four-­‐membered  1,2-­‐dioxetane  ring  with  two  adjacent  oxygen  atoms  yields  two  

carbonyl   groups,  one  of  which   is   in  an  electronically  excited   state  and  emits  a  photon   (at  

420  nm)  on  relaxation.  Bis(adamantyl)-­‐substituted  1,2-­‐dioxetane  is  thermally  stable  at  room  

temperature  and  has  been  successfully  activated  mechanically  by  straining  polymer  samples  

that   covalently   incorporate   the   luminescent   unit   in   poly(methylacrylate)   chains.  

Experiments   have   shown   that   dioxetanes   can   be   used   as   mechanochemical-­‐probe   for  

spatiotemporal  mapping  and  chain  scission  in  polymers.  A  limitation  of  using  the  dioxetane  

unit   as   a  mechanophore   is   that   for   a   given  dioxetane,   the  barrier   to   scission   is   fixed.   The  

currently   employed   bisadamantyl   dioxetane   has   a   scission   barrier   of   155   kJ/mol,   and   is  

thermally  stable  up  to  ~200  oC.16,18  Lowering  the  barrier  for  decomposition  by  reducing  the  

steric  bulk  of  the  dioxetane  will  also  reduce  its  thermal  stability.  Therefore,  it  is  of  interest  

to   develop   an   alternative   scheme   for   mechanoluminescene   in   which   luminescence   is  

induced  chemically  by  deprotonation  of  a  chemiluminescent  probe  by  a  mechanochemically  

released  base,  such  as  an  NHC-­‐group.  Such  a  scheme  was  first  explored  in  collaboration  with  

Jessica   Clough   in   our   group,   by   studying   the   release   of   NHC   from   polymeric   Ag-­‐carbene  

complexes,   to   initiate   base   induced   chemiluminescence   of   2-­‐Coumaranone.   Although  

luminescence   was   indeed   observed,   the   thermal   lability   of   the   Ag   complexes   prevented  

detailed   analysis   of   the   mechanochemical   effects.     In   the   current   Chapter,   the   use   of  

Pd(NHC)2Cl2   complexes   for   mechanochemical   release   of   NHC   for   the   base   induced  

chemiluminescence  of  coumaranone  is  investigated.  

2-­‐Coumaranones  

Lofthouse   et   al.   reported   the   synthesis   and   chemiluminescence   of   2-­‐coumaranone  

derivatives   in   1979.19   They   reported   that   dimethylformamide   solutions   of   coumaranone  

derivatives   emit   violet   light   in   the  presence  of   triethylamine,   at   room   temperature  under  

air.    Later,  Schramm  et  al.  reported  a  systematic  study  of  the  mechanism  of  this  reaction.  By  

isolating   the   decomposition   products   they   established   that   the   chemiluminescent   step  

involves   the   formation  and  decomposition  of  an  unstable  1,2-­‐dioxetanone   intermediate.20  

Because   the  proposed  1,2-­‐dioxetanone   intermediate  decomposes   rapidly,   it   has  not  been  

isolated  so  far,  but  indirect  proof  for  its  existence  comes  from  the  study  of  charge  transfer  

complexes  with  perylene.  21,22    

Investigation   of   decomposition   mechanisms   for   1,2-­‐dioxetanone   under   sonication  

conditions  is  a  subject  of  a  separate  study  in  our  laboratories  and  beyond  the  scope  of  this  

thesis.    

Mechanically induced chemiluminescence

51

3

Page 59: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  discussions  

Chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  complexes  in  the  presence  of  coumaranone  

In   order   to   investigate   chain   scission   in   polymers,   Pd(NHC-­‐pTHF)2Cl2   (0.2   mM)   and  

coumaranone   (0.2  mM)  were  sonicated   in   toluene   for  60  min  using  a   sonication  probe.   It  

has   been   shown   that   chemiluminescent   decomposition   of   coumaranone   requires  

deprotonation   followed   by   a   reaction   with   molecular   oxygen.20   Therefore,   dry   air   was  

bubbled   through  the  solution  during  sonication  providing  a  constant  stream  of  gas,  which  

contains   both     molecular   oxygen   required   for   dioxetanone   formation   and   nitrogen   as  

cavitation   gas.   The   double-­‐jacketed   sonication   vessel   was   cooled   down   to   2oC   by   water  

circulation  from  a  thermostat-­‐controlled  bath.  The  temperature  inside  the  sonication  vessel  

was  monitored  with  a   thermocouple  and   recorded  as  below  30   oC  during   sonication  after  

thermal  equilibration  for  approximately  5  min.      

Molecular   weights   of   complexes   were   determined   with   MALDI-­‐TOF   as   shown   in   the  

previous  chapter.  Bimodal  distributions  obtained  with  GPC  were  fitted  with  double  Gaussian  

function,   which   gives   the   best   fit   for   selective   midpoint   scission   of   a   weak   metal–ligand  

coordination   bonds,   to   determine   initial  weight   fractions   of   Pd(NHC-­‐pTHF)2Cl2   complexes.  

The  two  peaks  of  the  bimodal  distribution  were  assumed  to  have  the  same  polydispersity  so  

peak  widths  were   fixed.   Areas   under   GPC   traces  were   calculated   from   the   fits   and  were  

normalized   to   total  peak  area   (i.e.  A1+A2=1).  Concentration  of   starting  material  C1   at  each  

sonication  time  t  is  proportional  to  the  relative  peak  area  in  RI  detection:  

𝐶𝐶! 𝑡𝑡 = 𝐴𝐴! 𝑡𝑡  

Polymer   degradation   during   sonication   follows   first-­‐order   kinetics   and   weight   fraction   of  

starting  material  C1  decays  exponentially  with  the  equation;    𝐶𝐶! 𝑡𝑡 = 𝐶𝐶! 0 ×𝑒𝑒 !!!"!  where  

ksc  is  mechanochemical  scission  rate  coefficient  and  t  is  sonication  time.  Exponential  fitting  

of   C1(t)/C1(0)   vs   t   graph   reveals   the   scission   rate   coefficient   ksc   (R2=0.99)   that   ranges  

between  0.005  and  0.065  min-­‐1.  Chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  complexes  under  air  in  

the   presence   of   2-­‐coumaranone   increases   from   20%   to   98%   for   the   initial   MW   of     the  

polymer  complexes  16  kDa  and  50kDa  respectively.    

GPC   traces   showed   the   formation   of   polymer   species   with   half   of   the   initial  MW   during  

sonication   (Figure   1),   confirming   that   scission   takes   place   and   is   not   reversible,   as   is  

expected   when   free   NHC   is   protonated   by   the   presence   of   2-­‐coumaranone.   GPC   traces  

taken  after  different  time  intervals  during  sonication  show  two  distinct  peaks  corresponding  

to   initial   and   broken   polymer   fractions.   Such   behavior   is   usually   not   observed   for  

polydisperse   polymers   without   a   weak   bond   (see   for   example   ref   23)   and   suggests   that  

scission  is  selective  for  the  mid-­‐chain  coordination  bond.    

In   the   presence   of   2-­‐coumaranone   under   air,   Pd(NHC-­‐pTHF18k)2Cl2   showed   ≈90%   chain  

scission  after  1h  of  continuous  sonication,  while  55%  scission  was  observed  after  sonicating  

for  1h  under  CH4  with  AcOH  and  MeCN  as   trapping  agents.  The  highest  molecular  weight  

complex  Pd(NHC-­‐pTHF25k)2Cl2  was  broken  completely  (≥  98%)   in  1h,  with  25  kDa  polymer  

as   the   only   polymeric   product   and   complete   disappearance   of   the   50kDa   peak  

corresponding  to  the  complex  (Figure  1d).    

   

Chapter 3

52

Page 60: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  discussions  

Chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  complexes  in  the  presence  of  coumaranone  

In   order   to   investigate   chain   scission   in   polymers,   Pd(NHC-­‐pTHF)2Cl2   (0.2   mM)   and  

coumaranone   (0.2  mM)  were  sonicated   in   toluene   for  60  min  using  a   sonication  probe.   It  

has   been   shown   that   chemiluminescent   decomposition   of   coumaranone   requires  

deprotonation   followed   by   a   reaction   with   molecular   oxygen.20   Therefore,   dry   air   was  

bubbled   through  the  solution  during  sonication  providing  a  constant  stream  of  gas,  which  

contains   both     molecular   oxygen   required   for   dioxetanone   formation   and   nitrogen   as  

cavitation   gas.   The   double-­‐jacketed   sonication   vessel   was   cooled   down   to   2oC   by   water  

circulation  from  a  thermostat-­‐controlled  bath.  The  temperature  inside  the  sonication  vessel  

was  monitored  with  a   thermocouple  and   recorded  as  below  30   oC  during   sonication  after  

thermal  equilibration  for  approximately  5  min.      

Molecular   weights   of   complexes   were   determined   with   MALDI-­‐TOF   as   shown   in   the  

previous  chapter.  Bimodal  distributions  obtained  with  GPC  were  fitted  with  double  Gaussian  

function,   which   gives   the   best   fit   for   selective   midpoint   scission   of   a   weak   metal–ligand  

coordination   bonds,   to   determine   initial  weight   fractions   of   Pd(NHC-­‐pTHF)2Cl2   complexes.  

The  two  peaks  of  the  bimodal  distribution  were  assumed  to  have  the  same  polydispersity  so  

peak  widths  were   fixed.   Areas   under   GPC   traces  were   calculated   from   the   fits   and  were  

normalized   to   total  peak  area   (i.e.  A1+A2=1).  Concentration  of   starting  material  C1   at  each  

sonication  time  t  is  proportional  to  the  relative  peak  area  in  RI  detection:  

𝐶𝐶! 𝑡𝑡 = 𝐴𝐴! 𝑡𝑡  

Polymer   degradation   during   sonication   follows   first-­‐order   kinetics   and   weight   fraction   of  

starting  material  C1  decays  exponentially  with  the  equation;    𝐶𝐶! 𝑡𝑡 = 𝐶𝐶! 0 ×𝑒𝑒 !!!"!  where  

ksc  is  mechanochemical  scission  rate  coefficient  and  t  is  sonication  time.  Exponential  fitting  

of   C1(t)/C1(0)   vs   t   graph   reveals   the   scission   rate   coefficient   ksc   (R2=0.99)   that   ranges  

between  0.005  and  0.065  min-­‐1.  Chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  complexes  under  air  in  

the   presence   of   2-­‐coumaranone   increases   from   20%   to   98%   for   the   initial   MW   of     the  

polymer  complexes  16  kDa  and  50kDa  respectively.    

GPC   traces   showed   the   formation   of   polymer   species   with   half   of   the   initial  MW   during  

sonication   (Figure   1),   confirming   that   scission   takes   place   and   is   not   reversible,   as   is  

expected   when   free   NHC   is   protonated   by   the   presence   of   2-­‐coumaranone.   GPC   traces  

taken  after  different  time  intervals  during  sonication  show  two  distinct  peaks  corresponding  

to   initial   and   broken   polymer   fractions.   Such   behavior   is   usually   not   observed   for  

polydisperse   polymers   without   a   weak   bond   (see   for   example   ref   23)   and   suggests   that  

scission  is  selective  for  the  mid-­‐chain  coordination  bond.    

In   the   presence   of   2-­‐coumaranone   under   air,   Pd(NHC-­‐pTHF18k)2Cl2   showed   ≈90%   chain  

scission  after  1h  of  continuous  sonication,  while  55%  scission  was  observed  after  sonicating  

for  1h  under  CH4  with  AcOH  and  MeCN  as   trapping  agents.  The  highest  molecular  weight  

complex  Pd(NHC-­‐pTHF25k)2Cl2  was  broken  completely  (≥  98%)   in  1h,  with  25  kDa  polymer  

as   the   only   polymeric   product   and   complete   disappearance   of   the   50kDa   peak  

corresponding  to  the  complex  (Figure  1d).    

   

Mechanically induced chemiluminescence

53

3

Page 61: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   1:   GPC   traces   for   Pd(NHC-­‐pTHF)2Cl2   taken   during   sonication   in   air   saturated   toluene   (10  

mg/ml)   in   the   presence   of   2-­‐coumaranone   (0.2  mM).   Internal   temperature  was   kept   below   30oC.  

Aliquots  were  taken  at  given  time  intervals.  Toluene  was  evaporated  and  residues  were  submitted  

to  GPC  in  THF  (2  mg/ml).  Initial  molecular  weight  for  polymers:  a)  16  kDa  b)  24  kDa  c)  36  kDa  d)  50  

kDa  

Table   1:  Scission  rates  and  total  scission  after  1h  sonication  for  Pd(NHC-­‐pTHF)2Cl2  polymers     in   the  

presence  of  coumaranone  under  air.    

MW  (kg/mol)   ksc  (min-­‐1)   Scission  (1  h)  

16   0.0055   28  %  

24   0.0150   60  %  

36   0.0360   88  %  

50   0.0600   98  %  

   

Chain   scission  was  determined   from  GPC   traces  using   the  method  discussed  above  and   is  

plotted  as   a   function  of   time   (Figure  2).   Scission   rates   ksc  were  determined  by   fitting   first  

order  rate  constants  to  the  data  in  Figure  2a  with  the  method  presented  in  Chapter  2,  and  

are  given  in  Table  1.  A  value  of  13.5  kDa  was  determined  for  the  limiting  molecular  weight  

(Mlim)  from  linear  extrapolation  of  a  plot  of  ksc  vs  MW  (Figure  2b).  This  value  is  significantly  

lower  than  the  Mlim  determined  previously  for  the  same  complex  when  sonicated  under  CH4  

(20  kDa).    

Empirically,   ksc   in   sonication   usually   increases   with   MW   following   the   equation  𝑘𝑘!"  =

 𝑘𝑘!  (𝑀𝑀𝑀𝑀  −  𝑀𝑀!"#)ƛ  26    with    kd  corresponding  to  the  slope  of  a  plot  of  ksc  vs  MW.  Figure  2b  

shows  that  the  kd   for  the  system  under  air   is  significantly  higher  than  for  sonication  under  

methane  (Table  2).    

Table  2:  kd  (mol  x  g-­‐1  x  min-­‐1)  for  Pd(NHC-­‐pTHF)2Cl2  determined  for  the  sonications  under  air  and  CH4.  

𝒌𝒌𝒅𝒅𝑷𝑷𝑷𝑷 𝒂𝒂𝒂𝒂𝒂𝒂   𝒌𝒌𝒅𝒅𝑷𝑷𝑷𝑷(𝑪𝑪𝑪𝑪𝟒𝟒)  

(1.75 ± 0.1)×10!!   (7.4 ± 0.3)×10!!  

 

In   Chapter   2,   the   kinetics   of   ultrasound   induced   chain   scission   in   Pd(NHC-­‐pTHF)2Cl2  were  

reported.  It  was  shown  that  scission  rates  are  strongly  dependent  on  the  heat  capacity  and  

solubility   of   the   saturation   gas   because   these   parameters   influence   the   formation   and  

implosion  of   cavitation  bubbles,  which  eventually   is   the   source  of  elongational   stresses   in  

solution.24,25   The   molecular   weight   dependence   of   scission   rate   and   limiting   molecular  

Chapter 3

54

Page 62: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   1:   GPC   traces   for   Pd(NHC-­‐pTHF)2Cl2   taken   during   sonication   in   air   saturated   toluene   (10  

mg/ml)   in   the   presence   of   2-­‐coumaranone   (0.2  mM).   Internal   temperature  was   kept   below   30oC.  

Aliquots  were  taken  at  given  time  intervals.  Toluene  was  evaporated  and  residues  were  submitted  

to  GPC  in  THF  (2  mg/ml).  Initial  molecular  weight  for  polymers:  a)  16  kDa  b)  24  kDa  c)  36  kDa  d)  50  

kDa  

Table   1:  Scission  rates  and  total  scission  after  1h  sonication  for  Pd(NHC-­‐pTHF)2Cl2  polymers     in   the  

presence  of  coumaranone  under  air.    

MW  (kg/mol)   ksc  (min-­‐1)   Scission  (1  h)  

16   0.0055   28  %  

24   0.0150   60  %  

36   0.0360   88  %  

50   0.0600   98  %  

   

Chain   scission  was  determined   from  GPC   traces  using   the  method  discussed  above  and   is  

plotted  as   a   function  of   time   (Figure  2).   Scission   rates   ksc  were  determined  by   fitting   first  

order  rate  constants  to  the  data  in  Figure  2a  with  the  method  presented  in  Chapter  2,  and  

are  given  in  Table  1.  A  value  of  13.5  kDa  was  determined  for  the  limiting  molecular  weight  

(Mlim)  from  linear  extrapolation  of  a  plot  of  ksc  vs  MW  (Figure  2b).  This  value  is  significantly  

lower  than  the  Mlim  determined  previously  for  the  same  complex  when  sonicated  under  CH4  

(20  kDa).    

Empirically,   ksc   in   sonication   usually   increases   with   MW   following   the   equation  𝑘𝑘!"  =

 𝑘𝑘!  (𝑀𝑀𝑀𝑀  −  𝑀𝑀!"#)ƛ  26    with    kd  corresponding  to  the  slope  of  a  plot  of  ksc  vs  MW.  Figure  2b  

shows  that  the  kd   for  the  system  under  air   is  significantly  higher  than  for  sonication  under  

methane  (Table  2).    

Table  2:  kd  (mol  x  g-­‐1  x  min-­‐1)  for  Pd(NHC-­‐pTHF)2Cl2  determined  for  the  sonications  under  air  and  CH4.  

𝒌𝒌𝒅𝒅𝑷𝑷𝑷𝑷 𝒂𝒂𝒂𝒂𝒂𝒂   𝒌𝒌𝒅𝒅𝑷𝑷𝑷𝑷(𝑪𝑪𝑪𝑪𝟒𝟒)  

(1.75 ± 0.1)×10!!   (7.4 ± 0.3)×10!!  

 

In   Chapter   2,   the   kinetics   of   ultrasound   induced   chain   scission   in   Pd(NHC-­‐pTHF)2Cl2  were  

reported.  It  was  shown  that  scission  rates  are  strongly  dependent  on  the  heat  capacity  and  

solubility   of   the   saturation   gas   because   these   parameters   influence   the   formation   and  

implosion  of   cavitation  bubbles,  which  eventually   is   the   source  of  elongational   stresses   in  

solution.24,25   The   molecular   weight   dependence   of   scission   rate   and   limiting   molecular  

Mechanically induced chemiluminescence

55

3

Page 63: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

weight  for  Pd(NHC-­‐pTHF)2Cl2  for  sonication  under  CH4  were  given  in  Chapter  2  are  different  

than  those  under  air.    

Groote   has   previously   studied   the   effect   of   saturation   gas   on   scission   rate   for   Ag(NHC-­‐

pTHF)2PF6   polymeric   complex   and   reported   that   scission   proceeds  with   comparable   rates  

under  N2  and  CH4.24  Therefore,   it  can  be  concluded  that  the  presence  of  20%  of  O2,  has  a  

strong   effect   on   scission   rate.   This   may   be   due   to   the   higher   solubility   of   O2   in   toluene  

compared   to   N2   since   this   means   cavitation   bubbles   involve   mainly   O2   rather   than   N2.27  

Reactive   sonochemical   impurities   produced   due   to   pyrolysis   of   the   content   of   cavitation  

bubbles   in   hot   spots28   might   result   in   chemical   breakage   of   polymer   and   increase   the  

scission  rate.    

 

Figure   2:   a)   Chain   scission   rates   for   Pd(NHC-­‐pTHF)2Cl2   complexes   under   air   in   the   presence   of  

coumaranone.   b)   Scission   rate   constants   ksc   plotted   against   initial   molecular   weights   of   polymers  

under   air   (with   coumaranone)   and   under   CH4   (with   CH3COOH   and   CH3CN).   MW   of   polymer  

complexes  are  16  kDa,  24  kDa,  36  kDa  and  50  kDa.  

 

 

 

 

 

Chapter 3

56

Page 64: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

weight  for  Pd(NHC-­‐pTHF)2Cl2  for  sonication  under  CH4  were  given  in  Chapter  2  are  different  

than  those  under  air.    

Groote   has   previously   studied   the   effect   of   saturation   gas   on   scission   rate   for   Ag(NHC-­‐

pTHF)2PF6   polymeric   complex   and   reported   that   scission   proceeds  with   comparable   rates  

under  N2  and  CH4.24  Therefore,   it  can  be  concluded  that  the  presence  of  20%  of  O2,  has  a  

strong   effect   on   scission   rate.   This   may   be   due   to   the   higher   solubility   of   O2   in   toluene  

compared   to   N2   since   this   means   cavitation   bubbles   involve   mainly   O2   rather   than   N2.27  

Reactive   sonochemical   impurities   produced   due   to   pyrolysis   of   the   content   of   cavitation  

bubbles   in   hot   spots28   might   result   in   chemical   breakage   of   polymer   and   increase   the  

scission  rate.    

 

Figure   2:   a)   Chain   scission   rates   for   Pd(NHC-­‐pTHF)2Cl2   complexes   under   air   in   the   presence   of  

coumaranone.   b)   Scission   rate   constants   ksc   plotted   against   initial   molecular   weights   of   polymers  

under   air   (with   coumaranone)   and   under   CH4   (with   CH3COOH   and   CH3CN).   MW   of   polymer  

complexes  are  16  kDa,  24  kDa,  36  kDa  and  50  kDa.  

 

 

 

 

 

Mechanically induced chemiluminescence

57

3

Page 65: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Light  emission  

Mechanobase   induced   chemiluminescence   was   studied   by   sonication   of   air   saturated  

toluene   solutions   containing   coumaranone   1   and   Pd(NHC-­‐pTHF)2Cl2   or   small   molecule  

model  complex  Pd(NHC-­‐EtBut)2Cl2,  and  monitoring  light  emission  with  a  photodiode.  During  

experiments,   the   sonication   flask   was   kept   in   the   dark   to   prevent   interference   of  

background  light.  The  double-­‐jacketed  sonication  vessel  was  cooled  down  to  2oC  by  water  

circulation   from   a   thermostat-­‐   controlled   bath   to   keep   the   temperature   inside   the  

sonication  vessel  constant  below  30  oC.    

When  a  10  mg/ml   (0.2  mM)   solution  of   Pd(NHC-­‐pTHF25k)2Cl2  in   toluene  was   sonicated   in  

the   presence   of   2-­‐coumaranone   (0.2   mM),   light   intensity   increased   slowly   as   shown   in  

Figure   3,   indicating   chemiluminescent   decomposition   of   coumaranone.   Light   intensity  

reached   a   maximum   level   after   about   300   s.,   after   which   it   slowly   decreased.   When  

sonication   was   stopped   after   850   s.,   the   photodiode   response   dropped   back   to   its  

background  value  in  approximately  70s.    

 

Figure   3:  A  0.2  mM  solution  of  Pd(NHC-­‐pTHF25k)2Cl2  in  air   saturated   toluene  was  sonicated   in   the  

presence  of  2-­‐coumaranone  (0.2  mM).  Temperature  inside  the  sonication  vessel  was  kept  below  30  oC  Luminescence  was  monitored  with  a  photodiode  (inset),  which  was  placed  under  the  sonication  

flask  during  sonication.  Light  intensity  was  measured  as  the  change  in  current.      

 

In   a   control   experiment   (Scheme  1),   a   (0.2  mM)   solution   of   low  molecular  weight   analog  

Pd(NHC-­‐EtBut)2Cl2,   model   complex,   was   sonicated   in   the   presence   of   coumaranone   (0.2  

mM)  and  methoxy-­‐end-­‐capped  pTHF   (30  kDa,  0.2  mM).  Photodiode   response  did  not   rise  

above  the  background  level  during  sonication.  Additonally,  refluxing  the  mixture  of  Pd(NHC-­‐

pTHF25k)2Cl2   (10   mg/ml)   and   coumaranone   in   toluene   did   not   result   in   detectable  

chemiluminescence.   These   observations   rule   out   the   contribution   of   thermal   ligand-­‐

dissociation  to  base  induced  chemiluminescence.    

To  confirm  that  the  of  presence  of  O2  is  essential  in  mechano-­‐chemiluminescence,  a  mixture  

of  Pd(NHC-­‐pTHF25k)2Cl2  and  coumaranone  was  sonicated  under  CH4  otherwise  identical  to  

above  mentioned  sonication  conditions.  No  chemiluminescence  was  observed.  

In   order   to   test   its   thermal   stability,   Pd(NHC-­‐pTHF25k)2Cl2   (10  mg/ml)     and   coumaranone  

(0.2   mM)   were   kept   in   a   toluene   solution   at   RT   for   30   days.   Then   a   0.1   ml   aliquot   was  

submitted  to  GPC  and  the  Mw  distribution  was  unchanged.  When  the  stored  solution  was  

sonicated,  the  maximum  light  intensity  was  identical  to  that  of  a  freshly  prepared  solution  

 

Scheme  1:  Mechano-­‐chemiluminescence  a  result  of  decomposition  of  coumaranone  when  sonicated  

in  the  presence  of  Pd(NHC-­‐pTHF)2Cl2.  Light  emission  was  not  detected  when  PTHF  and  the  Pd-­‐NHC  

complex  were  both  present  in  the  solution  but  not  covalently  attached  to  each  other.        

 

N N

Pd

NN

ClCl R¹

O

HNOR²

OO

hvN N

Pd

NN

ClCl

O

OR¹

O

HNOR²

OO

OOn

n

n

hvx

-Pd(NHC/pTHF12k)₂Cl₂--Pd(NHC/pTHF18k)₂Cl₂--Pd(NHC/pTHF25k)₂Cl₂

PTHF:-50-kDA

+

+ +

Chapter 3

58

Page 66: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Light  emission  

Mechanobase   induced   chemiluminescence   was   studied   by   sonication   of   air   saturated  

toluene   solutions   containing   coumaranone   1   and   Pd(NHC-­‐pTHF)2Cl2   or   small   molecule  

model  complex  Pd(NHC-­‐EtBut)2Cl2,  and  monitoring  light  emission  with  a  photodiode.  During  

experiments,   the   sonication   flask   was   kept   in   the   dark   to   prevent   interference   of  

background  light.  The  double-­‐jacketed  sonication  vessel  was  cooled  down  to  2oC  by  water  

circulation   from   a   thermostat-­‐   controlled   bath   to   keep   the   temperature   inside   the  

sonication  vessel  constant  below  30  oC.    

When  a  10  mg/ml   (0.2  mM)   solution  of   Pd(NHC-­‐pTHF25k)2Cl2  in   toluene  was   sonicated   in  

the   presence   of   2-­‐coumaranone   (0.2   mM),   light   intensity   increased   slowly   as   shown   in  

Figure   3,   indicating   chemiluminescent   decomposition   of   coumaranone.   Light   intensity  

reached   a   maximum   level   after   about   300   s.,   after   which   it   slowly   decreased.   When  

sonication   was   stopped   after   850   s.,   the   photodiode   response   dropped   back   to   its  

background  value  in  approximately  70s.    

 

Figure   3:  A  0.2  mM  solution  of  Pd(NHC-­‐pTHF25k)2Cl2  in  air   saturated   toluene  was  sonicated   in   the  

presence  of  2-­‐coumaranone  (0.2  mM).  Temperature  inside  the  sonication  vessel  was  kept  below  30  oC  Luminescence  was  monitored  with  a  photodiode  (inset),  which  was  placed  under  the  sonication  

flask  during  sonication.  Light  intensity  was  measured  as  the  change  in  current.      

 

In   a   control   experiment   (Scheme  1),   a   (0.2  mM)   solution   of   low  molecular  weight   analog  

Pd(NHC-­‐EtBut)2Cl2,   model   complex,   was   sonicated   in   the   presence   of   coumaranone   (0.2  

mM)  and  methoxy-­‐end-­‐capped  pTHF   (30  kDa,  0.2  mM).  Photodiode   response  did  not   rise  

above  the  background  level  during  sonication.  Additonally,  refluxing  the  mixture  of  Pd(NHC-­‐

pTHF25k)2Cl2   (10   mg/ml)   and   coumaranone   in   toluene   did   not   result   in   detectable  

chemiluminescence.   These   observations   rule   out   the   contribution   of   thermal   ligand-­‐

dissociation  to  base  induced  chemiluminescence.    

To  confirm  that  the  of  presence  of  O2  is  essential  in  mechano-­‐chemiluminescence,  a  mixture  

of  Pd(NHC-­‐pTHF25k)2Cl2  and  coumaranone  was  sonicated  under  CH4  otherwise  identical  to  

above  mentioned  sonication  conditions.  No  chemiluminescence  was  observed.  

In   order   to   test   its   thermal   stability,   Pd(NHC-­‐pTHF25k)2Cl2   (10  mg/ml)     and   coumaranone  

(0.2   mM)   were   kept   in   a   toluene   solution   at   RT   for   30   days.   Then   a   0.1   ml   aliquot   was  

submitted  to  GPC  and  the  Mw  distribution  was  unchanged.  When  the  stored  solution  was  

sonicated,  the  maximum  light  intensity  was  identical  to  that  of  a  freshly  prepared  solution  

 

Scheme  1:  Mechano-­‐chemiluminescence  a  result  of  decomposition  of  coumaranone  when  sonicated  

in  the  presence  of  Pd(NHC-­‐pTHF)2Cl2.  Light  emission  was  not  detected  when  PTHF  and  the  Pd-­‐NHC  

complex  were  both  present  in  the  solution  but  not  covalently  attached  to  each  other.        

 

N N

Pd

NN

ClCl R¹

O

HNOR²

OO

hvN N

Pd

NN

ClCl

O

OR¹

O

HNOR²

OO

OOn

n

n

hvx

-Pd(NHC/pTHF12k)₂Cl₂--Pd(NHC/pTHF18k)₂Cl₂--Pd(NHC/pTHF25k)₂Cl₂

PTHF:-50-kDA

+

+ +

Mechanically induced chemiluminescence

59

3

Page 67: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

The  changes  in  chemiluminescence  intensity  with  time  in  Figure  3  reflect  the  kinetics  of  the  

individual   steps   of   the   process,  which   consists   of   scission   of   the   carbene  metal   complex,  

deprotonation   of   coumaranone,   oxidation   of   the   anion   to   dioxetanone,   and   finally  

decomposition  of  the  dioxetanone  under  emission  of  a  photon,  with  the  light  intensity  being  

proportional   to   the  concentration  of  dioxetanone.   If  scission  would  be  the  slowest  step   in  

the  process,  light  intensity  would  be  proportional  to  the  rate  at  which  the  complex  breaks,  

and  simple  first  order  decay  would  be  observed  with  a  rate  constant  corresponding  to  the  

scission  rate  constant  of  1  x  10-­‐3  s-­‐1.  The  initial  rise  in  intensity  in  Figure  3  therefore  shows  

that   scission   is   not   the   slowest   step,   and   reflects   the   accumulation   of   an   intermediate,  

either   the   anion   or   the   2-­‐dioxetanone.   The   latter   has   a   reported   decomposition   rate  

constant  kD  =  1.08(±0.01)  x10-­‐2  s-­‐1  in  MeCN  at  25oC.21  This  rate  constant  matches  reasonably  

well  with  the  fast  decrease  of  signal  when  sonication  was  stopped.  Since  deprotonation  of  

the   coumaranone   may   be   assumed   to   be   fast,   we   conclude   that   decomposition   of  

intermediate  is  the  slowest  step.  Simulation  of  the  process  with  kinetic  modeling  software,  

using   the   parameters   in   Scheme   2   indeed   reproduces   qualitatively   the   observed   rise   and  

decay  in  light  intensity  (Figure  4).      

 

Scheme  2:  Schematic  representation  of  mechanochemiluminescence,  estimated  rate  constants  and  

proposed   chemiluminescent   decomposition   mechanism   in   the   presence   of   base   and   oxygen.  

Ultrasound  induced  chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  yields  free  carbene  which  abstracts  a  proton  

from  2-­‐coumaranone  and  then  it  decomposes  through  a  luminescent  pathway.  

 

 

 

F

O

HN

OBu

O

O O2

1-2 dioxetanone2-coumaranone

hv

F

O

HN

OBu

O

OF

O

OO

O

HN

BuO O

CO2

F

OH

NH

OBu

O O

+

1 32

3

N N

Pd

NN

ClCl

O

O

n

n

N NOn

Pd

NN

ClCl

On

Ultrasound ksc$:$0.001$s)¹

k₁$:$0.030$s)¹

k₃$>>$k)₃

k₁$:$0.011$s)¹

k₄$>>$k)₄

[PMP]₀6:60.26mM

[Coum]₀6:60.26mM

Chapter 3

60

Page 68: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

The  changes  in  chemiluminescence  intensity  with  time  in  Figure  3  reflect  the  kinetics  of  the  

individual   steps   of   the   process,  which   consists   of   scission   of   the   carbene  metal   complex,  

deprotonation   of   coumaranone,   oxidation   of   the   anion   to   dioxetanone,   and   finally  

decomposition  of  the  dioxetanone  under  emission  of  a  photon,  with  the  light  intensity  being  

proportional   to   the  concentration  of  dioxetanone.   If  scission  would  be  the  slowest  step   in  

the  process,  light  intensity  would  be  proportional  to  the  rate  at  which  the  complex  breaks,  

and  simple  first  order  decay  would  be  observed  with  a  rate  constant  corresponding  to  the  

scission  rate  constant  of  1  x  10-­‐3  s-­‐1.  The  initial  rise  in  intensity  in  Figure  3  therefore  shows  

that   scission   is   not   the   slowest   step,   and   reflects   the   accumulation   of   an   intermediate,  

either   the   anion   or   the   2-­‐dioxetanone.   The   latter   has   a   reported   decomposition   rate  

constant  kD  =  1.08(±0.01)  x10-­‐2  s-­‐1  in  MeCN  at  25oC.21  This  rate  constant  matches  reasonably  

well  with  the  fast  decrease  of  signal  when  sonication  was  stopped.  Since  deprotonation  of  

the   coumaranone   may   be   assumed   to   be   fast,   we   conclude   that   decomposition   of  

intermediate  is  the  slowest  step.  Simulation  of  the  process  with  kinetic  modeling  software,  

using   the   parameters   in   Scheme   2   indeed   reproduces   qualitatively   the   observed   rise   and  

decay  in  light  intensity  (Figure  4).      

 

Scheme  2:  Schematic  representation  of  mechanochemiluminescence,  estimated  rate  constants  and  

proposed   chemiluminescent   decomposition   mechanism   in   the   presence   of   base   and   oxygen.  

Ultrasound  induced  chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  yields  free  carbene  which  abstracts  a  proton  

from  2-­‐coumaranone  and  then  it  decomposes  through  a  luminescent  pathway.  

 

 

 

F

O

HN

OBu

O

O O2

1-2 dioxetanone2-coumaranone

hv

F

O

HN

OBu

O

OF

O

OO

O

HN

BuO O

CO2

F

OH

NH

OBu

O O

+

1 32

3

N N

Pd

NN

ClCl

O

O

n

n

N NOn

Pd

NN

ClCl

On

Ultrasound ksc$:$0.001$s)¹

k₁$:$0.030$s)¹

k₃$>>$k)₃

k₁$:$0.011$s)¹

k₄$>>$k)₄

[PMP]₀6:60.26mM

[Coum]₀6:60.26mM

Mechanically induced chemiluminescence

61

3

Page 69: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Example   of   simulated   light   intensity   (proportional   to   [Cox])   with   following   kinetic  parameters:    

PMP  =  PM  +  P  k1  =  0.001  s-­‐1  (ksc  for  Pd(NHC-­‐pTHF25k)2Cl2  taken  from  Table  1),  k-­‐1  =  108  M-­‐1s-­‐1  

(backward  reaction  was  assumed  to  be  diffusion  controlled)  

P   +   C   =   PH   +   Canion   k2   =   108   s-­‐1   (assumed   to   be   diffusion   controlled),   k-­‐2   =   10-­‐4   M-­‐1s-­‐1  

(assumption:  k2>>k-­‐2)  

Canion  +  O2  =  Cox    k3  =  0.03  s-­‐1  (assumption:  k3>>k-­‐3)    

Cox  =  Cdecomp    k4  =  0.011  s-­‐1  (k4  is  taken  from  ref  21,  assumption:  k4>>k-­‐4,)    

Backward   rates   are   included   because   gepasi   cannot   handle   zero   reverse   rates.   Initial  

concentrations  were  taken  as  [PMP]  =  0.2  mM  [C]  =  0.2  mM  [O2]  =  1  M  (value  was  set  high  

to  keep  [O2]  constant).  

 

Figure  4:  Simulated  change  in  light  intensity  (proportional  to  [Cox])   is  compared  with  experimental  

data.  GEPASI21  was  used  for  simulation  with  kinetic  parameters  in  Scheme  1.    

 

Figure   5:   Comparison   of   the   light   intensity   during   sonication   of   Pd(NHC-­‐pTHF)2Cl2  polymers   in   the  

presence  of    2-­‐coumaranone.  MW  for  polymers  are  50  kDa,  36  kDa  and  24  kDa.    

Conclusions  

Chain   scission   in   the   presence   of   coumaranone   has   been   investigated.   Base   induced  

chemiluminescent  property  of  coumaranone  provided  a  molecular  probe  to  monitor  chain  

scission   in  situ.  Kinetic  studies   revealed  that   the  NHC-­‐Pd  bond   is  mechanically  more   labile  

under   air   than   under   CH4.   The   lowest   molecular   weight   for   polymeric   complex   Pd(NHC-­‐

pTHF)2Cl2  to  break  under  ultrasound  induced  strain  was  calculated  as  13.5  kDa.  Above  that  

MW   the   rate   of   mechanochemical   chain   scission   increases   linearly   with   initial   Mw.  

Sonicating   Pd(NHC-­‐pTHF)2Cl2   complexes   reveals   free   NHC-­‐pTHF   ligands   which   abstracts  

proton   from   coumaranone   that   is   followed   by   a   dioxetanone   and   its   subsequent  

chemiluminescent   decomposition.   Mechanically   induced   release   of   a   base   to   activate  

chemiluminescence  could  be  monitored  by  a  photodiode.  Control  experiments  showed  that  

the  chemiluminescence  is  a  result  of  mechanical  activation  NHC-­‐Pd  bond  to  yield  free  NHC.  

It   has   been   shown  here   that   Pd(NHC-­‐pTHF)2Cl2  complexes   are   thermally   stable   sources  of  

base  that  can  be  released  under  mechanical  stress  in  due  course.    

These   preliminary   results   represent   the   first   promising   example   of   mechano-­‐

chemiluminescence   and   in   the   presence   of   2-­‐coumaranone,   mechanochemical   chain  

Chapter 3

62

Page 70: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Example   of   simulated   light   intensity   (proportional   to   [Cox])   with   following   kinetic  parameters:    

PMP  =  PM  +  P  k1  =  0.001  s-­‐1  (ksc  for  Pd(NHC-­‐pTHF25k)2Cl2  taken  from  Table  1),  k-­‐1  =  108  M-­‐1s-­‐1  

(backward  reaction  was  assumed  to  be  diffusion  controlled)  

P   +   C   =   PH   +   Canion   k2   =   108   s-­‐1   (assumed   to   be   diffusion   controlled),   k-­‐2   =   10-­‐4   M-­‐1s-­‐1  

(assumption:  k2>>k-­‐2)  

Canion  +  O2  =  Cox    k3  =  0.03  s-­‐1  (assumption:  k3>>k-­‐3)    

Cox  =  Cdecomp    k4  =  0.011  s-­‐1  (k4  is  taken  from  ref  21,  assumption:  k4>>k-­‐4,)    

Backward   rates   are   included   because   gepasi   cannot   handle   zero   reverse   rates.   Initial  

concentrations  were  taken  as  [PMP]  =  0.2  mM  [C]  =  0.2  mM  [O2]  =  1  M  (value  was  set  high  

to  keep  [O2]  constant).  

 

Figure  4:  Simulated  change  in  light  intensity  (proportional  to  [Cox])   is  compared  with  experimental  

data.  GEPASI21  was  used  for  simulation  with  kinetic  parameters  in  Scheme  1.    

 

Figure   5:   Comparison   of   the   light   intensity   during   sonication   of   Pd(NHC-­‐pTHF)2Cl2  polymers   in   the  

presence  of    2-­‐coumaranone.  MW  for  polymers  are  50  kDa,  36  kDa  and  24  kDa.    

Conclusions  

Chain   scission   in   the   presence   of   coumaranone   has   been   investigated.   Base   induced  

chemiluminescent  property  of  coumaranone  provided  a  molecular  probe  to  monitor  chain  

scission   in  situ.  Kinetic  studies   revealed  that   the  NHC-­‐Pd  bond   is  mechanically  more   labile  

under   air   than   under   CH4.   The   lowest   molecular   weight   for   polymeric   complex   Pd(NHC-­‐

pTHF)2Cl2  to  break  under  ultrasound  induced  strain  was  calculated  as  13.5  kDa.  Above  that  

MW   the   rate   of   mechanochemical   chain   scission   increases   linearly   with   initial   Mw.  

Sonicating   Pd(NHC-­‐pTHF)2Cl2   complexes   reveals   free   NHC-­‐pTHF   ligands   which   abstracts  

proton   from   coumaranone   that   is   followed   by   a   dioxetanone   and   its   subsequent  

chemiluminescent   decomposition.   Mechanically   induced   release   of   a   base   to   activate  

chemiluminescence  could  be  monitored  by  a  photodiode.  Control  experiments  showed  that  

the  chemiluminescence  is  a  result  of  mechanical  activation  NHC-­‐Pd  bond  to  yield  free  NHC.  

It   has   been   shown  here   that   Pd(NHC-­‐pTHF)2Cl2  complexes   are   thermally   stable   sources  of  

base  that  can  be  released  under  mechanical  stress  in  due  course.    

These   preliminary   results   represent   the   first   promising   example   of   mechano-­‐

chemiluminescence   and   in   the   presence   of   2-­‐coumaranone,   mechanochemical   chain  

Mechanically induced chemiluminescence

63

3

Page 71: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

scission   can   be   visually   monitored,   which   could   be   used   for   mapping   deformations   in  

polymer  samples.      

Experimental  

General  

All   chemicals   were   purchased   from   commercial   sources   used   without   further   purification,   unless  

specified  otherwise.  Dry  tetrahydrofuran  (THF,  HPLC  grade)  was  degassed  with  argon  and  purified  by  

passage  through  activated  alumina  solvent  column.  Toluene  was  dried  over  4Å  molecular  sieves  for  

at   least   12h   prior   to   use.   Gel   permeation   chromatography   (GPC)   was   performed   on   a   Shimadzu  

LC10-­‐AD,   using   Polymer   Laboratories   PL  Gel   5μm  MIXED-­‐C   and  MIXEDD   columns     (linear   range  of  

MW:  200–2000000  g/mol),  a  Shimadzu  SPD-­‐M10A  UV-­‐vis  detector  at  254  nm  and  RID-­‐10A  refractive  

index  detector,   and  THF  as  eluent  at  a   flow   rate  of  1  mL/min   (20   °C).  Polystyrene   standards  were  

used  for  calibration.      

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   vessel   with   a   volume   of   10   mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   A   silicon   photodiode   (Hanamatsu,   diameter   of   photosensitive   area   7   mm)  

positioned  underneath   the  vessel  during   sonication   for  mechano-­‐chemiluminescence  experiments.  

The  set-­‐up  was  covered  to  exclude  background  light.  The  temperature  in  the  vessel  was  maintained  

with   a   Lauda   E300   cooling   bath   and  measured  using   a   0.5  mm  diameter   thermocouple.   Solutions  

were   sonicated   continuously.   Temperature   of   the   solution   was   checked   by   thermocouple   and  

recorded   as   below   30oC   after   the   thermal   equilibrium   was   achieved   in   the   first   5   mins.   During  

sonication  air  was  bubbled  through  the  solution  via  teflon  tubing.  Aliquots  of  100  µL  were  taken  at  

different  time  intervals  toluene  was  removed  under  reduced  pressure  and  residues  were  dissolved  

in  THF  and  submitted  to  GPC.     In  order  to  calculate  mechanochemical  scission  rates  GPC  traces  for  

samples  taken  during  sonication  were  fitted  using  Origin  8.5  to  following  equation:  

𝑰𝑰 𝒙𝒙 =𝑨𝑨𝟏𝟏

𝒘𝒘 𝝅𝝅 𝟐𝟐𝒆𝒆!𝟐𝟐

𝒙𝒙!𝒙𝒙𝒄𝒄,𝟏𝟏𝒘𝒘

𝟐𝟐

+𝑨𝑨𝟐𝟐

𝒘𝒘 𝝅𝝅 𝟐𝟐𝒆𝒆!𝟐𝟐

𝒙𝒙!𝒙𝒙𝒄𝒄,𝟐𝟐𝒘𝒘

𝟐𝟐

 

Where;    

A1  =  Area  of  peak  1,  high  molecular  weight  peak  

A2  =  Area  of  peak  2,  low  molecular  weight  peak  

w  =  2σ  =  Width  of  both  peaks  

X  =  Retention  time  

Xc,1  =  Retention  time  at  the  center  of  peak  1  

Xc,2  =  Retention  time  at  the  center  of  peak  2  

All  peaks  were  normalized  to  total  peak  area  (A1+A2=1).  All  GPC  traces  during  sonication  were  fitted  

with  the  indicated  formula,  fixing  values  for  Xc,1,  Xc,2,  and  w  to  the  same  values  for  each  samples.      

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Chapter 3

64

Page 72: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

scission   can   be   visually   monitored,   which   could   be   used   for   mapping   deformations   in  

polymer  samples.      

Experimental  

General  

All   chemicals   were   purchased   from   commercial   sources   used   without   further   purification,   unless  

specified  otherwise.  Dry  tetrahydrofuran  (THF,  HPLC  grade)  was  degassed  with  argon  and  purified  by  

passage  through  activated  alumina  solvent  column.  Toluene  was  dried  over  4Å  molecular  sieves  for  

at   least   12h   prior   to   use.   Gel   permeation   chromatography   (GPC)   was   performed   on   a   Shimadzu  

LC10-­‐AD,   using   Polymer   Laboratories   PL  Gel   5μm  MIXED-­‐C   and  MIXEDD   columns     (linear   range  of  

MW:  200–2000000  g/mol),  a  Shimadzu  SPD-­‐M10A  UV-­‐vis  detector  at  254  nm  and  RID-­‐10A  refractive  

index  detector,   and  THF  as  eluent  at  a   flow   rate  of  1  mL/min   (20   °C).  Polystyrene   standards  were  

used  for  calibration.      

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   vessel   with   a   volume   of   10   mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   A   silicon   photodiode   (Hanamatsu,   diameter   of   photosensitive   area   7   mm)  

positioned  underneath   the  vessel  during   sonication   for  mechano-­‐chemiluminescence  experiments.  

The  set-­‐up  was  covered  to  exclude  background  light.  The  temperature  in  the  vessel  was  maintained  

with   a   Lauda   E300   cooling   bath   and  measured  using   a   0.5  mm  diameter   thermocouple.   Solutions  

were   sonicated   continuously.   Temperature   of   the   solution   was   checked   by   thermocouple   and  

recorded   as   below   30oC   after   the   thermal   equilibrium   was   achieved   in   the   first   5   mins.   During  

sonication  air  was  bubbled  through  the  solution  via  teflon  tubing.  Aliquots  of  100  µL  were  taken  at  

different  time  intervals  toluene  was  removed  under  reduced  pressure  and  residues  were  dissolved  

in  THF  and  submitted  to  GPC.     In  order  to  calculate  mechanochemical  scission  rates  GPC  traces  for  

samples  taken  during  sonication  were  fitted  using  Origin  8.5  to  following  equation:  

𝑰𝑰 𝒙𝒙 =𝑨𝑨𝟏𝟏

𝒘𝒘 𝝅𝝅 𝟐𝟐𝒆𝒆!𝟐𝟐

𝒙𝒙!𝒙𝒙𝒄𝒄,𝟏𝟏𝒘𝒘

𝟐𝟐

+𝑨𝑨𝟐𝟐

𝒘𝒘 𝝅𝝅 𝟐𝟐𝒆𝒆!𝟐𝟐

𝒙𝒙!𝒙𝒙𝒄𝒄,𝟐𝟐𝒘𝒘

𝟐𝟐

 

Where;    

A1  =  Area  of  peak  1,  high  molecular  weight  peak  

A2  =  Area  of  peak  2,  low  molecular  weight  peak  

w  =  2σ  =  Width  of  both  peaks  

X  =  Retention  time  

Xc,1  =  Retention  time  at  the  center  of  peak  1  

Xc,2  =  Retention  time  at  the  center  of  peak  2  

All  peaks  were  normalized  to  total  peak  area  (A1+A2=1).  All  GPC  traces  during  sonication  were  fitted  

with  the  indicated  formula,  fixing  values  for  Xc,1,  Xc,2,  and  w  to  the  same  values  for  each  samples.      

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Mechanically induced chemiluminescence

65

3

Page 73: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

References  

(1)     Fèvre,  M.;  Pinaud,  J.;  Gnanou,  Y.;  Vignolle,   J.;  Taton,  D.  Chem.  Soc.  Rev.  2013,  42   (5),  2142–

2172.  

(2)     Marion,  N.;  Díez-­‐González,  S.;  Nolan,  S.  P.  Angew.  Chem.  Int.  Ed.  2007,  46  (17),  2988–3000.  

(3)     Hopkinson,  M.  N.;  Richter,  C.;  Schedler,  M.;  Glorius,  F.  Nature  2014,  510  (7506),  485–496.  

(4)     Moore,  J.  L.;  Rovis,  T.  Top.  Curr.  Chem.  2010,  291.  

(5)     Bantu,  B.;  Pawar,  G.  M.;  Decker,  U.;  Wurst,  K.;  Schmidt,  A.  M.;  Buchmeiser,  M.  R.  Chem.  –  Eur.  

J.  2009,  15  (13),  3103–3109.  

(6)     Naumann,  S.;  Schmidt,  F.  G.;  Schowner,  R.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4  

(9),  2731–2740.  

(7)     Naumann,  S.;  Schmidt,  F.  G.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4   (15),  4172–

4181.  

(8)     Naumann,  S.;  Schmidt,  F.  G.;  Speiser,  M.;  Böhl,  M.;  Epple,  S.;  Bonten,  C.;  Buchmeiser,  M.  R.  

Macromolecules  2013,  46  (21),  8426–8433.  

(9)     Naumann,  S.;  Speiser,  M.;  Schowner,  R.;  Giebel,  E.;  Buchmeiser,  M.  R.  Macromolecules  2014,  

47  (14),  4548–4556.  

(10)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(11)     Karthikeyan,  S.;  Potisek,  S.  L.;  Piermattei,  A.;  Sijbesma,  R.  P.  J.  Am.  Chem.  Soc.  2008,  130  (45),  

14968–14969.  

(12)     Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat  Chem  2010,  2  (6),  436–437.  

(13)     Beyer,  M.  K.  J.  Chem.  Phys.  2000,  112  (17),  7307–7312.  

(14)     Davis,  D.  A.;  Hamilton,  A.;   Yang,   J.;   Cremar,   L.  D.;  Van  Gough,  D.;  Potisek,   S.   L.;  Ong,  M.  T.;  

Braun,  P.  V.;  Martínez,  T.  J.;  White,  S.  R.;  Moore,  J.  S.;  Sottos,  N.  R.  Nature  2009,  459  (7243),  

68–72.  

(15)     Black  Ramirez,  A.   L.;  Kean,  Z.  S.;  Orlicki,   J.  A.;  Champhekar,  M.;  Elsakr,  S.  M.;  Krause,  W.  E.;  

Craig,  S.  L.  Nat.  Chem.  2013,  5  (9),  757–761.  

(16)     Chen,  Y.;  Spiering,  A.  J.  H.;  Karthikeyan,  S.;  Peters,  G.  W.  M.;  Meijer,  E.  W.;  Sijbesma,  R.  P.  Nat.  

Chem.  2012,  4  (7),  559–562.  

(17)     Turro,  N.  J.;  Lechtken,  P.;  Schore,  N.  E.;  Schuster,  G.;  Steinmetzer,  H.  C.;  Yekta,  A.  Acc.  Chem.  

Res.  1974,  7  (4),  97–105.  

(18)     Schuster,  G.  B.;  Turro,  N.  J.;  Steinmetzer,  H.  C.;  Schaap,  A.  P.;  Faler,  G.;  Adam,  W.;  Liu,  J.  C.  J.  

Am.  Chem.  Soc.  1975,  97  (24),  7110–7118.  

(19)     Lofthouse,  G.  J.;  Suschitzky,  H.;  Wakefield,  B.  J.;  Whittaker,  R.  A.;  Tuck,  B.  J.  Chem.  Soc.  [Perkin  

1]  1979,  No.  0,  1634–1639.  

(20)     Schramm,   S.;   Weiss,   D.;   Navizet,   I.;   Roca-­‐Sanjuán,   D.;   Brandl,   W.;   Beckert,   R.;   Görls,   H.  

ARKIVOC  2013,  iii,  174–188.  

(21)     Ciscato,   L.   F.   M.   L.;   Bartoloni,   F.   H.;   Colavite,   A.   S.;   Weiss,   D.;   Beckert,   R.;   Schramm,   S.  

Photochem.  Photobiol.  Sci.  2013,  13  (1),  32–37.  

(22)     Almeida   de   Oliveira,   M.;   Bartoloni,   F.   H.;   Augusto,   F.   A.;   Ciscato,   L.   F.   M.   L.;   Bastos,   E.   L.;  

Baader,  W.  J.  J.  Org.  Chem.  2012,  77  (23),  10537–10544.  

(23)     Odell,  J.  A.;  Keller,  A.  J.  Polym.  Sci.  Part  B  Polym.  Phys.  1986,  24  (9),  1889–1916.  

(24)     Rooze,   J.;   Groote,   R.;   Jakobs,   R.   T.   M.;   Sijbesma,   R.   P.;   van   Iersel,   M.   M.;   Rebrov,   E.   V.;  

Schouten,  J.  C.;  Keurentjes,  J.  T.  F.  J.  Phys.  Chem.  B  2011,  115  (38),  11038–11043.  

(25)     Groote,  R.;  Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  ACS  Macro  Lett.  2012,  1  (8),  1012–1015.  

(26)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(27)     Field,  L.  R.;  Wilhelm,  E.;  Battino,  R.  J.  Chem.  Thermodyn.  1974,  6  (3),  237–243.  

(28)     Szwarc,  M.  J.  Chem.  Phys.  1948,  16  (2),  128–136.  

 

 

Chapter 3

66

Page 74: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

References  

(1)     Fèvre,  M.;  Pinaud,  J.;  Gnanou,  Y.;  Vignolle,   J.;  Taton,  D.  Chem.  Soc.  Rev.  2013,  42   (5),  2142–

2172.  

(2)     Marion,  N.;  Díez-­‐González,  S.;  Nolan,  S.  P.  Angew.  Chem.  Int.  Ed.  2007,  46  (17),  2988–3000.  

(3)     Hopkinson,  M.  N.;  Richter,  C.;  Schedler,  M.;  Glorius,  F.  Nature  2014,  510  (7506),  485–496.  

(4)     Moore,  J.  L.;  Rovis,  T.  Top.  Curr.  Chem.  2010,  291.  

(5)     Bantu,  B.;  Pawar,  G.  M.;  Decker,  U.;  Wurst,  K.;  Schmidt,  A.  M.;  Buchmeiser,  M.  R.  Chem.  –  Eur.  

J.  2009,  15  (13),  3103–3109.  

(6)     Naumann,  S.;  Schmidt,  F.  G.;  Schowner,  R.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4  

(9),  2731–2740.  

(7)     Naumann,  S.;  Schmidt,  F.  G.;  Frey,  W.;  Buchmeiser,  M.  R.  Polym.  Chem.  2013,  4   (15),  4172–

4181.  

(8)     Naumann,  S.;  Schmidt,  F.  G.;  Speiser,  M.;  Böhl,  M.;  Epple,  S.;  Bonten,  C.;  Buchmeiser,  M.  R.  

Macromolecules  2013,  46  (21),  8426–8433.  

(9)     Naumann,  S.;  Speiser,  M.;  Schowner,  R.;  Giebel,  E.;  Buchmeiser,  M.  R.  Macromolecules  2014,  

47  (14),  4548–4556.  

(10)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(11)     Karthikeyan,  S.;  Potisek,  S.  L.;  Piermattei,  A.;  Sijbesma,  R.  P.  J.  Am.  Chem.  Soc.  2008,  130  (45),  

14968–14969.  

(12)     Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat  Chem  2010,  2  (6),  436–437.  

(13)     Beyer,  M.  K.  J.  Chem.  Phys.  2000,  112  (17),  7307–7312.  

(14)     Davis,  D.  A.;  Hamilton,  A.;   Yang,   J.;   Cremar,   L.  D.;  Van  Gough,  D.;  Potisek,   S.   L.;  Ong,  M.  T.;  

Braun,  P.  V.;  Martínez,  T.  J.;  White,  S.  R.;  Moore,  J.  S.;  Sottos,  N.  R.  Nature  2009,  459  (7243),  

68–72.  

(15)     Black  Ramirez,  A.   L.;  Kean,  Z.  S.;  Orlicki,   J.  A.;  Champhekar,  M.;  Elsakr,  S.  M.;  Krause,  W.  E.;  

Craig,  S.  L.  Nat.  Chem.  2013,  5  (9),  757–761.  

(16)     Chen,  Y.;  Spiering,  A.  J.  H.;  Karthikeyan,  S.;  Peters,  G.  W.  M.;  Meijer,  E.  W.;  Sijbesma,  R.  P.  Nat.  

Chem.  2012,  4  (7),  559–562.  

(17)     Turro,  N.  J.;  Lechtken,  P.;  Schore,  N.  E.;  Schuster,  G.;  Steinmetzer,  H.  C.;  Yekta,  A.  Acc.  Chem.  

Res.  1974,  7  (4),  97–105.  

(18)     Schuster,  G.  B.;  Turro,  N.  J.;  Steinmetzer,  H.  C.;  Schaap,  A.  P.;  Faler,  G.;  Adam,  W.;  Liu,  J.  C.  J.  

Am.  Chem.  Soc.  1975,  97  (24),  7110–7118.  

(19)     Lofthouse,  G.  J.;  Suschitzky,  H.;  Wakefield,  B.  J.;  Whittaker,  R.  A.;  Tuck,  B.  J.  Chem.  Soc.  [Perkin  

1]  1979,  No.  0,  1634–1639.  

(20)     Schramm,   S.;   Weiss,   D.;   Navizet,   I.;   Roca-­‐Sanjuán,   D.;   Brandl,   W.;   Beckert,   R.;   Görls,   H.  

ARKIVOC  2013,  iii,  174–188.  

(21)     Ciscato,   L.   F.   M.   L.;   Bartoloni,   F.   H.;   Colavite,   A.   S.;   Weiss,   D.;   Beckert,   R.;   Schramm,   S.  

Photochem.  Photobiol.  Sci.  2013,  13  (1),  32–37.  

(22)     Almeida   de   Oliveira,   M.;   Bartoloni,   F.   H.;   Augusto,   F.   A.;   Ciscato,   L.   F.   M.   L.;   Bastos,   E.   L.;  

Baader,  W.  J.  J.  Org.  Chem.  2012,  77  (23),  10537–10544.  

(23)     Odell,  J.  A.;  Keller,  A.  J.  Polym.  Sci.  Part  B  Polym.  Phys.  1986,  24  (9),  1889–1916.  

(24)     Rooze,   J.;   Groote,   R.;   Jakobs,   R.   T.   M.;   Sijbesma,   R.   P.;   van   Iersel,   M.   M.;   Rebrov,   E.   V.;  

Schouten,  J.  C.;  Keurentjes,  J.  T.  F.  J.  Phys.  Chem.  B  2011,  115  (38),  11038–11043.  

(25)     Groote,  R.;  Jakobs,  R.  T.  M.;  Sijbesma,  R.  P.  ACS  Macro  Lett.  2012,  1  (8),  1012–1015.  

(26)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(27)     Field,  L.  R.;  Wilhelm,  E.;  Battino,  R.  J.  Chem.  Thermodyn.  1974,  6  (3),  237–243.  

(28)     Szwarc,  M.  J.  Chem.  Phys.  1948,  16  (2),  128–136.  

 

 

Mechanically induced chemiluminescence

67

3

Page 75: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

   

Page 76: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter 4Determination of ligand exchange dynamics and sonication induced ligand exchange rate in Imidazole-Pd centered coordination polymers

Thermal ligand exchange between imidazole Palladium (II) complexes has been investigated

in the presence of excess free imidazole ligand in CHCl3. Ligand exchange between bound

and free ligands in a mixture of Pd(EtIm)2Cl2 and EtIm was shown to occur through an asso-

ciative process via intermediate [Pd(EtIm)3Cl]Cl. NMR and MALDI-TOF established the struc-

ture of the intermediate. In contrast to the thermal process, ligand exchange was dissocia-

tive when mixtures of polymer functionalized complex Pd(Im-pTHF)2Cl2 and low molecular

weight complex Pd(DodecIm)2Cl2 were sonicated in solution. 40% of exchange was observed

after 1h of continuous sonication at 20 °C.

   

Page 77: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Coordination   polymers   (CPs),   with   their   dynamic   reversible   nature,   find   applications   as  

smart   materials,   for   instance,   as   stimuli-­‐responsive,   self-­‐healing   and   photoactive  

polymers.1,2  Recent  attention  in  this  research  area  is  focused  on  the  characterization  of  the  

dynamic   properties   of   CPs   by   exploring   their   exchange   kinetics,3   ring–chain   equilibria,4  

solvent  interactions,5  and  new  metal–  ligand  combinations.6    

We   have   shown   that   the   coordination   sphere   of   transition-­‐metal   complexes   in   the  main  

chain  of  CPs  can  be  manipulated  by  means  of  mechanical  force.7  As  demonstrated  by  us  and  

other   research   groups,   transition  metal   CPs   can  be  used   as  mechanoresponsive  polymers  

and   provide   opportunity   for   mechanical   release   of   reactive   groups   that   can   be   used   in  

further  reactions,  e.g.  as  catalytically  active  sites.8    

Ultrasound   is   one   of   the   most   efficient   techniques   to   create   forces   that   break   bonds   of  

polymer  backbones  in  solution.9  In  order  to  effectively  transduce  elongational  strain  fields  in  

solution,  high  molecular  weight  polymers  are  required.10  Incorporation  of  weak  bonds  into  a  

polymer  chain  increases  the  sensitivity  of  the  chain  to  mechanical  force  and  thus  decreases  

the   molecular   weight   threshold   for   mechanochemical   chain   scission.   Groote   et   al.  

investigated   the   scission   of   metal-­‐ligand   bonds   in   supramolecular   polymer   complexes   by  

ultrasound  and  showed  that  the  force  required  to  break  metal−ligand  bond  is  much  lower  

than  the  force  that  is  typically  required  to  break  covalent  bonds.11  Thus,  ultrasonication  of  

reversible  coordination  polymers  enables  the  specific  rupture  of  metal   ligand  coordination  

bonds  since  force  selectively  breaks  the  weakest  bond  on  the  chain.  

One  of  the  first  reports  on  mechanochemistry  of  CP’s  involved  high-­‐molecular-­‐weight  linear  

polymers   of   alkyldicyclohexylphosphine   (ADChP)   telechelic   polytetrahydrofuran   with  

palladium(II)  dichloride.7,12  Molecular  weights  of  these  polymers  were  altered  reversibly  by  

ultrasonic  chain  scission.  In  order  to  scavenge  the  reactive  chain  ends  created  by  ultrasonic  

scission,   coordinatively   unsaturated   Pd(II)   and   phosphine,   bisalkyldiphenylphosphine  

(ADPhP)  complex  of  palladium(II)  dichloride  was  added.  This  low  molecular  weight  complex  

shows   high   reactivity   towards   ADChP   ligands   and   coordinatively   unsaturated   Pd,   but   not  

towards   the   coordination   polymer.   High   nucleophilicity   of   ADChP   resulted   in   complete  

displacement  of  ADPhP  ligand.4  These  results  aroused  our  curiosity  towards  understanding  

the  mechanisms  and  kinetics  of  ligand  exchange  triggered  by  ultrasound.  

Insight  in  the  dynamics  of  ligand  exchange  in  coordination  polymers  is  required  for  rational  

use  in  diverse  applications  such  as  responsive  materials,  where  the  ligand  exchange  kinetics  

determine  mechanical   properties,   and  mechano-­‐catalysis,  where   relative   rates   of   scission  

and  association  determine  the  steady  state  concentration  of  active  species.    

Generally,  ligand  exchange  occurs  by  one  of  two  mechanisms:  dissociative  or  associative.13–

16   In   a   dissociative   mechanism   one   of   the   ligands   coordinated   to   the   metal   center  

dissociates,  leaving  an  electron  deficient  complex.  Then,  a  free  ligand  coordinates  to  metal,  

resulting   in   ligand  exchange.   In  transition  metal  complexes,  dissociative   ligand  exchange  is  

favored   in   18e-­‐   systems   to   avoid   the   formation   of   an   energetically   unfavorable   20e-­‐  

intermediate.  On  the  other  hand,  ligand  exchange  occurs  through  an  associative  mechanism  

for   16e-­‐   square   planar   complexes.   The   intermediate   in   this   case   is   an   18e-­‐   complex   and  

therefore  provides  a  lower  energy  route  to  the  product  than  the  14e-­‐  intermediate,  formed  

via  dissociative  ligand  exchange.    

In   the   current   Chapter,   the   thermal   ligand   exchange  mechanism   (Scheme   1)   and   rate   of  

ultrasound-­‐induced   scission   of   Pd-­‐Imidazole   CP’s   are   investigated.   Pd-­‐Imidazole  

coordination  complexes  were  selected   to   learn  how   ligand  exchange  could  be   initiated  by  

mechanical  force.  This  knowledge  can  be  used  in  further  studies  such  as  solid-­‐state  catalysis  

and   self-­‐healing  materials.   Initial   attempts  and   some  preliminary   results   for   the   latter   are  

introduced  in  Chapter  6.    

Chapter 4

70

Page 78: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Coordination   polymers   (CPs),   with   their   dynamic   reversible   nature,   find   applications   as  

smart   materials,   for   instance,   as   stimuli-­‐responsive,   self-­‐healing   and   photoactive  

polymers.1,2  Recent  attention  in  this  research  area  is  focused  on  the  characterization  of  the  

dynamic   properties   of   CPs   by   exploring   their   exchange   kinetics,3   ring–chain   equilibria,4  

solvent  interactions,5  and  new  metal–  ligand  combinations.6    

We   have   shown   that   the   coordination   sphere   of   transition-­‐metal   complexes   in   the  main  

chain  of  CPs  can  be  manipulated  by  means  of  mechanical  force.7  As  demonstrated  by  us  and  

other   research   groups,   transition  metal   CPs   can  be  used   as  mechanoresponsive  polymers  

and   provide   opportunity   for   mechanical   release   of   reactive   groups   that   can   be   used   in  

further  reactions,  e.g.  as  catalytically  active  sites.8    

Ultrasound   is   one   of   the   most   efficient   techniques   to   create   forces   that   break   bonds   of  

polymer  backbones  in  solution.9  In  order  to  effectively  transduce  elongational  strain  fields  in  

solution,  high  molecular  weight  polymers  are  required.10  Incorporation  of  weak  bonds  into  a  

polymer  chain  increases  the  sensitivity  of  the  chain  to  mechanical  force  and  thus  decreases  

the   molecular   weight   threshold   for   mechanochemical   chain   scission.   Groote   et   al.  

investigated   the   scission   of   metal-­‐ligand   bonds   in   supramolecular   polymer   complexes   by  

ultrasound  and  showed  that  the  force  required  to  break  metal−ligand  bond  is  much  lower  

than  the  force  that  is  typically  required  to  break  covalent  bonds.11  Thus,  ultrasonication  of  

reversible  coordination  polymers  enables  the  specific  rupture  of  metal   ligand  coordination  

bonds  since  force  selectively  breaks  the  weakest  bond  on  the  chain.  

One  of  the  first  reports  on  mechanochemistry  of  CP’s  involved  high-­‐molecular-­‐weight  linear  

polymers   of   alkyldicyclohexylphosphine   (ADChP)   telechelic   polytetrahydrofuran   with  

palladium(II)  dichloride.7,12  Molecular  weights  of  these  polymers  were  altered  reversibly  by  

ultrasonic  chain  scission.  In  order  to  scavenge  the  reactive  chain  ends  created  by  ultrasonic  

scission,   coordinatively   unsaturated   Pd(II)   and   phosphine,   bisalkyldiphenylphosphine  

(ADPhP)  complex  of  palladium(II)  dichloride  was  added.  This  low  molecular  weight  complex  

shows   high   reactivity   towards   ADChP   ligands   and   coordinatively   unsaturated   Pd,   but   not  

towards   the   coordination   polymer.   High   nucleophilicity   of   ADChP   resulted   in   complete  

displacement  of  ADPhP  ligand.4  These  results  aroused  our  curiosity  towards  understanding  

the  mechanisms  and  kinetics  of  ligand  exchange  triggered  by  ultrasound.  

Insight  in  the  dynamics  of  ligand  exchange  in  coordination  polymers  is  required  for  rational  

use  in  diverse  applications  such  as  responsive  materials,  where  the  ligand  exchange  kinetics  

determine  mechanical   properties,   and  mechano-­‐catalysis,  where   relative   rates   of   scission  

and  association  determine  the  steady  state  concentration  of  active  species.    

Generally,  ligand  exchange  occurs  by  one  of  two  mechanisms:  dissociative  or  associative.13–

16   In   a   dissociative   mechanism   one   of   the   ligands   coordinated   to   the   metal   center  

dissociates,  leaving  an  electron  deficient  complex.  Then,  a  free  ligand  coordinates  to  metal,  

resulting   in   ligand  exchange.   In  transition  metal  complexes,  dissociative   ligand  exchange  is  

favored   in   18e-­‐   systems   to   avoid   the   formation   of   an   energetically   unfavorable   20e-­‐  

intermediate.  On  the  other  hand,  ligand  exchange  occurs  through  an  associative  mechanism  

for   16e-­‐   square   planar   complexes.   The   intermediate   in   this   case   is   an   18e-­‐   complex   and  

therefore  provides  a  lower  energy  route  to  the  product  than  the  14e-­‐  intermediate,  formed  

via  dissociative  ligand  exchange.    

In   the   current   Chapter,   the   thermal   ligand   exchange  mechanism   (Scheme   1)   and   rate   of  

ultrasound-­‐induced   scission   of   Pd-­‐Imidazole   CP’s   are   investigated.   Pd-­‐Imidazole  

coordination  complexes  were  selected   to   learn  how   ligand  exchange  could  be   initiated  by  

mechanical  force.  This  knowledge  can  be  used  in  further  studies  such  as  solid-­‐state  catalysis  

and   self-­‐healing  materials.   Initial   attempts  and   some  preliminary   results   for   the   latter   are  

introduced  in  Chapter  6.    

Determination of ligand exchange dynamics

71

4

Page 79: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 Scheme  1:  Potential  ligand  exchange  pathways  between  complex  Pd(EtIm)2Cl2  and  free  ligand  EtIm  

 

 

 

 

 

 

 

 

 

 

 

NN Pd NCl

N

Cl+ N N

NN Pd NCl

N

Cl

+

N

N

NN Pd NCl

N

Cl

N

N

NN Pd NN

ClN

N+ Cl

'

NN Pd N

N

ClN

N

Cl

NN Pd NCl

N

Cl

+N

N

Dissociative

Associative

Results  and  Discussions  

Synthesis  of  Pd-­‐Imidazole  complexes  

   

Scheme  2:  Synthesis  of  polymer  attached  complex  and  model  complexes.  

Im-­‐pTHF  ligands  were  prepared  by  terminating  poly  tetrahydrofuran  (pTHF)  with  imidazole  

(Im)  (Scheme  2).  Cationic  ring  opening  polymerization  of  THF  was  performed  as  described  in  

Chapter   2.   Polymerization   was   terminated   by   nucleophilic   substitution   of   the   sodium  

imidazolate,  obtained  by  deprotonation  of  1H-­‐imidazole  with  NaH  in  THF.  Once  Im-­‐pTHF  is  

formed,  the  nucleophilic  N(3)  nitrogen  can  react  further  with  cationic  chain  ends  to  form  1,3  

pTHF-­‐imidazolium   salt.   Therefore,   sodium   imidazolate   end-­‐capper   was   used   five-­‐fold   in  

excess   relative   to   methyl   triflate   initiator.   Molecular   weights   of   resulting   polymers   were  

analyzed   by   1H   NMR,  MALDI-­‐TOF   and   gel   permeation   chromatography   (GPC).   MW   value  

obtained   in   MALDI-­‐TOF   was   used   for   the   further   calculations.   Depending   on   the  

polymerization  time  Im-­‐pTHFs  with  MW  of  18  kDa  and  35  kDa  were  prepared.    

 

 

Chapter 4

72

Page 80: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 Scheme  1:  Potential  ligand  exchange  pathways  between  complex  Pd(EtIm)2Cl2  and  free  ligand  EtIm  

 

 

 

 

 

 

 

 

 

 

 

NN Pd NCl

N

Cl+ N N

NN Pd NCl

N

Cl

+

N

N

NN Pd NCl

N

Cl

N

N

NN Pd NN

ClN

N+ Cl

'

NN Pd N

N

ClN

N

Cl

NN Pd NCl

N

Cl

+N

N

Dissociative

Associative

Results  and  Discussions  

Synthesis  of  Pd-­‐Imidazole  complexes  

   

Scheme  2:  Synthesis  of  polymer  attached  complex  and  model  complexes.  

Im-­‐pTHF  ligands  were  prepared  by  terminating  poly  tetrahydrofuran  (pTHF)  with  imidazole  

(Im)  (Scheme  2).  Cationic  ring  opening  polymerization  of  THF  was  performed  as  described  in  

Chapter   2.   Polymerization   was   terminated   by   nucleophilic   substitution   of   the   sodium  

imidazolate,  obtained  by  deprotonation  of  1H-­‐imidazole  with  NaH  in  THF.  Once  Im-­‐pTHF  is  

formed,  the  nucleophilic  N(3)  nitrogen  can  react  further  with  cationic  chain  ends  to  form  1,3  

pTHF-­‐imidazolium   salt.   Therefore,   sodium   imidazolate   end-­‐capper   was   used   five-­‐fold   in  

excess   relative   to   methyl   triflate   initiator.   Molecular   weights   of   resulting   polymers   were  

analyzed   by   1H   NMR,  MALDI-­‐TOF   and   gel   permeation   chromatography   (GPC).   MW   value  

obtained   in   MALDI-­‐TOF   was   used   for   the   further   calculations.   Depending   on   the  

polymerization  time  Im-­‐pTHFs  with  MW  of  18  kDa  and  35  kDa  were  prepared.    

 

 

Determination of ligand exchange dynamics

73

4

Page 81: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Polymer  functionalized  imidazolyl  ligands  Im-­‐pTHF,  EtIm  and  DodecIm  were  coordinated  to  

Pd   via   exchange   of   the   acetonitrile   (MeCN)   ligand   in   Pd(MeCN)2Cl2   to   give  

mechanoresponsive   complex   Pd(Im-­‐pTHF)2Cl2   and   low   molecular   weight   complexes  

Pd(EtIm)2Cl2  and  Pd(DodecIm)2Cl2.  MeCN-­‐Pd  coordination  is  sufficiently  labile  to  allow  ligand  

exchange   at   room   temperature   in   dichloromethane   (DCM).   Coordination   of   imidazole  

ligands  was  evident   from  the  change   in  chemical  shifts   for  H(2),  H(4)  and  H(5)  protons  on  

the  imidazole  heterocycle.  In  the  Pd(Im-­‐pTHF)2Cl2  complex,  chemical  shifts  for  the  imidazolyl  

moiety   were   identical   to   those   in   the   small   molecule   counterparts   Pd(EtIm)2Cl2   and  

Pd(DodecIm)2Cl2.   1H  NMR   spectra  were   in   good  agreement  with   literature  data  published  

previously.17    

Im-­‐pTHF  coordination  reactions  were  completed  after  2h  as  molecular  weight  distributions  

monitored   by   GPC   did   not   change   after   2h.   GPC   also   showed   that   reaction   of   polymer  

functionalized  ligand  Im-­‐pTHF  (Mn  =  18  kDa)  resulted  in  doubling  of  the  molecular  weight  to  

36  kDa  (Figure  1).  This  indicates  that  two  imidazoles  are  coordinated  to  Pd  although  it  has  

been  reported  that  coordination  of  three  imidazole  ligands  to  Pd  (II)  may  also  take  place  at  

higher  ligand  to  metal  ratio  and  specific  reaction  conditions.17    

 

Figure  1:  1H  NMR  spectra  (left,  in  CD2Cl2)  and  GPC  traces  (right,  in  THF)  for  Im-­‐pTHF18k  and  Pd(Im-­‐

pTHF18k)2Cl2.    

 

 

 

Ligand  exchange  between  EtIm  and  Pd(EtIm)2Cl2  

Structure  of  the  intermediate    

In  order  to  determine  the  ligand  exchange  mechanism  for  palladium-­‐imidazole  complexes,  

free   ligand  EtIm   (L)  and  Pd(EtIm)2Cl2   (C)  were  mixed   in  a  1:1  molar   ratio   (0.09  M  each)   in  

CDCl3.  The  1H-­‐NMR  spectrum  (Figure  2)  revealed  the  coexistence  of  three  magnetically  non-­‐

equivalent   species,   L,   coordination   complexes   C   and   intermediate,   [Pd(EtIm)3Cl]Cl   (I).  

MALDI-­‐TOF   spectrum   indicated   that   tris-­‐imidazole   coordinated   complex   [Pd(EtIm)3Cl]Cl  

formed  when  free  ligand  and  Pd(EtIm)2Cl2  were  mixed  (Figure  3).    

 Figure  2:    Stacked  1H  NMR  spectra  for  EtIm  (b),  Pd(EtIm)2Cl2  (c)  and  [Pd(EtIm)3Cl]Cl  (a)  in  CDCl3  (400  

MHz).    

Chapter 4

74

Page 82: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Polymer  functionalized  imidazolyl  ligands  Im-­‐pTHF,  EtIm  and  DodecIm  were  coordinated  to  

Pd   via   exchange   of   the   acetonitrile   (MeCN)   ligand   in   Pd(MeCN)2Cl2   to   give  

mechanoresponsive   complex   Pd(Im-­‐pTHF)2Cl2   and   low   molecular   weight   complexes  

Pd(EtIm)2Cl2  and  Pd(DodecIm)2Cl2.  MeCN-­‐Pd  coordination  is  sufficiently  labile  to  allow  ligand  

exchange   at   room   temperature   in   dichloromethane   (DCM).   Coordination   of   imidazole  

ligands  was  evident   from  the  change   in  chemical  shifts   for  H(2),  H(4)  and  H(5)  protons  on  

the  imidazole  heterocycle.  In  the  Pd(Im-­‐pTHF)2Cl2  complex,  chemical  shifts  for  the  imidazolyl  

moiety   were   identical   to   those   in   the   small   molecule   counterparts   Pd(EtIm)2Cl2   and  

Pd(DodecIm)2Cl2.   1H  NMR   spectra  were   in   good  agreement  with   literature  data  published  

previously.17    

Im-­‐pTHF  coordination  reactions  were  completed  after  2h  as  molecular  weight  distributions  

monitored   by   GPC   did   not   change   after   2h.   GPC   also   showed   that   reaction   of   polymer  

functionalized  ligand  Im-­‐pTHF  (Mn  =  18  kDa)  resulted  in  doubling  of  the  molecular  weight  to  

36  kDa  (Figure  1).  This  indicates  that  two  imidazoles  are  coordinated  to  Pd  although  it  has  

been  reported  that  coordination  of  three  imidazole  ligands  to  Pd  (II)  may  also  take  place  at  

higher  ligand  to  metal  ratio  and  specific  reaction  conditions.17    

 

Figure  1:  1H  NMR  spectra  (left,  in  CD2Cl2)  and  GPC  traces  (right,  in  THF)  for  Im-­‐pTHF18k  and  Pd(Im-­‐

pTHF18k)2Cl2.    

 

 

 

Ligand  exchange  between  EtIm  and  Pd(EtIm)2Cl2  

Structure  of  the  intermediate    

In  order  to  determine  the  ligand  exchange  mechanism  for  palladium-­‐imidazole  complexes,  

free   ligand  EtIm   (L)  and  Pd(EtIm)2Cl2   (C)  were  mixed   in  a  1:1  molar   ratio   (0.09  M  each)   in  

CDCl3.  The  1H-­‐NMR  spectrum  (Figure  2)  revealed  the  coexistence  of  three  magnetically  non-­‐

equivalent   species,   L,   coordination   complexes   C   and   intermediate,   [Pd(EtIm)3Cl]Cl   (I).  

MALDI-­‐TOF   spectrum   indicated   that   tris-­‐imidazole   coordinated   complex   [Pd(EtIm)3Cl]Cl  

formed  when  free  ligand  and  Pd(EtIm)2Cl2  were  mixed  (Figure  3).    

 Figure  2:    Stacked  1H  NMR  spectra  for  EtIm  (b),  Pd(EtIm)2Cl2  (c)  and  [Pd(EtIm)3Cl]Cl  (a)  in  CDCl3  (400  

MHz).    

Determination of ligand exchange dynamics

75

4

Page 83: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  3:  MALDI-­‐TOF  spectrum  of   [Pd(EtIm)3Cl]Cl  taken  in  CHCl3,  1mg/ml.   Inset:  chemical  structure  

of    [Pd(EtIm)3Cl]Cl  and  its  simulated  mass  spectrum.    

2D  EXSY  spectra  

The   mechanism   of   exchange   between   free   and   bound   Im   was   investigated   using   2D  

exchange   spectroscopy   (EXSY18,19).   To  monitor   the   ligand  exchange,   complex  2  was  mixed  

with  free  ligand  in  deuterated  chloroform  (CDCl3).  2D  EXSY  spectra  were  taken  in  CDCl3  at  RT  

with   several   different   mixing   times   (tmix).   Spectra   indicate   that   free   Im   (which   is   not  

coordinated   to   Pd   center,   abbreviated   as   L)   exchanges  with   Im   coordinated   to   Pd   on   the  

time  scale  of  typical  tmix  of  100-­‐500  ms.  When  tmix  was  lower  than  200  ms,  spectra  showed  

cross  peaks   for  exchange  of   L  with   intermediate   [Pd(EtIm)3Cl]Cl   (I)  and  for  Pd(EtIm)2Cl2  (C)    

with   I.   However,   only   at   higher   mixing   times   (>200   ms)   cross   peaks   for   L-­‐C   exchange  

becomes  visible  (Figure  4).    

 

 

 

 

 

 

 

Figure  4:  2D  EXSY  spectra  of  0.09  M  solution  of  N-­‐ethyl  Imidazole  and  Pd(EtIm)2Cl2  in  CDCl3  at  tmix  =  

200  ms  and  tmix  =  500  ms.    Circles  were  used  to  highlight  cross  peaks  appear  when  tmix  is  high  (≥  300  

ms).  

In  order  to  investigate  exchange  rate  constants,  ratios  of  cross  peaks  to  diagonal  peaks  were  

calculated   using   absolute   peak   volumes   at   each   tmix.     C-­‐L   crosspeaks   are   absent   at   low  

mixing  times  and  only  appear  at  mixing  times  above  200  ms.  The  absence  of  direct  exchange  

between  C  and  L   in  the  presence  of  C-­‐I  and  L-­‐I  cross  peaks  demonstrates  the  intermediacy  

of   I   in   the   ligand  exchange  process  of  Pd(EtIm)2Cl2.  Dissociative   ligand  exchange   rates  are  

too  slow  to  give  cross-­‐peaks  in  the  spectrum.    

   

NN Pd NN

ClN

N

N NNN Pd N

ClN

Cl

L C I

+ Cl)

Chapter 4

76

Page 84: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  3:  MALDI-­‐TOF  spectrum  of   [Pd(EtIm)3Cl]Cl  taken  in  CHCl3,  1mg/ml.   Inset:  chemical  structure  

of    [Pd(EtIm)3Cl]Cl  and  its  simulated  mass  spectrum.    

2D  EXSY  spectra  

The   mechanism   of   exchange   between   free   and   bound   Im   was   investigated   using   2D  

exchange   spectroscopy   (EXSY18,19).   To  monitor   the   ligand  exchange,   complex  2  was  mixed  

with  free  ligand  in  deuterated  chloroform  (CDCl3).  2D  EXSY  spectra  were  taken  in  CDCl3  at  RT  

with   several   different   mixing   times   (tmix).   Spectra   indicate   that   free   Im   (which   is   not  

coordinated   to   Pd   center,   abbreviated   as   L)   exchanges  with   Im   coordinated   to   Pd   on   the  

time  scale  of  typical  tmix  of  100-­‐500  ms.  When  tmix  was  lower  than  200  ms,  spectra  showed  

cross  peaks   for  exchange  of   L  with   intermediate   [Pd(EtIm)3Cl]Cl   (I)  and  for  Pd(EtIm)2Cl2  (C)    

with   I.   However,   only   at   higher   mixing   times   (>200   ms)   cross   peaks   for   L-­‐C   exchange  

becomes  visible  (Figure  4).    

 

 

 

 

 

 

 

Figure  4:  2D  EXSY  spectra  of  0.09  M  solution  of  N-­‐ethyl  Imidazole  and  Pd(EtIm)2Cl2  in  CDCl3  at  tmix  =  

200  ms  and  tmix  =  500  ms.    Circles  were  used  to  highlight  cross  peaks  appear  when  tmix  is  high  (≥  300  

ms).  

In  order  to  investigate  exchange  rate  constants,  ratios  of  cross  peaks  to  diagonal  peaks  were  

calculated   using   absolute   peak   volumes   at   each   tmix.     C-­‐L   crosspeaks   are   absent   at   low  

mixing  times  and  only  appear  at  mixing  times  above  200  ms.  The  absence  of  direct  exchange  

between  C  and  L   in  the  presence  of  C-­‐I  and  L-­‐I  cross  peaks  demonstrates  the  intermediacy  

of   I   in   the   ligand  exchange  process  of  Pd(EtIm)2Cl2.  Dissociative   ligand  exchange   rates  are  

too  slow  to  give  cross-­‐peaks  in  the  spectrum.    

   

NN Pd NN

ClN

N

N NNN Pd N

ClN

Cl

L C I

+ Cl)

Determination of ligand exchange dynamics

77

4

Page 85: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  5:  Graph  of  cross-­‐peak  intensities  relative  to  diagonal  intensities  vs  tmix  for  exchange  between  

C-­‐I,    L-­‐I  (left)  and  L-­‐C  (right).  

On  the  other  hand,  for  short  mixing  times,  the  intensity  of  C-­‐I  and  L-­‐I  cross  peaks  increases  

linearly   with   tmix   indicating   that   these   are   direct   exchange   processes.   Rate   constants   for  

these  processes  were  calculated  from  the  slope  of  the  transfer  function  ф  vs  tmix  (Figure  6)  

as  described  by  Perrin  and  Dwyer.18  The  (pseudo)  first  order  rate  constants  were  found  to  

be  kL-­‐I  =  3.7  s–1  (r2  =  0.999)  and  kC-­‐I  =  3.2  s–1  (r2  =  0.999).    

 

Figure   6:   In   order   to   determine   rate  

constants  for  two  site  exchange  between  

L-­‐I   and   C-­‐I   by   the   equation  𝑘𝑘 = !!!"#

×ф,  

the  transfer  functions  (ф)  were  calculated  

using   the   expression:ф = ln !!!!!!

 where                            

𝑟𝑟 =   !!!!!!!!!"!!!"

.   Rate   constants   were  

obtained  from  the  slope  of  a  plot  of  ф  vs  

tmix.  

 

 

 

Sonication  induced  ligand  exchange  

Sonication   induced   ligand   exchange   was   studied   in   toluene   using   polymer   functionalized  

complex   Pd(Im-­‐pTHF18)2Cl2   and   a   small   molecule   model   complex   at   a   1:100   mole   ratio.  

Dodecyl   Imidazole   complex   Pd(DodecIm)2Cl2   was   used   because   of   its   high   solubility   in  

toluene.  MWs  of  the  complexes  are  sufficiently  different  to  allow  monitoring  of  the  ligand  

exchange  with  GPC.  0.1  mM  solutions  of  polymers  were  sonicated  using  a  sonication  probe  

and  solutions  were  saturated  with  CH4  by  bubbling  the  gas  through  the  solution  starting  15  

minutes  prior   to  sonication.  The  temperature   inside   the  sonication  vessel  was  constant  at  

25  oC  to  prevent  thermal   reactions.  This  was  done  by  circulation  of  water  at  2oC  between  

the  walls  of  the  double-­‐jacketed  sonication  vessel.        

Mechanochemically   induced   ligand   exchange   resulted   in   fragmentation   of   Pd(Im-­‐

pTHF18)2Cl2.  Emergence  of  a  new  peak  in  GPC  at  a  MW  of  18  kDa  was  monitored  over  time.  

This   new   peak   is   consistent   with   the   formation   of   product   with   two   different   imidazole  

ligands   coordinated   to   the   same   metal   center.   17%   of   the   initial   Pd(Im-­‐pTHF18k)2Cl2    

experienced  ligand  exchange  after  2h  of  continuous  sonication  as  monitored  by  GPC.  On  the  

other  hand,  without  sonication,  the  MW  distribution  of  the  mixture  did  not  change  over  the  

course  of  2  h  at  RT,  indicating  that  thermal  ligand  exchange  between  the  complexes  is  slow.      

 

Figure   7:  Ultrasound   induced   ligand   exchange   between   Pd(Im-­‐pTHF18k)2Cl2   (0.1   mM)   and   Pd(Im-­‐

Dodec)2Cl2  without   sonication   (left)   and  with   sonication   (right).  Molar   ratio   between  polymer   and  

small  molecule  is  1:100.  

Chapter 4

78

Page 86: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  5:  Graph  of  cross-­‐peak  intensities  relative  to  diagonal  intensities  vs  tmix  for  exchange  between  

C-­‐I,    L-­‐I  (left)  and  L-­‐C  (right).  

On  the  other  hand,  for  short  mixing  times,  the  intensity  of  C-­‐I  and  L-­‐I  cross  peaks  increases  

linearly   with   tmix   indicating   that   these   are   direct   exchange   processes.   Rate   constants   for  

these  processes  were  calculated  from  the  slope  of  the  transfer  function  ф  vs  tmix  (Figure  6)  

as  described  by  Perrin  and  Dwyer.18  The  (pseudo)  first  order  rate  constants  were  found  to  

be  kL-­‐I  =  3.7  s–1  (r2  =  0.999)  and  kC-­‐I  =  3.2  s–1  (r2  =  0.999).    

 

Figure   6:   In   order   to   determine   rate  

constants  for  two  site  exchange  between  

L-­‐I   and   C-­‐I   by   the   equation  𝑘𝑘 = !!!"#

×ф,  

the  transfer  functions  (ф)  were  calculated  

using   the   expression:ф = ln !!!!!!

 where                            

𝑟𝑟 =   !!!!!!!!!"!!!"

.   Rate   constants   were  

obtained  from  the  slope  of  a  plot  of  ф  vs  

tmix.  

 

 

 

Sonication  induced  ligand  exchange  

Sonication   induced   ligand   exchange   was   studied   in   toluene   using   polymer   functionalized  

complex   Pd(Im-­‐pTHF18)2Cl2   and   a   small   molecule   model   complex   at   a   1:100   mole   ratio.  

Dodecyl   Imidazole   complex   Pd(DodecIm)2Cl2   was   used   because   of   its   high   solubility   in  

toluene.  MWs  of  the  complexes  are  sufficiently  different  to  allow  monitoring  of  the  ligand  

exchange  with  GPC.  0.1  mM  solutions  of  polymers  were  sonicated  using  a  sonication  probe  

and  solutions  were  saturated  with  CH4  by  bubbling  the  gas  through  the  solution  starting  15  

minutes  prior   to  sonication.  The  temperature   inside   the  sonication  vessel  was  constant  at  

25  oC  to  prevent  thermal   reactions.  This  was  done  by  circulation  of  water  at  2oC  between  

the  walls  of  the  double-­‐jacketed  sonication  vessel.        

Mechanochemically   induced   ligand   exchange   resulted   in   fragmentation   of   Pd(Im-­‐

pTHF18)2Cl2.  Emergence  of  a  new  peak  in  GPC  at  a  MW  of  18  kDa  was  monitored  over  time.  

This   new   peak   is   consistent   with   the   formation   of   product   with   two   different   imidazole  

ligands   coordinated   to   the   same   metal   center.   17%   of   the   initial   Pd(Im-­‐pTHF18k)2Cl2    

experienced  ligand  exchange  after  2h  of  continuous  sonication  as  monitored  by  GPC.  On  the  

other  hand,  without  sonication,  the  MW  distribution  of  the  mixture  did  not  change  over  the  

course  of  2  h  at  RT,  indicating  that  thermal  ligand  exchange  between  the  complexes  is  slow.      

 

Figure   7:  Ultrasound   induced   ligand   exchange   between   Pd(Im-­‐pTHF18k)2Cl2   (0.1   mM)   and   Pd(Im-­‐

Dodec)2Cl2  without   sonication   (left)   and  with   sonication   (right).  Molar   ratio   between  polymer   and  

small  molecule  is  1:100.  

Determination of ligand exchange dynamics

79

4

Page 87: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   8:   Ultrasound   induced   ligand   exchange   between   Pd(Im-­‐pTHF35k)2Cl2   (0.1   mM)  and   Pd(Im-­‐

Dodec)2Cl2.  Molar  ratio  between  polymer  and  small  molecule  is  1:100.  

Pd(Im-­‐pTHF35k)2Cl2  polymer  was  also  tested  in  sonication  induced  ligand  exchange.  After  1h  

of  continuous  sonication,  40%  of  complexes  were  exchanged,  a  significantly  higher  fraction  

than  for  the  36k  complex.  The  molecular  weight  dependence  of  the  rate  demonstrates  that  

bond  scission  is  a  mechano-­‐chemical  process  and  not  the  result  of  local  heating  or  reaction  

with  solvent  decomposition  products.  

 

 

 

 

 

 

 

Rate  of  sonication  induced  chain  scission  

 

Figure   9:  Concentration  change  of  Pd(Im-­‐pTHF)2Cl2  complexes  during  sonication   in   the  presence  of  

Pd(DodecIm)2Cl2.  Red  lines  represent  linear  fitting.    

 The   rate   of   sonication   induced   chain   scission  was   determined   as   described   in   Chapter   2.  

Samples  were  taken  during  sonication  and  submitted  to  the  GPC.  Since  area  under  the  GPC  

trace  is  proportional  to  the  weight  fraction  of  the  polymer,  change  in  polymer  concentration  

could  be  found  by  deconvolution  of  the  GPC  traces.  Regardless  of  initial  molecular  weights  

of  polymer  complex,  the  concentration  of  initial  complex  decreased  linearly  with  sonication  

time   (Figure  9),   suggesting   that   the  consumption  of   complex   is  a   zeroth  order  process,  of  

which  the  rate  does  not  depend  on  the  concentration  of  complex.  This  behavior  is   in  stark  

contrast   to   the   experiments   described   in   previous   chapters,   where   the   concentration   of  

polymer   decayed   exponentially   during   sonication,   from   which   it   was   concluded   that  

mechanochemical  chain  scission  follows  first  order  reaction  kinetics.    

Thus,   in   the   current   experiments,   the   rate-­‐determining   step   appears   not   to   be   the  

mechanochemical  dissociation  of   the  polymer   ligand,  but  a   later   step.   In  order   to  analyze  

the  process  in  more  detail,  rates  of  separate  steps  were  calculated,  and  the  overall  process  

was  simulated.  

Sonication  induced  ligand  exchange  between  Pd(Im-­‐pTHF)2Cl2  and  Pd(Im-­‐Dodec)2Cl2  involves  

three   main   steps:   i)   dissociation   of   polymeric   ligand   under   stress,   ii)   association   of   free  

Chapter 4

80

Page 88: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   8:   Ultrasound   induced   ligand   exchange   between   Pd(Im-­‐pTHF35k)2Cl2   (0.1   mM)  and   Pd(Im-­‐

Dodec)2Cl2.  Molar  ratio  between  polymer  and  small  molecule  is  1:100.  

Pd(Im-­‐pTHF35k)2Cl2  polymer  was  also  tested  in  sonication  induced  ligand  exchange.  After  1h  

of  continuous  sonication,  40%  of  complexes  were  exchanged,  a  significantly  higher  fraction  

than  for  the  36k  complex.  The  molecular  weight  dependence  of  the  rate  demonstrates  that  

bond  scission  is  a  mechano-­‐chemical  process  and  not  the  result  of  local  heating  or  reaction  

with  solvent  decomposition  products.  

 

 

 

 

 

 

 

Rate  of  sonication  induced  chain  scission  

 

Figure   9:  Concentration  change  of  Pd(Im-­‐pTHF)2Cl2  complexes  during  sonication   in   the  presence  of  

Pd(DodecIm)2Cl2.  Red  lines  represent  linear  fitting.    

 The   rate   of   sonication   induced   chain   scission  was   determined   as   described   in   Chapter   2.  

Samples  were  taken  during  sonication  and  submitted  to  the  GPC.  Since  area  under  the  GPC  

trace  is  proportional  to  the  weight  fraction  of  the  polymer,  change  in  polymer  concentration  

could  be  found  by  deconvolution  of  the  GPC  traces.  Regardless  of  initial  molecular  weights  

of  polymer  complex,  the  concentration  of  initial  complex  decreased  linearly  with  sonication  

time   (Figure  9),   suggesting   that   the  consumption  of   complex   is  a   zeroth  order  process,  of  

which  the  rate  does  not  depend  on  the  concentration  of  complex.  This  behavior  is   in  stark  

contrast   to   the   experiments   described   in   previous   chapters,   where   the   concentration   of  

polymer   decayed   exponentially   during   sonication,   from   which   it   was   concluded   that  

mechanochemical  chain  scission  follows  first  order  reaction  kinetics.    

Thus,   in   the   current   experiments,   the   rate-­‐determining   step   appears   not   to   be   the  

mechanochemical  dissociation  of   the  polymer   ligand,  but  a   later   step.   In  order   to  analyze  

the  process  in  more  detail,  rates  of  separate  steps  were  calculated,  and  the  overall  process  

was  simulated.  

Sonication  induced  ligand  exchange  between  Pd(Im-­‐pTHF)2Cl2  and  Pd(Im-­‐Dodec)2Cl2  involves  

three   main   steps:   i)   dissociation   of   polymeric   ligand   under   stress,   ii)   association   of   free  

Determination of ligand exchange dynamics

81

4

Page 89: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

ligand  to  Pd(Im-­‐Dodec)2Cl2  to  form  the  tris-­‐imidazole  intermediate  as  mentioned  above,  and  

iii)  dissociation  of  one  of  the  three  ligands.    

Relevant  reactions  during  sonication  are  (Scheme  3);  

𝑃𝑃𝑃𝑃𝑃𝑃   ↔ 𝑃𝑃𝑃𝑃 + 𝑃𝑃                                                                         1  

𝑃𝑃𝑃𝑃  + 𝐿𝐿   → 𝑃𝑃𝑃𝑃𝑃𝑃                                                                         2  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝑃𝑃 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     3  

𝑃𝑃𝑃𝑃𝑃𝑃 + 𝐿𝐿 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     4  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝐿𝐿 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     5  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝑃𝑃 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     6  

𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑃𝑃 ↔ 𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃                                                                     7  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝐿𝐿 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                         8  

𝑃𝑃:𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝  𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙  𝐼𝐼𝐼𝐼 − 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝,𝑀𝑀:𝑀𝑀𝑀𝑀𝑀𝑀𝑀𝑀𝑀𝑀  𝑐𝑐𝑐𝑐𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛   𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃! ,  

 𝐿𝐿: 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠  𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙  (𝐼𝐼𝐼𝐼 − 𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷)  

 

 

 

Scheme   3:   Plausible  mechanism   for  ultrasound   induced   ligand  exchange  between  Pd(Im-­‐pTHF)2Cl2  

and  Pd(Im-­‐Dodec)2Cl2.    

 

 

NNPd N NCl

N

NC₁₂H₂₅

Cl*

NN Pd N NCl

ClO O N N C₁₂H₂₅ NN Pd N N

Cl

ClO C₁₂H₂₅+

n n n

On

On

NN Pd N NCl

ClO O +

n nNN Pd N N

Cl

ClC₁₂H₂₅ C₁₂H₂₅

N N O +n

NN Pd N NCl

ClC₁₂H₂₅ C₁₂H₂₅

NNPd N NCl

N

N

Cl*

C₁₂H₂₅ C₁₂H₂₅

O

n

NN PdCl

ClO

nN N C₁₂H₂₅

NN Pd N NCl

ClO C₁₂H₂₅+

PMP LML

P

L

PMLL

PML

PMPL

PM PMLL

Chapter 4

82

Page 90: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

ligand  to  Pd(Im-­‐Dodec)2Cl2  to  form  the  tris-­‐imidazole  intermediate  as  mentioned  above,  and  

iii)  dissociation  of  one  of  the  three  ligands.    

Relevant  reactions  during  sonication  are  (Scheme  3);  

𝑃𝑃𝑃𝑃𝑃𝑃   ↔ 𝑃𝑃𝑃𝑃 + 𝑃𝑃                                                                         1  

𝑃𝑃𝑃𝑃  + 𝐿𝐿   → 𝑃𝑃𝑃𝑃𝑃𝑃                                                                         2  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝑃𝑃 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     3  

𝑃𝑃𝑃𝑃𝑃𝑃 + 𝐿𝐿 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     4  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝐿𝐿 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     5  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝑃𝑃 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                     6  

𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑃𝑃 ↔ 𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃                                                                     7  

𝐿𝐿𝐿𝐿𝐿𝐿 + 𝐿𝐿 ↔ 𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿                                                                         8  

𝑃𝑃:𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝  𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙  𝐼𝐼𝐼𝐼 − 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝,𝑀𝑀:𝑀𝑀𝑀𝑀𝑀𝑀𝑀𝑀𝑀𝑀  𝑐𝑐𝑐𝑐𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛   𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃! ,  

 𝐿𝐿: 𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠  𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙  (𝐼𝐼𝐼𝐼 − 𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷)  

 

 

 

Scheme   3:   Plausible  mechanism   for  ultrasound   induced   ligand  exchange  between  Pd(Im-­‐pTHF)2Cl2  

and  Pd(Im-­‐Dodec)2Cl2.    

 

 

NNPd N NCl

N

NC₁₂H₂₅

Cl*

NN Pd N NCl

ClO O N N C₁₂H₂₅ NN Pd N N

Cl

ClO C₁₂H₂₅+

n n n

On

On

NN Pd N NCl

ClO O +

n nNN Pd N N

Cl

ClC₁₂H₂₅ C₁₂H₂₅

N N O +n

NN Pd N NCl

ClC₁₂H₂₅ C₁₂H₂₅

NNPd N NCl

N

N

Cl*

C₁₂H₂₅ C₁₂H₂₅

O

n

NN PdCl

ClO

nN N C₁₂H₂₅

NN Pd N NCl

ClO C₁₂H₂₅+

PMP LML

P

L

PMLL

PML

PMPL

PM PMLL

Determination of ligand exchange dynamics

83

4

Page 91: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Relative  peak  areas  in  1H  NMR  spectrum  of  the  small  molecule  intermediate  (Figure  2)  give  a  

1:5:5  ratio  of  equilibrium  concentrations  of  I:C:L.  Since  the  initial  concentrations  of  C  and  L  

are  0.09  M,  

𝐿𝐿 !   = 0.09𝑀𝑀,   𝐶𝐶 !   = 0.09𝑀𝑀,  [𝐼𝐼] = [𝐶𝐶] 5  

𝐼𝐼   = 0.015  𝑀𝑀  𝑎𝑎𝑎𝑎𝑎𝑎 𝐶𝐶  , 𝐿𝐿   = 0.075  

the  association  constant  can  be  calculated  as  

𝐾𝐾!   =  [𝐼𝐼]

[𝐶𝐶]×[𝐸𝐸]  =0.0150.075! ≅ 2.7  𝑀𝑀!!  

Rate  constants  in  scission  experiments  can  be  written  as:  

 

Pd(Im-­‐pTHF-­‐35k)2Cl2   showed   40%   chain   scission   after   1h   sonication,   so   the   scission   rate  

constant  k1  should  be  higher  than  10-­‐4  s-­‐1.  The  reverse  reaction  and  reaction  (2)  are  assumed  

to  be  diffusion  controlled.  We  used  the  diffusion  controlled  association  constant  calculated  

previously  for  UPy  dimers  in  toluene  (k-­‐1  =  108  M-­‐1s-­‐1)20  

𝐾𝐾 =𝑘𝑘!!

𝑘𝑘! + 𝑘𝑘!=  

𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 𝑃𝑃  

where  k2  is  sum  of  rates  for  reactions  that  involves  polymeric  ligand,  P.  Those  reactions  are  

(3),  (6)  and  (7).    

The   steps   in   the   exchange   process   were   modeled   in   the   metabolic   modeling   program  

GEPASI21   in   order   to   verify   that   the   combination   of   elementary   reaction   rate   constants  

discussed  above  results  in  a  linear  decrease  of  the  polymer  concentration  during  sonication.  

When  the  rate  constants  and  equilibrium  concentrations  as  calculated  above  were  used  in  

simulation,   the   polymer   concentration   indeed   decreased   linearly   in   time   (Figure   10).   The  

simulations   showed   that   this   is   caused  by   the   slow  kinetics  of   the   ligand  association   step  

(i.e.  third  ligand  coordinates  to  Pd(NIm)2  complex,  reaction  (3)  in  Scheme  3).    

P M P P M +k₁

k&₁

P

 

Figure   10:   Simulated   plot   for   change   in   polymer   concentration   during   sonication   vs.   time   using  

GEPASI  compared  to  the  data  that  was  obtained  experimentally.    

Conclusions  

The  ligand  exchange  mechanism  for  imidazole  Palladium  complexes  has  been  investigated  in  

the   presence   of   excess   free   imidazole   ligand   in   CHCl3   as   an   associative   process.   When  

Pd(EtIm)2Cl2   and   EtIm  were  mixed,   a   tris   imidazolyl   intermediate   complex   [Pd(EtIm)3Cl]Cl    

was  identified  by  NMR  and  MALDI-­‐TOF.  It  has  been  shown  with  2D  EXSY  that  there  is  direct  

magnetization   transfer   between   Pd(EtIm)2Cl2   and   the   intermediate.   However,   exchange  

between   Pd(EtIm)2Cl2   and   EtIm   is   indirect,   and   proceeds   via   the   stable   intermediate  

[Pd(EtIm)3Cl]Cl.    

Complexes   with  monotelechelic   polymeric   ligands   Pd(Im-­‐pTHF)2Cl2   were   sonicated   in   the  

presence   of   Pd(Im-­‐Dodec)2Cl2   and   GPC   traces   showed   that   during   sonication   polymeric  

ligands  dissociate  from  the  central  metal.      

Previously,   the   rate   of   sonication   induced   chain   scission   in   coordination   complexes   was  

investigated   for  NHC  bearing  Pd(NHC-­‐pTHF)2Cl2   in   the  presence  of   trapping   agents   (AcOH  

and  MeCN).  These  trapping  agents  coordinate  to  the  metal  center  and  protonate  the  free  

carbene  released  when  the  bond  between  metal  and  ligand  is  broken.  Although  it  has  been  

established  in  previous  chapters  that  the  bond  scission  observed  during  sonication  of  these  

complexes   is  mechanically   induced,   it   is   still   arguable   that   trapping   agents   play   a   role   in  

thermally   induced   chain   scission   by   destabilizing   the   coordination   bond.   However,   the  

ultrasound   induced   ligand   exchange   presented   in   this   chapter   occurs   in   the   absence   of  

Chapter 4

84

Page 92: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Relative  peak  areas  in  1H  NMR  spectrum  of  the  small  molecule  intermediate  (Figure  2)  give  a  

1:5:5  ratio  of  equilibrium  concentrations  of  I:C:L.  Since  the  initial  concentrations  of  C  and  L  

are  0.09  M,  

𝐿𝐿 !   = 0.09𝑀𝑀,   𝐶𝐶 !   = 0.09𝑀𝑀,  [𝐼𝐼] = [𝐶𝐶] 5  

𝐼𝐼   = 0.015  𝑀𝑀  𝑎𝑎𝑎𝑎𝑎𝑎 𝐶𝐶  , 𝐿𝐿   = 0.075  

the  association  constant  can  be  calculated  as  

𝐾𝐾!   =  [𝐼𝐼]

[𝐶𝐶]×[𝐸𝐸]  =0.0150.075! ≅ 2.7  𝑀𝑀!!  

Rate  constants  in  scission  experiments  can  be  written  as:  

 

Pd(Im-­‐pTHF-­‐35k)2Cl2   showed   40%   chain   scission   after   1h   sonication,   so   the   scission   rate  

constant  k1  should  be  higher  than  10-­‐4  s-­‐1.  The  reverse  reaction  and  reaction  (2)  are  assumed  

to  be  diffusion  controlled.  We  used  the  diffusion  controlled  association  constant  calculated  

previously  for  UPy  dimers  in  toluene  (k-­‐1  =  108  M-­‐1s-­‐1)20  

𝐾𝐾 =𝑘𝑘!!

𝑘𝑘! + 𝑘𝑘!=  

𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 𝑃𝑃  

where  k2  is  sum  of  rates  for  reactions  that  involves  polymeric  ligand,  P.  Those  reactions  are  

(3),  (6)  and  (7).    

The   steps   in   the   exchange   process   were   modeled   in   the   metabolic   modeling   program  

GEPASI21   in   order   to   verify   that   the   combination   of   elementary   reaction   rate   constants  

discussed  above  results  in  a  linear  decrease  of  the  polymer  concentration  during  sonication.  

When  the  rate  constants  and  equilibrium  concentrations  as  calculated  above  were  used  in  

simulation,   the   polymer   concentration   indeed   decreased   linearly   in   time   (Figure   10).   The  

simulations   showed   that   this   is   caused  by   the   slow  kinetics  of   the   ligand  association   step  

(i.e.  third  ligand  coordinates  to  Pd(NIm)2  complex,  reaction  (3)  in  Scheme  3).    

P M P P M +k₁

k&₁

P

 

Figure   10:   Simulated   plot   for   change   in   polymer   concentration   during   sonication   vs.   time   using  

GEPASI  compared  to  the  data  that  was  obtained  experimentally.    

Conclusions  

The  ligand  exchange  mechanism  for  imidazole  Palladium  complexes  has  been  investigated  in  

the   presence   of   excess   free   imidazole   ligand   in   CHCl3   as   an   associative   process.   When  

Pd(EtIm)2Cl2   and   EtIm  were  mixed,   a   tris   imidazolyl   intermediate   complex   [Pd(EtIm)3Cl]Cl    

was  identified  by  NMR  and  MALDI-­‐TOF.  It  has  been  shown  with  2D  EXSY  that  there  is  direct  

magnetization   transfer   between   Pd(EtIm)2Cl2   and   the   intermediate.   However,   exchange  

between   Pd(EtIm)2Cl2   and   EtIm   is   indirect,   and   proceeds   via   the   stable   intermediate  

[Pd(EtIm)3Cl]Cl.    

Complexes   with  monotelechelic   polymeric   ligands   Pd(Im-­‐pTHF)2Cl2   were   sonicated   in   the  

presence   of   Pd(Im-­‐Dodec)2Cl2   and   GPC   traces   showed   that   during   sonication   polymeric  

ligands  dissociate  from  the  central  metal.      

Previously,   the   rate   of   sonication   induced   chain   scission   in   coordination   complexes   was  

investigated   for  NHC  bearing  Pd(NHC-­‐pTHF)2Cl2   in   the  presence  of   trapping   agents   (AcOH  

and  MeCN).  These  trapping  agents  coordinate  to  the  metal  center  and  protonate  the  free  

carbene  released  when  the  bond  between  metal  and  ligand  is  broken.  Although  it  has  been  

established  in  previous  chapters  that  the  bond  scission  observed  during  sonication  of  these  

complexes   is  mechanically   induced,   it   is   still   arguable   that   trapping   agents   play   a   role   in  

thermally   induced   chain   scission   by   destabilizing   the   coordination   bond.   However,   the  

ultrasound   induced   ligand   exchange   presented   in   this   chapter   occurs   in   the   absence   of  

Determination of ligand exchange dynamics

85

4

Page 93: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

trapping   agents,   and   thermal   processes   would   lead   to   the   formation   of   an   equilibrium  

mixture   of   starting   material   and   product   Pd[(Im-­‐pTHF)(Im-­‐Dodec)]Cl2.   The   selective  

formation   of   product   therefore   further   proves   that   the   scission   is  mechanochemical.   The  

direction  of  the  reaction  is  governed  by  the  fact  that  the  coordination  bond  in  the  product  

Pd[(Im-­‐pTHF)(Im-­‐Dodec)]Cl2  is  not    responsive  to  mechanical  stress  because  it  is  at  the  end  

of  the  polymer  chain,  where  solvodynamic  forces  are  too  low  to  cause  scission.  

Experimental  

General  

All   chemicals   were   purchased   from   commercial   sources   used   without   further   purification   unless  

specified   otherwise.   Dry   tetrahydrofuran   (THF,   HPLC   grade)   and   dichloromethane   (DCM)   were  

degassed  with  argon  and  purified  by  passage  through  activated  alumina  solvent  column  prior  to  use.  

Sodium  hydride  was  60%  dispersion  in  mineral  oil  as  obtained  from  commercial  source  washed  with  

n-­‐hexane   under   Ar   in   a   Schlenk   flask   prior   to   use.   A   Varian   400MR   or   a   Varian   Mercury   400  

spectrometer   was   used   to   record   1H   NMR   (400   MHz)   Chemical   shifts   are   reported   in   ppm   and  

referenced   to   chemical   shifts   for   tetramethylsilane   or   residual   solvents.   2D   EXSY   spectra   were  

recorded   on   Varian   Inova   500   spectrometer   with   NOESY   pulse   sequence.   Gel   permeation  

chromatography   (GPC)  was  performed  on  a  Shimadzu  LC10-­‐AD,  using  Polymer  Laboratories  PL  Gel  

5μm  MIXED-­‐C  and  MIXEDD  columns     (linear   range  of  MW:  200–2000000  g/mol),   a  Shimadzu  SPD-­‐

M10A  UV-­‐vis    detector  at  254  nm  and  RID-­‐10A  refractive  index  detector,  and  THF  as  eluent  at  a  flow  

rate  of  1  mL/min  (20  °C).  Polystyrene  standards  were  used  for  calibration.        

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   reactor   with   a   volume   of   10  mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   The   temperature   in   the   reactor   was  maintained  with   a   Lauda   E300   cooling  

bath  and  measured  using  a  0.5  mm  diameter  thermocouple.  Solutions  were  sonicated  continuously,  

temperature  of  the  solution  were  checked  by  thermocouple  was  constant  at  25oC  after  the  thermal  

equilibrium  was  achieved   in  the  first  3-­‐5  mins.  During  sonication  saturation  gas   (CH4)  was  bubbled  

through  solution  via  teflon  tubing.  Aliquots  of  100  µL  were  taken  at  different  time  intervals.  Toluene  

was   removed  under   reduced  pressure   and   residues  were   dissolved   in   THF   and   submitted   to  GPC.  

Results   were   analyzed   and   concentration   change   during   sonication   was   determined   by   double  

Gaussian  de-­‐convolution  method.    

Synthesis  of  α-­‐(N-­‐imidazole)-­‐ω-­‐methoxy  poly(tetrahydrofuran)  

Polymer   ligands   (Im-­‐pTHF)   were   synthesized   via   cationic   ring-­‐opening   polymerization   of  

tetrahydrofuran   (THF).22   THF   (100  mL)   and  DTBP   (200   µL,   0.92  mmol)  were   added  methyl   triflate  

(100  µL,  0.91  mmol)  inside  a  Schlenk  round-­‐bottom  flask  under  Ar  to  initiate  the  polymerization.  End  

capper  imidazolyl  was  prepared  in  a  separate  round  bottom  flask.  Imidazole  (306  mg,  4.5  mmol)  was  

added  in  portions  onto  NaH  (400  mg,  10  mmol)  dispersion  in  dry  THF  (10  ml)  under  Ar.  After  stirring  

for  defined  time  as  described  in  chapter  2,  the  polymerization  was  terminated  by  adding  imidazolyl  

solution  in  one  portion  to  polymerization  flask.  After  20  mins,  10  ml  MeOH  was  added,  the  solution  

was  diluted  to  app  ¼  of  its  initial  volume  under  reduced  pressure  and  precipitated  in  water  (400  mL)  

overnight   at   ambient   temperature.   White   polymer   was   washed   with   water,   dissolved   in   diethyl  

ether  (200  mL),  dried  over  MgSO4  and  precipitated  overnight  at  –30  °C,  white  powder  was  filtered  

washed  with  cold  Et2O  and  yielded  ligands  as  white  powder.  In  order  to  remove  traces  of  solvents,  

ligands  were   left  under  vacuum  at  ambient  temperature  overnight  prior  to  use.  Molecular  weights  

were  determined  by  MALDI-­‐TOF  as  18  kDa  and  35  kDa  depending  on   the  polymerization   time   (2h  

and  3.5h  respectively).  1H  NMR  Im-­‐pTHF18k  [CD2Cl2,  400  MHz]:  7.84  ppm  (s,   Im),  7.17  ppm  (s,   Im),  

7.06  ppm  (s,  Im),  4.08  ppm  (t,  N-­‐CH2,  J:  8Hz),  3.0-­‐3.6  ppm  (br  O-­‐CH2-­‐),  1.3-­‐2.2  ppm  (br,  OCH2-­‐CH2-­‐).  

Synthesis  of  Pd(II)–imidazole  polymer  complexes  Pd(Im-­‐pTHF)2Cl2.  

Im-­‐pTHF  (400  mg)  was  dissolved  in  DCM  (10  ml)  and  added  Pd(MeCN)2Cl2  (0.5  eq.).  Ligand  exchange  

yielded  desired  polymer  attached  mechanophores  after  2h.  Solvent  was  evaporated  under  reduced  

pressure.   1H  NMR  Pd(Im-­‐pTHF18k)2Cl2   [CD2Cl2,   400  MHz]:  8.03  ppm   (s,   Im),   7.38  ppm   (s,   Im),   6.67  

ppm  (s,  Im),  3.98  ppm  (t,  N-­‐CH2,  J:  8Hz),  3.0-­‐3.6  ppm  (br  O-­‐CH2-­‐),  1.3-­‐2.2  ppm  (br,  OCH2-­‐CH2-­‐).  

 

 

 

 

 

 

 

 

Chapter 4

86

Page 94: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

trapping   agents,   and   thermal   processes   would   lead   to   the   formation   of   an   equilibrium  

mixture   of   starting   material   and   product   Pd[(Im-­‐pTHF)(Im-­‐Dodec)]Cl2.   The   selective  

formation   of   product   therefore   further   proves   that   the   scission   is  mechanochemical.   The  

direction  of  the  reaction  is  governed  by  the  fact  that  the  coordination  bond  in  the  product  

Pd[(Im-­‐pTHF)(Im-­‐Dodec)]Cl2  is  not    responsive  to  mechanical  stress  because  it  is  at  the  end  

of  the  polymer  chain,  where  solvodynamic  forces  are  too  low  to  cause  scission.  

Experimental  

General  

All   chemicals   were   purchased   from   commercial   sources   used   without   further   purification   unless  

specified   otherwise.   Dry   tetrahydrofuran   (THF,   HPLC   grade)   and   dichloromethane   (DCM)   were  

degassed  with  argon  and  purified  by  passage  through  activated  alumina  solvent  column  prior  to  use.  

Sodium  hydride  was  60%  dispersion  in  mineral  oil  as  obtained  from  commercial  source  washed  with  

n-­‐hexane   under   Ar   in   a   Schlenk   flask   prior   to   use.   A   Varian   400MR   or   a   Varian   Mercury   400  

spectrometer   was   used   to   record   1H   NMR   (400   MHz)   Chemical   shifts   are   reported   in   ppm   and  

referenced   to   chemical   shifts   for   tetramethylsilane   or   residual   solvents.   2D   EXSY   spectra   were  

recorded   on   Varian   Inova   500   spectrometer   with   NOESY   pulse   sequence.   Gel   permeation  

chromatography   (GPC)  was  performed  on  a  Shimadzu  LC10-­‐AD,  using  Polymer  Laboratories  PL  Gel  

5μm  MIXED-­‐C  and  MIXEDD  columns     (linear   range  of  MW:  200–2000000  g/mol),   a  Shimadzu  SPD-­‐

M10A  UV-­‐vis    detector  at  254  nm  and  RID-­‐10A  refractive  index  detector,  and  THF  as  eluent  at  a  flow  

rate  of  1  mL/min  (20  °C).  Polystyrene  standards  were  used  for  calibration.        

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   reactor   with   a   volume   of   10  mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   The   temperature   in   the   reactor   was  maintained  with   a   Lauda   E300   cooling  

bath  and  measured  using  a  0.5  mm  diameter  thermocouple.  Solutions  were  sonicated  continuously,  

temperature  of  the  solution  were  checked  by  thermocouple  was  constant  at  25oC  after  the  thermal  

equilibrium  was  achieved   in  the  first  3-­‐5  mins.  During  sonication  saturation  gas   (CH4)  was  bubbled  

through  solution  via  teflon  tubing.  Aliquots  of  100  µL  were  taken  at  different  time  intervals.  Toluene  

was   removed  under   reduced  pressure   and   residues  were   dissolved   in   THF   and   submitted   to  GPC.  

Results   were   analyzed   and   concentration   change   during   sonication   was   determined   by   double  

Gaussian  de-­‐convolution  method.    

Synthesis  of  α-­‐(N-­‐imidazole)-­‐ω-­‐methoxy  poly(tetrahydrofuran)  

Polymer   ligands   (Im-­‐pTHF)   were   synthesized   via   cationic   ring-­‐opening   polymerization   of  

tetrahydrofuran   (THF).22   THF   (100  mL)   and  DTBP   (200   µL,   0.92  mmol)  were   added  methyl   triflate  

(100  µL,  0.91  mmol)  inside  a  Schlenk  round-­‐bottom  flask  under  Ar  to  initiate  the  polymerization.  End  

capper  imidazolyl  was  prepared  in  a  separate  round  bottom  flask.  Imidazole  (306  mg,  4.5  mmol)  was  

added  in  portions  onto  NaH  (400  mg,  10  mmol)  dispersion  in  dry  THF  (10  ml)  under  Ar.  After  stirring  

for  defined  time  as  described  in  chapter  2,  the  polymerization  was  terminated  by  adding  imidazolyl  

solution  in  one  portion  to  polymerization  flask.  After  20  mins,  10  ml  MeOH  was  added,  the  solution  

was  diluted  to  app  ¼  of  its  initial  volume  under  reduced  pressure  and  precipitated  in  water  (400  mL)  

overnight   at   ambient   temperature.   White   polymer   was   washed   with   water,   dissolved   in   diethyl  

ether  (200  mL),  dried  over  MgSO4  and  precipitated  overnight  at  –30  °C,  white  powder  was  filtered  

washed  with  cold  Et2O  and  yielded  ligands  as  white  powder.  In  order  to  remove  traces  of  solvents,  

ligands  were   left  under  vacuum  at  ambient  temperature  overnight  prior  to  use.  Molecular  weights  

were  determined  by  MALDI-­‐TOF  as  18  kDa  and  35  kDa  depending  on   the  polymerization   time   (2h  

and  3.5h  respectively).  1H  NMR  Im-­‐pTHF18k  [CD2Cl2,  400  MHz]:  7.84  ppm  (s,   Im),  7.17  ppm  (s,   Im),  

7.06  ppm  (s,  Im),  4.08  ppm  (t,  N-­‐CH2,  J:  8Hz),  3.0-­‐3.6  ppm  (br  O-­‐CH2-­‐),  1.3-­‐2.2  ppm  (br,  OCH2-­‐CH2-­‐).  

Synthesis  of  Pd(II)–imidazole  polymer  complexes  Pd(Im-­‐pTHF)2Cl2.  

Im-­‐pTHF  (400  mg)  was  dissolved  in  DCM  (10  ml)  and  added  Pd(MeCN)2Cl2  (0.5  eq.).  Ligand  exchange  

yielded  desired  polymer  attached  mechanophores  after  2h.  Solvent  was  evaporated  under  reduced  

pressure.   1H  NMR  Pd(Im-­‐pTHF18k)2Cl2   [CD2Cl2,   400  MHz]:  8.03  ppm   (s,   Im),   7.38  ppm   (s,   Im),   6.67  

ppm  (s,  Im),  3.98  ppm  (t,  N-­‐CH2,  J:  8Hz),  3.0-­‐3.6  ppm  (br  O-­‐CH2-­‐),  1.3-­‐2.2  ppm  (br,  OCH2-­‐CH2-­‐).  

 

 

 

 

 

 

 

 

Determination of ligand exchange dynamics

87

4

Page 95: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

References  

(1)     Beck,  J.  B.;  Rowan,  S.  J.  J.  Am.  Chem.  Soc.  2003,  125  (46),  13922–13923.  

(2)     Dobrawa,  R.;  Würthner,  F.  J.  Polym.  Sci.  Part  Polym.  Chem.  2005,  43  (21),  4981–4995.  

(3)     Yount,  W.  C.;  Juwarker,  H.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2003,  125  (50),  15302–15303.  

(4)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2003,  No.  13,  1494–1495.  

(5)     Vermonden,  T.;  van  der  Gucht,  J.;  de  Waard,  P.;  Marcelis,  A.  T.  M.;  Besseling,  N.  A.  M.;  

Sudhölter,  E.  J.  R.;  Fleer,  G.  J.;  Cohen  Stuart,  M.  A.  Macromolecules  2003,  36  (19),  7035–7044.  

(6)     Michelsen,  U.;  Hunter,  C.  A.  Angew.  Chem.  Int.  Ed.  2000,  39  (4),  764–767.  

(7)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.-­‐Int.  Ed.  2004,  43  (34),  4460–4462.  

(8)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(9)     Suslick,  K.  S.;  Price,  G.  J.  Annu.  Rev.  Mater.  Sci.  1999,  29  (1),  295–326.  

(10)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(11)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(12)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(13)     Basolo,  F.;  Chatt,  J.;  Gray,  H.  B.;  Pearson,  R.  G.;  Shaw,  B.  L.  J.  Chem.  Soc.  Resumed  1961,  No.  0,  

2207–2215.  

(14)     Cramer,  R.  J.  Am.  Chem.  Soc.  1972,  94  (16),  5681–5685.  

(15)     Johnson,  L.  K.;  Killian,  C.  M.;  Brookhart,  M.  J.  Am.  Chem.  Soc.  1995,  117  (23),  6414–6415.  

(16)     Ghosh,  P.  K.;  Mandal,  H.  K.;  Mahapatra,  A.  Int.  J.  Chem.  Kinet.  2011,  43  (3),  130–140.  

(17)     Szulmanowicz,  M.  S.;  Zawartka,  W.;  Gniewek,  A.;  Trzeciak,  A.  M.  Inorganica  Chim.  Acta  2010,  

363  (15),  4346–4354.  

(18)     Perrin,  C.  L.;  Dwyer,  T.  J.  Chem.  Rev.  1990,  90  (6),  935–967.  

(19)     Dwyer,  T.  J.;  Norman,  J.  E.;  Jasien,  P.  G.  J.  Chem.  Educ.  1998,  75  (12),  1635.  

(20)     Söntjens,  S.  H.  M.;  Sijbesma,  R.  P.;  van  Genderen,  M.  H.  P.;  Meijer,  E.  W.  J.  Am.  Chem.  Soc.  

2000,  122  (31),  7487–7493.  

(21)     Mendes,  P.  Trends  Biochem.  Sci.  1997,  22  (9),  361–363.  

(22)     Madras,  G.;  Chung,  G.  Y.;  Smith,  J.  M.;  McCoy,  B.  J.  Ind.  Eng.  Chem.  Res.  1997,  36  (6),  2019–

2024.  

(23)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(24)     Odell,  J.  A.;  Keller,  A.  J.  Polym.  Sci.  Part  B  Polym.  Phys.  1986,  24  (9),  1889–1916.  

(25)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

 

Chapter 4

88

Page 96: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter 5Mechanochemically induced, directed ligand exchange in polymeric Pd(II) complexes

Mechanochemically induced ligand exchange of Pd(II) complexes was used to direct the for-

mation of heterocomplexes. Symmetric complexes with high and low molecular weight po-

lymer-attached ligands were mixed in solution and sonicated. When one of the complexes

has a molecular weight higher than the threshold (Mlim) for mechanochemical chain scission,

while the other is smaller, sonication leads to the directed formation of a heterocomplex

with two different ligands.

References  

(1)     Beck,  J.  B.;  Rowan,  S.  J.  J.  Am.  Chem.  Soc.  2003,  125  (46),  13922–13923.  

(2)     Dobrawa,  R.;  Würthner,  F.  J.  Polym.  Sci.  Part  Polym.  Chem.  2005,  43  (21),  4981–4995.  

(3)     Yount,  W.  C.;  Juwarker,  H.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2003,  125  (50),  15302–15303.  

(4)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2003,  No.  13,  1494–1495.  

(5)     Vermonden,  T.;  van  der  Gucht,  J.;  de  Waard,  P.;  Marcelis,  A.  T.  M.;  Besseling,  N.  A.  M.;  

Sudhölter,  E.  J.  R.;  Fleer,  G.  J.;  Cohen  Stuart,  M.  A.  Macromolecules  2003,  36  (19),  7035–7044.  

(6)     Michelsen,  U.;  Hunter,  C.  A.  Angew.  Chem.  Int.  Ed.  2000,  39  (4),  764–767.  

(7)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.-­‐Int.  Ed.  2004,  43  (34),  4460–4462.  

(8)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(9)     Suslick,  K.  S.;  Price,  G.  J.  Annu.  Rev.  Mater.  Sci.  1999,  29  (1),  295–326.  

(10)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(11)     Groote,  R.;  van  Haandel,  L.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2012,  50  (23),  

4929–4935.  

(12)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(13)     Basolo,  F.;  Chatt,  J.;  Gray,  H.  B.;  Pearson,  R.  G.;  Shaw,  B.  L.  J.  Chem.  Soc.  Resumed  1961,  No.  0,  

2207–2215.  

(14)     Cramer,  R.  J.  Am.  Chem.  Soc.  1972,  94  (16),  5681–5685.  

(15)     Johnson,  L.  K.;  Killian,  C.  M.;  Brookhart,  M.  J.  Am.  Chem.  Soc.  1995,  117  (23),  6414–6415.  

(16)     Ghosh,  P.  K.;  Mandal,  H.  K.;  Mahapatra,  A.  Int.  J.  Chem.  Kinet.  2011,  43  (3),  130–140.  

(17)     Szulmanowicz,  M.  S.;  Zawartka,  W.;  Gniewek,  A.;  Trzeciak,  A.  M.  Inorganica  Chim.  Acta  2010,  

363  (15),  4346–4354.  

(18)     Perrin,  C.  L.;  Dwyer,  T.  J.  Chem.  Rev.  1990,  90  (6),  935–967.  

(19)     Dwyer,  T.  J.;  Norman,  J.  E.;  Jasien,  P.  G.  J.  Chem.  Educ.  1998,  75  (12),  1635.  

(20)     Söntjens,  S.  H.  M.;  Sijbesma,  R.  P.;  van  Genderen,  M.  H.  P.;  Meijer,  E.  W.  J.  Am.  Chem.  Soc.  

2000,  122  (31),  7487–7493.  

(21)     Mendes,  P.  Trends  Biochem.  Sci.  1997,  22  (9),  361–363.  

(22)     Madras,  G.;  Chung,  G.  Y.;  Smith,  J.  M.;  McCoy,  B.  J.  Ind.  Eng.  Chem.  Res.  1997,  36  (6),  2019–

2024.  

(23)     Vijayalakshmi,  S.  P.;  Madras,  G.  Polym.  Degrad.  Stab.  2005,  90  (1),  116–122.  

(24)     Odell,  J.  A.;  Keller,  A.  J.  Polym.  Sci.  Part  B  Polym.  Phys.  1986,  24  (9),  1889–1916.  

(25)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

 

Page 97: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Coordination   polymers   (CPs)   have   been   used   as   mechanoresponsive   polymers   for  

mechanical   release   of   reactive   groups,   such   as   catalytically   active   sites,   luminescent   end  

groups  and  small  molecules.1–6  Our  research  group  has  shown  that  the  coordination  sphere  

of   transition-­‐metal  complexes   in   the  main  chain  of  CPs  can  be  manipulated   in  solution  by  

ultrasound.7  The   force   required   to  break  metal−ligand  bond   is  much   lower   than   the   force  

that  is  typically  required  to  break  covalent  bonds.8  Thus,  ultrasonication  of  CPs  enables  the  

specific  rupture  of  metal  ligand  coordination  bonds.  9  

Mechanical   chain   scission   in   reversible   bonds   drives   dynamic   systems   away   from  

equilibrium.  Paulusse  showed,  for  instance,  that  the  ring  chain  equilibrium  in  Pd-­‐phosphine  

coordination  polymers  is  shifted  reversibly  by  ultrasound.10  The  equilibrium  composition  of  

coordination  polymer  solutions  can  also  be  changed  by  ultrasound.  As  described  in  Chapter  

4,   Pd(Im)2Cl2   complexes   in   toluene   do   not   show   ligand   exchange   at   room   temperature.  

However,  when  free  imidazole  is  present,  it  replaces  one  of  the  coordinated  imidazoles  via  

an   associative   mechanism,   which   results   in   ligand   exchange.   This   offers   a   method   to  

selectively   change   the   composition   of   a   mixture   of   complexes   by   mechanochemical  

activation.  

 

Scheme  1:  PdCl2-­‐Imidazole  complexes  are  kinetically  stable  against   ligand  exchange  in  the  absence  

of  free  ligand  (top).  Sonication  dissociates  one  of  the  ligands  of  the  polymeric  complex  by  selectively  

breaking  the  metal  ligand  bond.  Then,  released  free  ligand  induces  associative  ligand  exchange.  

Provided  that  the  molecular  weight  (MW)  one  of  the  complexes   is  above  the   limiting  MW  

(Mlim),11,12  to  break  the  metal  ligand  bond  by  ultrasound,  ligands  can  be  exchanged  and  the  

equilibrium   will   shift   to   the   hetero-­‐complex.   Since   mechanical   force   on   the   weak  

coordinative   bond   increases   with   chain   length,   only   polymer   that   has   high   MW,   is  

mechanically   labile  and  breaks  upon  sonication.  Free   ligand  binds  to  other  complex   in   the  

NN Pd N NCl

ClO O

NN PdCl

ClO

nN N O NN Pd N N

Cl

ClC12H25 C12H25

NN Pd N NCl

ClO

nC12H25

+

NN Pd N NCl

ClC12H25 C12H25+ x

solution  and  results  in  a  solution  with  non-­‐equilibrium  composition  (Scheme  1).  In  principle,  

this   feature   of   mechanochemical   polymer   scission   allows   formation   of   block   copolymers  

from  a  mixture  of  two  symmetric  coordination  complexes  (Figure  1).    

 

Figure  1:  Cartoon  representation  for  the  mechanochemically  initiated  block  copolymerization.    

In  this  chapter  we  investigate  the  potential  of  sonication-­‐induced  directed  ligand  exchange  

towards  copolymerization.  Mechanical  stress  created  in  solution  by  implosion  of  cavitation  

bubbles  is  used  to  transfer  polymeric  ligands  from  one  metal  center  to  another.  This  leads  

to  hetero-­‐complexes  with  different  ligands  or  different  polymers  attached  to  them.  In  order  

to   investigate   mechanically   directed   ligand   exchange   three   sets   of   experiments   were  

conducted.  Firstly,  two  Pd(Im)2Cl2  complexes  with  different  molecular  weights  were  mixed  

and  the  solution  was  subjected  to  ultrasound.  Secondly,  pTHF  attached  Pd  complexes  with  

different   ligands,   NHC   and   Imidazole,   were   mixed   to   monitor   the   ultrasound-­‐induced  

formation  of  heterocomplex.  Finally,  Pd(Im-­‐pTHF)2Cl2  was  mixed  with  poly(methylacrylate)  

attached  pyridine  (Py-­‐pMA)  complex  of  palladium,  Pd(PyPMA)2Cl2,  and  sonicated  to  form  a  

block  copolymer  linked  by  Pd  coordination.  In  all  three  cases  change  in  MW  distribution  of  

mixtures  and  formation  of  new  species  were  monitored  by  GPC.  

 

 

 

 

 

 

Chapter 5

90

Page 98: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Coordination   polymers   (CPs)   have   been   used   as   mechanoresponsive   polymers   for  

mechanical   release   of   reactive   groups,   such   as   catalytically   active   sites,   luminescent   end  

groups  and  small  molecules.1–6  Our  research  group  has  shown  that  the  coordination  sphere  

of   transition-­‐metal  complexes   in   the  main  chain  of  CPs  can  be  manipulated   in  solution  by  

ultrasound.7  The   force   required   to  break  metal−ligand  bond   is  much   lower   than   the   force  

that  is  typically  required  to  break  covalent  bonds.8  Thus,  ultrasonication  of  CPs  enables  the  

specific  rupture  of  metal  ligand  coordination  bonds.  9  

Mechanical   chain   scission   in   reversible   bonds   drives   dynamic   systems   away   from  

equilibrium.  Paulusse  showed,  for  instance,  that  the  ring  chain  equilibrium  in  Pd-­‐phosphine  

coordination  polymers  is  shifted  reversibly  by  ultrasound.10  The  equilibrium  composition  of  

coordination  polymer  solutions  can  also  be  changed  by  ultrasound.  As  described  in  Chapter  

4,   Pd(Im)2Cl2   complexes   in   toluene   do   not   show   ligand   exchange   at   room   temperature.  

However,  when  free  imidazole  is  present,  it  replaces  one  of  the  coordinated  imidazoles  via  

an   associative   mechanism,   which   results   in   ligand   exchange.   This   offers   a   method   to  

selectively   change   the   composition   of   a   mixture   of   complexes   by   mechanochemical  

activation.  

 

Scheme  1:  PdCl2-­‐Imidazole  complexes  are  kinetically  stable  against   ligand  exchange  in  the  absence  

of  free  ligand  (top).  Sonication  dissociates  one  of  the  ligands  of  the  polymeric  complex  by  selectively  

breaking  the  metal  ligand  bond.  Then,  released  free  ligand  induces  associative  ligand  exchange.  

Provided  that  the  molecular  weight  (MW)  one  of  the  complexes   is  above  the   limiting  MW  

(Mlim),11,12  to  break  the  metal  ligand  bond  by  ultrasound,  ligands  can  be  exchanged  and  the  

equilibrium   will   shift   to   the   hetero-­‐complex.   Since   mechanical   force   on   the   weak  

coordinative   bond   increases   with   chain   length,   only   polymer   that   has   high   MW,   is  

mechanically   labile  and  breaks  upon  sonication.  Free   ligand  binds  to  other  complex   in   the  

NN Pd N NCl

ClO O

NN PdCl

ClO

nN N O NN Pd N N

Cl

ClC12H25 C12H25

NN Pd N NCl

ClO

nC12H25

+

NN Pd N NCl

ClC12H25 C12H25+ x

solution  and  results  in  a  solution  with  non-­‐equilibrium  composition  (Scheme  1).  In  principle,  

this   feature   of   mechanochemical   polymer   scission   allows   formation   of   block   copolymers  

from  a  mixture  of  two  symmetric  coordination  complexes  (Figure  1).    

 

Figure  1:  Cartoon  representation  for  the  mechanochemically  initiated  block  copolymerization.    

In  this  chapter  we  investigate  the  potential  of  sonication-­‐induced  directed  ligand  exchange  

towards  copolymerization.  Mechanical  stress  created  in  solution  by  implosion  of  cavitation  

bubbles  is  used  to  transfer  polymeric  ligands  from  one  metal  center  to  another.  This  leads  

to  hetero-­‐complexes  with  different  ligands  or  different  polymers  attached  to  them.  In  order  

to   investigate   mechanically   directed   ligand   exchange   three   sets   of   experiments   were  

conducted.  Firstly,  two  Pd(Im)2Cl2  complexes  with  different  molecular  weights  were  mixed  

and  the  solution  was  subjected  to  ultrasound.  Secondly,  pTHF  attached  Pd  complexes  with  

different   ligands,   NHC   and   Imidazole,   were   mixed   to   monitor   the   ultrasound-­‐induced  

formation  of  heterocomplex.  Finally,  Pd(Im-­‐pTHF)2Cl2  was  mixed  with  poly(methylacrylate)  

attached  pyridine  (Py-­‐pMA)  complex  of  palladium,  Pd(PyPMA)2Cl2,  and  sonicated  to  form  a  

block  copolymer  linked  by  Pd  coordination.  In  all  three  cases  change  in  MW  distribution  of  

mixtures  and  formation  of  new  species  were  monitored  by  GPC.  

 

 

 

 

 

 

Mechanochemically induced, directed ligand exchange

91

5

Page 99: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  discussions  

Polymer   (pTHF)   attached   imidalozyl   ligands   (Im-­‐pTHF)   were   prepared   by   terminating  

cationic   ring   opening   polymerization   of   THF   sodium   imidazolate.   pTHF-­‐Im   ligands   with  

molecular   weights   of   30k   and   6k   were   synthesized   by   adjusting   polymerization   time.   1H  

NMR,   GPC   and  MALDI-­‐TOF  were   used   to   characterize   the   ligands.  Molecular  weights   are  

given  as  the  peak  MW  determined  by  MALDI-­‐TOF.  

 

Figure  2:  GPC  traces  of  Pd(Im-­‐pTHF-­‐30k)2Cl2  (left)  and  Pd(Im-­‐pTHF-­‐6k)2Cl2  (right)  

Symmetrical  complexes  Pd(Im-­‐pTHF)2Cl2  were  prepared  by  simply  mixing  polymer  attached  

Im-­‐pTHF   ligands  with  Pd(MeCN)2Cl2.  GPC   traces   indicating  a  doubling  of  molecular  weight  

compared  to  the  free  ligands.  (Figure  2)  The  peaks  in  the  GPC  traces  of  the  60  kDa  and  12  

kDa   bis(imidazole)-­‐Pd   complexes   have   shoulders   on   the   low   MW   side.   1H   NMR   did   not  

reveal  peaks   for  un-­‐complexed   imidazole,  which  suggest   that   the  shoulders  correspond  to  

unfunctionalized  pTHF  not  end-­‐capped  with  imidazole.    

 

Figure   3:  GPC   trace   the  mixture   of   60   kDa   and   12   kDa   bis(imidazole)-­‐Pd   complexes   after   thermal  

equilibrium  has  been  reached  by  heating  a  mixture  of  Pd(Im-­‐pTHF30k)2Cl2  and  Pd(Im-­‐pTHF6k)2Cl2  at  

60  °C  for  8  h.  

When  the  two  coordination  polymers  of  60  kDa  and  12  kDa  were  mixed  in  toluene  at  a  1:1  

ratio,  no  ligand  exchange  was  observed  after  2  h,  thus  the  polymers  are  inert  to  each  other.    

However,   after   8h   at   60   oC   the   GPC   trace   showed   the   formation   of   a   new   species   with  

intermediate  molecular  weight  (Figure  3).  

A  possible  explanation  for  that  could  be  at  high  temperatures  complexes  are  more  dynamic  

and  exchange   their   ligands,   thus   a  new  coordination  polymer,  which  has  one   ligand   from  

each  complex,  was  formed  (monitored  by  GPC).  When  thermal  equilibrium  was  established,  

a  statistical  mixture  was  obtained  containing  three  polymers  with  different  MWs.  

 

Chapter 5

92

Page 100: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  discussions  

Polymer   (pTHF)   attached   imidalozyl   ligands   (Im-­‐pTHF)   were   prepared   by   terminating  

cationic   ring   opening   polymerization   of   THF   sodium   imidazolate.   pTHF-­‐Im   ligands   with  

molecular   weights   of   30k   and   6k   were   synthesized   by   adjusting   polymerization   time.   1H  

NMR,   GPC   and  MALDI-­‐TOF  were   used   to   characterize   the   ligands.  Molecular  weights   are  

given  as  the  peak  MW  determined  by  MALDI-­‐TOF.  

 

Figure  2:  GPC  traces  of  Pd(Im-­‐pTHF-­‐30k)2Cl2  (left)  and  Pd(Im-­‐pTHF-­‐6k)2Cl2  (right)  

Symmetrical  complexes  Pd(Im-­‐pTHF)2Cl2  were  prepared  by  simply  mixing  polymer  attached  

Im-­‐pTHF   ligands  with  Pd(MeCN)2Cl2.  GPC   traces   indicating  a  doubling  of  molecular  weight  

compared  to  the  free  ligands.  (Figure  2)  The  peaks  in  the  GPC  traces  of  the  60  kDa  and  12  

kDa   bis(imidazole)-­‐Pd   complexes   have   shoulders   on   the   low   MW   side.   1H   NMR   did   not  

reveal  peaks   for  un-­‐complexed   imidazole,  which  suggest   that   the  shoulders  correspond  to  

unfunctionalized  pTHF  not  end-­‐capped  with  imidazole.    

 

Figure   3:  GPC   trace   the  mixture   of   60   kDa   and   12   kDa   bis(imidazole)-­‐Pd   complexes   after   thermal  

equilibrium  has  been  reached  by  heating  a  mixture  of  Pd(Im-­‐pTHF30k)2Cl2  and  Pd(Im-­‐pTHF6k)2Cl2  at  

60  °C  for  8  h.  

When  the  two  coordination  polymers  of  60  kDa  and  12  kDa  were  mixed  in  toluene  at  a  1:1  

ratio,  no  ligand  exchange  was  observed  after  2  h,  thus  the  polymers  are  inert  to  each  other.    

However,   after   8h   at   60   oC   the   GPC   trace   showed   the   formation   of   a   new   species   with  

intermediate  molecular  weight  (Figure  3).  

A  possible  explanation  for  that  could  be  at  high  temperatures  complexes  are  more  dynamic  

and  exchange   their   ligands,   thus   a  new  coordination  polymer,  which  has  one   ligand   from  

each  complex,  was  formed  (monitored  by  GPC).  When  thermal  equilibrium  was  established,  

a  statistical  mixture  was  obtained  containing  three  polymers  with  different  MWs.  

 

Mechanochemically induced, directed ligand exchange

93

5

Page 101: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   4:   GPC   traces   for   polymer   sample   after   sonicating   the  mixture   of   Pd(Im-­‐pTHF30k)2Cl2   and  

Pd(Im-­‐pTHF6k)2Cl2  for  2h.  New  polymer  was  obtained  with  35  kDa  which  corresponds  to  the  MW  for  

Pd(NHC-­‐pTHF30k)(Im-­‐pTHF6k)Cl2.    

We   anticipated   that   the   product   distribution   could   be   biased   by  mechanical   initiation   of  

exchange  when  the  same  mixture  was  sonicated  instead  of  heated.  To  investigate  this,  a  1:1  

mixture  of  coordination  polymers  Pd(Im-­‐pTHF30k)2Cl2  and  Pd(Im-­‐pTHF6k)2Cl2  was  sonicated  

for  2h.  After  sonication  the  GPC  trace  revealed  a  single  peak  at  35  kDa  with  small  shoulders  

corresponding  to  initial  polymers.    

 

After   ligand  exchange,  the  metal  center   is  not   located  in  the  middle  of  the  polymer  chain.  

Since   the   solvodynamic   force   reaches   a   maximum   value   at   the   center   due   to   the  

centrosymmetric  nature  of  the  flow  field  with  respect  to  the  molecule,11,13,14  the  product  is  

not   responsive   to   the  mechanical   stress   although   its  MW   is   above   the  Mlim   to   break   Pd-­‐

imidazole  bond.    

 

Mechanochemical  synthesis  of  hetero-­‐complexes  

In  order   to   investigate   the  directed  mechanochemical   formation  of  hetero-­‐complexes   two  

structurally  different   ligands   coordinated   in   the   same  metal   center,  Pd(NHC-­‐pTHF)2Cl2  and  

Pd(Im-­‐pTHF)2Cl2,  were  mixed.  Both  complexes  are  kinetically  inert  at  room  temperature  i.e.  

ligand  dissociation  is  very  slow  and  the  complexes  don’t  exchange  ligands  without  heating.    

 

Scheme   2:   Route   for   synthesis   of   hetero-­‐complex   of   Pd.  Ultrasound   initiates   ligand   exchange   and  

product  possess  two  different  ligands  (NHC  and  imidazole)  coordinated  to  the  same  Pd.      

 

Figure   5:   GPC   traces   for   polymer   samples   taken   before   and   after   sonication,  mixture   of   Pd(NHC-­‐

pTHF18k)2Cl2  and   Pd(Im-­‐pTHF6k)2Cl2   (left),   mixture   of   Pd(NHC-­‐pTHF25k)2Cl2  and   Pd(Im-­‐pTHF6k)2Cl2  

(right).  

In   order   to   test   this,   Pd(NHC-­‐pTHF)2Cl2  complex   (36   kDa  or   50   kDa,   both   above  Mlim)   and  

Pd(Im-­‐pTHF)2Cl2   (12kDa,   below  Mlim)   were  mixed   in   a   1:1   ratio   in   toluene   and   sonicated.  

After   sonicating   for   2h,   a   significant   amount   of   complex   had   formed   with   a   MW  

corresponding  to  the  asymmetric  complex  (i.e.,  MWproduct  =  MWinitialP1/2  +  MWinitialP2/2).    The  

polymer   formed   after   sonication   of   12   kDa   Pd(Im-­‐pTHF)2Cl2   with   Pd(NHC-­‐pTHF25k)2Cl2  

(Figure   5)   showed   high   conversion   to   product,   with   only   small   shoulders   of   the   initial  

Chapter 5

94

Page 102: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure   4:   GPC   traces   for   polymer   sample   after   sonicating   the  mixture   of   Pd(Im-­‐pTHF30k)2Cl2   and  

Pd(Im-­‐pTHF6k)2Cl2  for  2h.  New  polymer  was  obtained  with  35  kDa  which  corresponds  to  the  MW  for  

Pd(NHC-­‐pTHF30k)(Im-­‐pTHF6k)Cl2.    

We   anticipated   that   the   product   distribution   could   be   biased   by  mechanical   initiation   of  

exchange  when  the  same  mixture  was  sonicated  instead  of  heated.  To  investigate  this,  a  1:1  

mixture  of  coordination  polymers  Pd(Im-­‐pTHF30k)2Cl2  and  Pd(Im-­‐pTHF6k)2Cl2  was  sonicated  

for  2h.  After  sonication  the  GPC  trace  revealed  a  single  peak  at  35  kDa  with  small  shoulders  

corresponding  to  initial  polymers.    

 

After   ligand  exchange,  the  metal  center   is  not   located  in  the  middle  of  the  polymer  chain.  

Since   the   solvodynamic   force   reaches   a   maximum   value   at   the   center   due   to   the  

centrosymmetric  nature  of  the  flow  field  with  respect  to  the  molecule,11,13,14  the  product  is  

not   responsive   to   the  mechanical   stress   although   its  MW   is   above   the  Mlim   to   break   Pd-­‐

imidazole  bond.    

 

Mechanochemical  synthesis  of  hetero-­‐complexes  

In  order   to   investigate   the  directed  mechanochemical   formation  of  hetero-­‐complexes   two  

structurally  different   ligands   coordinated   in   the   same  metal   center,  Pd(NHC-­‐pTHF)2Cl2  and  

Pd(Im-­‐pTHF)2Cl2,  were  mixed.  Both  complexes  are  kinetically  inert  at  room  temperature  i.e.  

ligand  dissociation  is  very  slow  and  the  complexes  don’t  exchange  ligands  without  heating.    

 

Scheme   2:   Route   for   synthesis   of   hetero-­‐complex   of   Pd.  Ultrasound   initiates   ligand   exchange   and  

product  possess  two  different  ligands  (NHC  and  imidazole)  coordinated  to  the  same  Pd.      

 

Figure   5:   GPC   traces   for   polymer   samples   taken   before   and   after   sonication,  mixture   of   Pd(NHC-­‐

pTHF18k)2Cl2  and   Pd(Im-­‐pTHF6k)2Cl2   (left),   mixture   of   Pd(NHC-­‐pTHF25k)2Cl2  and   Pd(Im-­‐pTHF6k)2Cl2  

(right).  

In   order   to   test   this,   Pd(NHC-­‐pTHF)2Cl2  complex   (36   kDa  or   50   kDa,   both   above  Mlim)   and  

Pd(Im-­‐pTHF)2Cl2   (12kDa,   below  Mlim)   were  mixed   in   a   1:1   ratio   in   toluene   and   sonicated.  

After   sonicating   for   2h,   a   significant   amount   of   complex   had   formed   with   a   MW  

corresponding  to  the  asymmetric  complex  (i.e.,  MWproduct  =  MWinitialP1/2  +  MWinitialP2/2).    The  

polymer   formed   after   sonication   of   12   kDa   Pd(Im-­‐pTHF)2Cl2   with   Pd(NHC-­‐pTHF25k)2Cl2  

(Figure   5)   showed   high   conversion   to   product,   with   only   small   shoulders   of   the   initial  

Mechanochemically induced, directed ligand exchange

95

5

Page 103: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

polymers.  When   Pd(NHC-­‐pTHF18k)2Cl2  was   used,   shoulders   were   higher   in   intensity   after  

sonicating   for   2h   indicating   that   the   conversion   was   lower.   Product   distribution   in   both  

experiments   unequivocally   shows   that   hetero-­‐complex   formation   is   a   directed,  

mechanochemical  process  rather  than  a  thermal  process,  which  would   lead  to  a  statistical  

mixture  of  symmetric  starting  material  and  asymmetric  complexes.    

As  a  control  experiment,  Pd(NHC-­‐pTHF18k)2Cl2  was  sonicated  in  the  presence  of  acetonitrile  

and  acetic  acid  to  trap  scission  products.  This  resulted  in  products  with  MW  around  18  kDa,  

with  a  peak  in  GPC  that  is  clearly  distinguishable  from  the  product  peak  when  a  mixture  of  

Pd(NHC-­‐pTHF18k)2Cl2  and  Pd(Im-­‐pTHF6k)2Cl2    was  sonicated  (Figure  6).  

 

Figure  6:  Comparison  of  the  GPC  traces  for  Pd(NHC-­‐pTHF18k)2Cl2  before  and  after  sonication  in  the  

presence  of  Pd(Im-­‐pTHF6k)2Cl2    or  the  trapping  agents  acetonitrile  and  acetic  acid.  

Mechanochemical  synthesis  of  block  copolymers  

In   order   to   investigate   the   mechanochemical   synthesis   of   block   copolymers,   pyridine-­‐

capped  poly  methyl  acrylate  based  polymeric  ligand  (PyPMA,  Scheme  3)  was  synthesized  via  

SET-­‐LRP   of   methyl   acrylate15   using   4-­‐pyridinyl-­‐2-­‐bromoisobutyrate   as   the   initiator.  

Combining   PyPMA  with   0.5   equiv   of   Pd(CH3CN)2Cl2   in   DCM   at   room   temperature   for   2   h  

resulted   in   a   doubling   of   MW,   as   determined   by   GPC,   that   was   consistent   with   the  

displacement  of  thermally  labile  CH3CN  by  pyridine-­‐based  ligands.    

 

Scheme  3:    Synthesis  of  PyPMA  ligand  and  Pd(PyPMA)2Cl2  i)  2-­‐bromo-­‐isobutyrylbromide,  CH2Cl2,  0oC-­‐

25oC,  ii)  MA,  Me6TREN,  Cu  wire,  RT,  iii)  Pd(CH3CN)2Cl2,  CH2Cl2,  RT.  

Pd(PyPMA-­‐45k)2Cl2  complex  (90  kDa)  and  Pd(Im-­‐pTHF)2Cl2  (10  kDa,  below  Mlim)  were  mixed  

in   a   1:1   ratio   in   toluene.   GPC   traces   for   polymers   before   mixing   revealed   monomodal  

distribution.   A   new   peak   corresponding   to   a   MW   of   50   kDa   appeared   upon   mixing   that  

suggests  a  ligand  swapping  between  species  without  sonication.  However,  concentrations  of  

different  species  did  not  change  during  2  h  at  room  temperature  (Figure  7)   indicating  that  

equilibrium   was   achieved   immediately.   After   sonicating   the   same   mixture   for   2   h,   a  

significant   amount   of   complex   had   formed   (Figure   8)   with   a   MW   corresponding   to   the  

asymmetric  complex  (50  kDa)  that  consists  of  two  different  ligands  (pyridine  and  imidazole)  

and  two  different  polymers  (PTHF  and  PMA).  Comparing  the  GPC  traces  of  hetero-­‐complex  

that   contains   block   copolymer   and   a   partially   dissociated   complex   showed   that   the  

molecular   weight   of   hetero-­‐complex   is   indeed   significantly   higher   than   the   dissociation  

product  of  Pd(PyPMA)2Cl2  (Figure  8).    

Chapter 5

96

Page 104: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

polymers.  When   Pd(NHC-­‐pTHF18k)2Cl2  was   used,   shoulders   were   higher   in   intensity   after  

sonicating   for   2h   indicating   that   the   conversion   was   lower.   Product   distribution   in   both  

experiments   unequivocally   shows   that   hetero-­‐complex   formation   is   a   directed,  

mechanochemical  process  rather  than  a  thermal  process,  which  would   lead  to  a  statistical  

mixture  of  symmetric  starting  material  and  asymmetric  complexes.    

As  a  control  experiment,  Pd(NHC-­‐pTHF18k)2Cl2  was  sonicated  in  the  presence  of  acetonitrile  

and  acetic  acid  to  trap  scission  products.  This  resulted  in  products  with  MW  around  18  kDa,  

with  a  peak  in  GPC  that  is  clearly  distinguishable  from  the  product  peak  when  a  mixture  of  

Pd(NHC-­‐pTHF18k)2Cl2  and  Pd(Im-­‐pTHF6k)2Cl2    was  sonicated  (Figure  6).  

 

Figure  6:  Comparison  of  the  GPC  traces  for  Pd(NHC-­‐pTHF18k)2Cl2  before  and  after  sonication  in  the  

presence  of  Pd(Im-­‐pTHF6k)2Cl2    or  the  trapping  agents  acetonitrile  and  acetic  acid.  

Mechanochemical  synthesis  of  block  copolymers  

In   order   to   investigate   the   mechanochemical   synthesis   of   block   copolymers,   pyridine-­‐

capped  poly  methyl  acrylate  based  polymeric  ligand  (PyPMA,  Scheme  3)  was  synthesized  via  

SET-­‐LRP   of   methyl   acrylate15   using   4-­‐pyridinyl-­‐2-­‐bromoisobutyrate   as   the   initiator.  

Combining   PyPMA  with   0.5   equiv   of   Pd(CH3CN)2Cl2   in   DCM   at   room   temperature   for   2   h  

resulted   in   a   doubling   of   MW,   as   determined   by   GPC,   that   was   consistent   with   the  

displacement  of  thermally  labile  CH3CN  by  pyridine-­‐based  ligands.    

 

Scheme  3:    Synthesis  of  PyPMA  ligand  and  Pd(PyPMA)2Cl2  i)  2-­‐bromo-­‐isobutyrylbromide,  CH2Cl2,  0oC-­‐

25oC,  ii)  MA,  Me6TREN,  Cu  wire,  RT,  iii)  Pd(CH3CN)2Cl2,  CH2Cl2,  RT.  

Pd(PyPMA-­‐45k)2Cl2  complex  (90  kDa)  and  Pd(Im-­‐pTHF)2Cl2  (10  kDa,  below  Mlim)  were  mixed  

in   a   1:1   ratio   in   toluene.   GPC   traces   for   polymers   before   mixing   revealed   monomodal  

distribution.   A   new   peak   corresponding   to   a   MW   of   50   kDa   appeared   upon   mixing   that  

suggests  a  ligand  swapping  between  species  without  sonication.  However,  concentrations  of  

different  species  did  not  change  during  2  h  at  room  temperature  (Figure  7)   indicating  that  

equilibrium   was   achieved   immediately.   After   sonicating   the   same   mixture   for   2   h,   a  

significant   amount   of   complex   had   formed   (Figure   8)   with   a   MW   corresponding   to   the  

asymmetric  complex  (50  kDa)  that  consists  of  two  different  ligands  (pyridine  and  imidazole)  

and  two  different  polymers  (PTHF  and  PMA).  Comparing  the  GPC  traces  of  hetero-­‐complex  

that   contains   block   copolymer   and   a   partially   dissociated   complex   showed   that   the  

molecular   weight   of   hetero-­‐complex   is   indeed   significantly   higher   than   the   dissociation  

product  of  Pd(PyPMA)2Cl2  (Figure  8).    

Mechanochemically induced, directed ligand exchange

97

5

Page 105: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  7:  GPC  traces  for  the  mixture  of  Pd(PyPMA-­‐45k)2Cl2  complex  (90  kDa)  and  Pd(Im-­‐pTHF6k)2Cl2  

showed  that  the  mixture  is  in  equilibrium  and  concentration  of  different  species  did  not  change  after  

120  min.  

As   a   control   experiment,   Pd(PyPMA-­‐45k)2Cl2   was   sonicated   and   scission   products   were  

trapped  with  CH3CN  and  trifluoroacetic  acid  (TFA).  Mid-­‐chain  scission  yields  polymers  with  

MW   around   45   kDa.   Comparing   the   GPC   traces   of   hetero-­‐complex   that   contains   block  

copolymer  and  a  partially  dissociated  complex  showed  that  the  molecular  weight  of  hetero-­‐

complex  is   indeed  significantly  higher  than  the  dissociation  product  of  Pd(PyPMA)2Cl2  thus,  

the   new   species   is   the   hetero-­‐complex   rather   than   the   fragments   of   Pd(PyPMA-­‐45k)2Cl2.  

(Figure  8).    

 

Figure   8:  GPC   traces   for   the  different   species:  a)  Pd(PyPMA-­‐45k)2Cl2  (black),  PyPMA-­‐45k   (red),   and  

partially   dissociated   complex   (blue).   b)  Mixture   of   Pd(PyPMA-­‐45k)2Cl2  and   Pd(Im-­‐pTHF6k)2Cl2  after  

sonicating   for   120   min   (red).   Dashed   line   corresponds   to   the   GPC   trace   for   partially   dissociated  

complex  Pd(PyPMA-­‐45k)2Cl2.  

Conclusions  

In  Chapters  2  and  4,  ultrasonic  scission  in  Pd(NHC-­‐pTHF)2Cl2  and  Pd(Im-­‐pTHF)2Cl2    complexes  

was  analyzed  in  detail.  In  the  current  chapter,  the  information  on  ligand  exchange  kinetics,  

limiting  molecular  weights  and  scission  rates  has  been  used  to  synthesize  hetero-­‐complexes.    

Mechano-­‐chemically   induced   hetero-­‐complex   formation   is   slower   than  mechanochemical  

bond  scission  of  Pd(NHC)2Cl2  complexes  in  the  presence  of  trapping  agents.    After  sonicating  

the   mixture   for   2h,   significant   amounts   of   starting   polymers   were   still   present.   An  

explanation  is  that  the  formation  of  triscoordinated  complexes,  which  are  the  intermediates  

in  the  formation  of  heterocomplexes,  is  slow  compared  to  trapping,  while  the  concentration  

of  the  second  polymer  is  also  lower  than  the  trapping  agents.    

Results   presented   in   this   chapter   show   that   mechanochemically   induced   block  

copolymerization  is  possible  by  attaching  desired  polymers  on  easily  modified  ligands.      

 

 

 

Chapter 5

98

Page 106: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  7:  GPC  traces  for  the  mixture  of  Pd(PyPMA-­‐45k)2Cl2  complex  (90  kDa)  and  Pd(Im-­‐pTHF6k)2Cl2  

showed  that  the  mixture  is  in  equilibrium  and  concentration  of  different  species  did  not  change  after  

120  min.  

As   a   control   experiment,   Pd(PyPMA-­‐45k)2Cl2   was   sonicated   and   scission   products   were  

trapped  with  CH3CN  and  trifluoroacetic  acid  (TFA).  Mid-­‐chain  scission  yields  polymers  with  

MW   around   45   kDa.   Comparing   the   GPC   traces   of   hetero-­‐complex   that   contains   block  

copolymer  and  a  partially  dissociated  complex  showed  that  the  molecular  weight  of  hetero-­‐

complex  is   indeed  significantly  higher  than  the  dissociation  product  of  Pd(PyPMA)2Cl2  thus,  

the   new   species   is   the   hetero-­‐complex   rather   than   the   fragments   of   Pd(PyPMA-­‐45k)2Cl2.  

(Figure  8).    

 

Figure   8:  GPC   traces   for   the  different   species:  a)  Pd(PyPMA-­‐45k)2Cl2  (black),  PyPMA-­‐45k   (red),   and  

partially   dissociated   complex   (blue).   b)  Mixture   of   Pd(PyPMA-­‐45k)2Cl2  and   Pd(Im-­‐pTHF6k)2Cl2  after  

sonicating   for   120   min   (red).   Dashed   line   corresponds   to   the   GPC   trace   for   partially   dissociated  

complex  Pd(PyPMA-­‐45k)2Cl2.  

Conclusions  

In  Chapters  2  and  4,  ultrasonic  scission  in  Pd(NHC-­‐pTHF)2Cl2  and  Pd(Im-­‐pTHF)2Cl2    complexes  

was  analyzed  in  detail.  In  the  current  chapter,  the  information  on  ligand  exchange  kinetics,  

limiting  molecular  weights  and  scission  rates  has  been  used  to  synthesize  hetero-­‐complexes.    

Mechano-­‐chemically   induced   hetero-­‐complex   formation   is   slower   than  mechanochemical  

bond  scission  of  Pd(NHC)2Cl2  complexes  in  the  presence  of  trapping  agents.    After  sonicating  

the   mixture   for   2h,   significant   amounts   of   starting   polymers   were   still   present.   An  

explanation  is  that  the  formation  of  triscoordinated  complexes,  which  are  the  intermediates  

in  the  formation  of  heterocomplexes,  is  slow  compared  to  trapping,  while  the  concentration  

of  the  second  polymer  is  also  lower  than  the  trapping  agents.    

Results   presented   in   this   chapter   show   that   mechanochemically   induced   block  

copolymerization  is  possible  by  attaching  desired  polymers  on  easily  modified  ligands.      

 

 

 

Mechanochemically induced, directed ligand exchange

99

5

Page 107: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Experimental    

General  

All  chemicals  were  purchased  from  commercial  sources  and  used  without  further  purification  unless  

specified  otherwise.  Toluene  was  dried  over  4A  molecular  sieves.  Gel  permeation  chromatography  

(GPC)  was  performed  on  a  Shimadzu  LC10-­‐AD,  using  Polymer  Laboratories  PL  Gel  5μm  MIXED-­‐C  and  

MIXEDD  columns    (linear  range  of  MW:  200–2000000  g/mol),  a  Shimadzu  SPD-­‐M10A  UV-­‐vis  detector  

at  254  nm  and  RID-­‐10A  refractive  index  detector,  and  THF  as  eluent  at  a  flow  rate  of  1  mL/min  (20  

°C).   Polystyrene   standards   were   used   for   calibration.   Pd(NHC-­‐pTHF)2Cl2  and   Pd(Im-­‐pTHF)2Cl2    were  

synthesized  according  to  procedures  presented  in  Chapter  2  and  Chapter  4  respectively.        

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   reactor   with   a   volume   of   10  mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   The   temperature   in   the   reactor  was  maintained  with   a   Lauda   E300   cooling  

bath   and   measured   using   a   0.5   mm   diameter   thermocouple.   5   ml   solutions   were   sonicated  

continuously;   temperature   of   the   solution  was   kept   constant   by   circulating  water   at   2oC.   Prior   to  

sonication  CH4  was  bubbled  through  solution  via  teflon  tubing.  Aliquots  of  100  µL  were  taken  after  

2h.  Toluene  was  removed  under  reduced  pressure;  residues  were  dissolved  in  THF  and  submitted  to  

GPC.    

Synthesis  of  4-­‐pyridinyl-­‐2-­‐bromo-­‐isobutyrate  (PyI)  

 To   a   solution   of   4-­‐hydroxypyridine   (1.00   g,   10.5   mmol)   in   CH2Cl2   (dry,   15   mL),   2-­‐bromo-­‐

isobutyrylbromide  (1.43  mL,  11.6  mmol)  was  added  drop-­‐wise  at  0  °C.  The  resulting  suspension  was  

allowed  to  warm  to  ambient  temperature  and  then  stirred  overnight  under  argon.  The  reaction  was  

then  diluted  with  CH2Cl2  (30  mL),  washed  with  saturated  NaHCO3  (3  ×  50  mL)  and  water  (3  ×  50  mL),  

dried  over  MgSO4  and  then  passed  through  a  short  column  of  neutral  alumina  (CH2Cl2).  The  solvent  

was  removed  under  reduced  pressure  to  afford  the  desired  product  (2.18  g,  8.94  mmol)  as  a  clear,  

viscous  liquid  in  85%  yield.  1H  NMR  (CDCl3,  400  MHz):  δ  8.64  (d,  3J  =  6.0  Hz,  2H),  7.14  (d,  3J  =  6.8  Hz,  

2H),  2.04  (s,  6H).    

 

 

Synthesis  of  PyPMA  

25  mL  Schlenk  flask  was  charged  with  methyl  acrylate  (MA)  (4.0  mL,  44  mmol),  10  mM  solution  of  

tris(2-­‐[dimethyl]-­‐aminoethyl)amine  (Me6TREN,  4.0  mL,  40  μmol)   in  DMSO  and  4-­‐pyridinyl-­‐2-­‐bromo-­‐

isobutyrate  (1)  (15.5  mg,  64  μmol).    Solution  was  degassed  by  freeze-­‐pump-­‐thaw  cycle  (3  times).  A  

magnetic  stir-­‐bar  wrapped  with  copper  wire,  and  inserted  into  the  solution.  The  solution  was  stirred  

at  ambient  temperature  for  1  h.  Polymerization  was  terminated  by  THF  (10  ml,  stabilized  by  BHT).  

The  resulting  deep  blue  solution  was  then  added  slowly  to  excess  methanol  (150  mL),  which  caused  

a  polymeric  material  to  precipitate  as  a  gummy  solid.  Centrifugation  was  performed  to  sediment  out  

the  polymer  and  the  supernatant  was  decanted.  The  residue  was  then  washed  with  methanol  (5  ×  

20  mL)  and  dried  under   reduced  pressure   to  afford   the  desired  polymer.  GPC   (THF)  Mn  =  50  kDa,  

MALDI-­‐TOF:  45  kDA.  1H  NMR  (CD2Cl2,  400  MHz):  δ  8.73  (d,  3J  =  8  Hz),  7.80  (d,  3J  =8  Hz),  3.51  (br),  2.30  

(br),  1.91  (br),  1.68  (br),  1.45  (br)    

Synthesis  of  Pd(PyPMA)2Cl2  

PyPMA  (400  mg,  8.90  μmol  )was  mixed  with  Pd(CH3CN)2Cl2  (1.15  mg,  4.45  μmol)  in  DCM  and  stirred  

at   room   temperature   for   2   h   under   argon.   Resulting   light   yellow   solution   was   filtered   over   filter  

paper  and  solvent  was  evaporated  under  reduced  pressure.  1H  NMR  (CD2Cl2,  400  MHz):  δ  8.80  (d,  3J  

=  8  Hz),  7.23  (d,  3J  =8  Hz),  3.51  (br),  2.30  (br),  1.91  (br),  1.68  (br),  1.45  (br)  

 

 

 

 

 

 

 

 

 

 

 

Chapter 5

100

Page 108: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Experimental    

General  

All  chemicals  were  purchased  from  commercial  sources  and  used  without  further  purification  unless  

specified  otherwise.  Toluene  was  dried  over  4A  molecular  sieves.  Gel  permeation  chromatography  

(GPC)  was  performed  on  a  Shimadzu  LC10-­‐AD,  using  Polymer  Laboratories  PL  Gel  5μm  MIXED-­‐C  and  

MIXEDD  columns    (linear  range  of  MW:  200–2000000  g/mol),  a  Shimadzu  SPD-­‐M10A  UV-­‐vis  detector  

at  254  nm  and  RID-­‐10A  refractive  index  detector,  and  THF  as  eluent  at  a  flow  rate  of  1  mL/min  (20  

°C).   Polystyrene   standards   were   used   for   calibration.   Pd(NHC-­‐pTHF)2Cl2  and   Pd(Im-­‐pTHF)2Cl2    were  

synthesized  according  to  procedures  presented  in  Chapter  2  and  Chapter  4  respectively.        

Sonication  experiments  

A   homemade,   double-­‐jacketed   glass   reactor   with   a   volume   of   10  mL   was   used   in   the   sonication  

experiments.  A  Sonics  and  Materials  20  kHz,  0.5   in.  diameter  titanium  alloy  ultrasound  probe  with  

half   wave   extension   (parts   630-­‐0220   and   630-­‐0410)   was   operated   using   a   Sonics   and   Materials  

VC750   power   supply.   The   temperature   in   the   reactor  was  maintained  with   a   Lauda   E300   cooling  

bath   and   measured   using   a   0.5   mm   diameter   thermocouple.   5   ml   solutions   were   sonicated  

continuously;   temperature   of   the   solution  was   kept   constant   by   circulating  water   at   2oC.   Prior   to  

sonication  CH4  was  bubbled  through  solution  via  teflon  tubing.  Aliquots  of  100  µL  were  taken  after  

2h.  Toluene  was  removed  under  reduced  pressure;  residues  were  dissolved  in  THF  and  submitted  to  

GPC.    

Synthesis  of  4-­‐pyridinyl-­‐2-­‐bromo-­‐isobutyrate  (PyI)  

 To   a   solution   of   4-­‐hydroxypyridine   (1.00   g,   10.5   mmol)   in   CH2Cl2   (dry,   15   mL),   2-­‐bromo-­‐

isobutyrylbromide  (1.43  mL,  11.6  mmol)  was  added  drop-­‐wise  at  0  °C.  The  resulting  suspension  was  

allowed  to  warm  to  ambient  temperature  and  then  stirred  overnight  under  argon.  The  reaction  was  

then  diluted  with  CH2Cl2  (30  mL),  washed  with  saturated  NaHCO3  (3  ×  50  mL)  and  water  (3  ×  50  mL),  

dried  over  MgSO4  and  then  passed  through  a  short  column  of  neutral  alumina  (CH2Cl2).  The  solvent  

was  removed  under  reduced  pressure  to  afford  the  desired  product  (2.18  g,  8.94  mmol)  as  a  clear,  

viscous  liquid  in  85%  yield.  1H  NMR  (CDCl3,  400  MHz):  δ  8.64  (d,  3J  =  6.0  Hz,  2H),  7.14  (d,  3J  =  6.8  Hz,  

2H),  2.04  (s,  6H).    

 

 

Synthesis  of  PyPMA  

25  mL  Schlenk  flask  was  charged  with  methyl  acrylate  (MA)  (4.0  mL,  44  mmol),  10  mM  solution  of  

tris(2-­‐[dimethyl]-­‐aminoethyl)amine  (Me6TREN,  4.0  mL,  40  μmol)   in  DMSO  and  4-­‐pyridinyl-­‐2-­‐bromo-­‐

isobutyrate  (1)  (15.5  mg,  64  μmol).    Solution  was  degassed  by  freeze-­‐pump-­‐thaw  cycle  (3  times).  A  

magnetic  stir-­‐bar  wrapped  with  copper  wire,  and  inserted  into  the  solution.  The  solution  was  stirred  

at  ambient  temperature  for  1  h.  Polymerization  was  terminated  by  THF  (10  ml,  stabilized  by  BHT).  

The  resulting  deep  blue  solution  was  then  added  slowly  to  excess  methanol  (150  mL),  which  caused  

a  polymeric  material  to  precipitate  as  a  gummy  solid.  Centrifugation  was  performed  to  sediment  out  

the  polymer  and  the  supernatant  was  decanted.  The  residue  was  then  washed  with  methanol  (5  ×  

20  mL)  and  dried  under   reduced  pressure   to  afford   the  desired  polymer.  GPC   (THF)  Mn  =  50  kDa,  

MALDI-­‐TOF:  45  kDA.  1H  NMR  (CD2Cl2,  400  MHz):  δ  8.73  (d,  3J  =  8  Hz),  7.80  (d,  3J  =8  Hz),  3.51  (br),  2.30  

(br),  1.91  (br),  1.68  (br),  1.45  (br)    

Synthesis  of  Pd(PyPMA)2Cl2  

PyPMA  (400  mg,  8.90  μmol  )was  mixed  with  Pd(CH3CN)2Cl2  (1.15  mg,  4.45  μmol)  in  DCM  and  stirred  

at   room   temperature   for   2   h   under   argon.   Resulting   light   yellow   solution   was   filtered   over   filter  

paper  and  solvent  was  evaporated  under  reduced  pressure.  1H  NMR  (CD2Cl2,  400  MHz):  δ  8.80  (d,  3J  

=  8  Hz),  7.23  (d,  3J  =8  Hz),  3.51  (br),  2.30  (br),  1.91  (br),  1.68  (br),  1.45  (br)  

 

 

 

 

 

 

 

 

 

 

 

Mechanochemically induced, directed ligand exchange

101

5

Page 109: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

References  

(1)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(2)     Chen,  Y.;  Spiering,  A.  J.  H.;  Karthikeyan,  S.;  Peters,  G.  W.  M.;  Meijer,  E.  W.;  Sijbesma,  R.  P.  Nat.  

Chem.  2012,  4  (7),  559–562.  

(3)     Larsen,  M.  B.;  Boydston,  A.  J.  J.  Am.  Chem.  Soc.  2013,  135  (22),  8189–8192.  

(4)     Sottos,  N.  R.  Nat.  Chem.  2014,  6  (5),  381–383.  

(5)     Diesendruck,   C.   E.;   Steinberg,   B.   D.;   Sugai,   N.;   Silberstein,  M.   N.;   Sottos,   N.   R.;  White,   S.   R.;  

Braun,  P.  V.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2012,  134  (30),  12446–12449.  

(6)     Vermonden,  T.;  van  Steenbergen,  M.  J.;  Besseling,  N.  A.  M.;  Marcelis,  A.  T.  M.;  Hennink,  W.  E.;  

Sudhölter,  E.  J.  R.;  Cohen  Stuart,  M.  A.  J.  Am.  Chem.  Soc.  2004,  126  (48),  15802–15808.  

(7)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(8)     Groote,   R.;   van   Haandel,   L.;   Sijbesma,   R.   P.   J.   Polym.   Sci.   Part   Polym.   Chem.   2012,   50   (23),  

4929–4935.  

(9)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2008,  No.  37,  4416–4418.  

(10)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.-­‐Int.  Ed.  2004,  43  (34),  4460–4462.  

(11)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(12)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  Angew.  Chem.  Int.  Ed.  2009,  48  (23),  4190–4193.  

(13)     Odell,  J.  A.;  Keller,  A.  J.  Polym.  Sci.  Part  B  Polym.  Phys.  1986,  24  (9),  1889–1916.  

(14)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

(15)     Percec,   V.;   Guliashvili,   T.;   Ladislaw,   J.   S.;   Wistrand,   A.;   Stjerndahl,   A.;   Sienkowska,   M.   J.;  

Monteiro,  M.  J.;  Sahoo,  S.  J.  Am.  Chem.  Soc.  2006,  128  (43),  14156–14165.  

 

 

 

 

Chapter 5

102

Page 110: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Chapter 6Transition metal bearing supramolecular polymer networks: Towards self-healing applications

Synthesis of polymethyl acrylate (PMA) with 5% imidazole (VIm) groups and initial attempts

to prepare self-healing p(MA-co-VIm)-metal films are reported. Pendant VIm serves as a

ligand that leads to reversible cross-linking upon complexation to metal salts such as Copper

and Palladium.

References  

(1)     Piermattei,  A.;  Karthikeyan,  S.;  Sijbesma,  R.  P.  Nat.  Chem.  2009,  1  (2),  133–137.  

(2)     Chen,  Y.;  Spiering,  A.  J.  H.;  Karthikeyan,  S.;  Peters,  G.  W.  M.;  Meijer,  E.  W.;  Sijbesma,  R.  P.  Nat.  

Chem.  2012,  4  (7),  559–562.  

(3)     Larsen,  M.  B.;  Boydston,  A.  J.  J.  Am.  Chem.  Soc.  2013,  135  (22),  8189–8192.  

(4)     Sottos,  N.  R.  Nat.  Chem.  2014,  6  (5),  381–383.  

(5)     Diesendruck,   C.   E.;   Steinberg,   B.   D.;   Sugai,   N.;   Silberstein,  M.   N.;   Sottos,   N.   R.;  White,   S.   R.;  

Braun,  P.  V.;  Moore,  J.  S.  J.  Am.  Chem.  Soc.  2012,  134  (30),  12446–12449.  

(6)     Vermonden,  T.;  van  Steenbergen,  M.  J.;  Besseling,  N.  A.  M.;  Marcelis,  A.  T.  M.;  Hennink,  W.  E.;  

Sudhölter,  E.  J.  R.;  Cohen  Stuart,  M.  A.  J.  Am.  Chem.  Soc.  2004,  126  (48),  15802–15808.  

(7)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  J.  Polym.  Sci.  Part  Polym.  Chem.  2006,  44  (19),  5445–5453.  

(8)     Groote,   R.;   van   Haandel,   L.;   Sijbesma,   R.   P.   J.   Polym.   Sci.   Part   Polym.   Chem.   2012,   50   (23),  

4929–4935.  

(9)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Chem.  Commun.  2008,  No.  37,  4416–4418.  

(10)     Paulusse,  J.  M.  J.;  Sijbesma,  R.  P.  Angew.  Chem.-­‐Int.  Ed.  2004,  43  (34),  4460–4462.  

(11)     Nakano,  A.;  Minoura,  Y.  Macromolecules  1975,  8  (5),  677–680.  

(12)     Ribas-­‐Arino,  J.;  Shiga,  M.;  Marx,  D.  Angew.  Chem.  Int.  Ed.  2009,  48  (23),  4190–4193.  

(13)     Odell,  J.  A.;  Keller,  A.  J.  Polym.  Sci.  Part  B  Polym.  Phys.  1986,  24  (9),  1889–1916.  

(14)     Odell,  J.  A.;  Muller,  A.  J.;  Narh,  K.  A.;  Keller,  A.  Macromolecules  1990,  23  (12),  3092–3103.  

(15)     Percec,   V.;   Guliashvili,   T.;   Ladislaw,   J.   S.;   Wistrand,   A.;   Stjerndahl,   A.;   Sienkowska,   M.   J.;  

Monteiro,  M.  J.;  Sahoo,  S.  J.  Am.  Chem.  Soc.  2006,  128  (43),  14156–14165.  

 

 

 

 

Page 111: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Supramolecular   polymer   networks   are   three-­‐dimensional   assemblies   of   macromolecules  

connected  by  non-­‐covalent  bonds.1  These  networks  form  a  useful  class  of  materials  with  the  

potential  utility  caused  by  the  reversibility  of  their  constituent  supramolecular  bonds.2  Such  

bonds  formed  through  hydrogen  bonding  or  transition  metal  complexation  are  quite  strong  

but   under   stress   they   are   broken   more   easily   than   covalent   bonds.3–5   Their   reversibility  

renders   supramolecular  polymer  networks  useful   for  applications  as   self-­‐healing   scaffolds,  

or  in  shape-­‐memory  materials.6,7  

Up  to  date,  self-­‐healing  materials  have  been  developed  using  various  reversible  bonds  such  

as   Diels-­‐Alder   adducts,8   hydrogen   bonding,9,10   ionic   interactions,11   π-­‐π   interactions,12   or  

host-­‐guest   interactions.13   In   addition,   reversible   coordination   bonds   have   been   used   to  

obtain   self-­‐healing   materials   since   metal-­‐ligand   interaction   is   less   sensitive   to   moisture  

compared   to   hydrogen   bonds.   For   the   application   of  metallo-­‐supramolecular   polymers   in  

self-­‐healing  materials,  mostly  nitrogen-­‐based  aromatic  ligands  were  used  (Figure  1).  Metal-­‐

ligand  bond  in  these  materials  was  either  used  for  chain  extension  or  as  a  cross-­‐linker.  5,14–23  

We   have   shown   previously   that   bisimidazole-­‐Pd   complexes   do   not   show   any   significant  

ligand  exchange  between  two  different  Pd  centers  at  room  temperature.  However,  excess  

free   imidazole,   which   is   formed  when   Im-­‐Pd   bonds   break   under   stress,   initiates   a   ligand  

exchange  reaction  and  replaces  one  of  the  imidazoles  coordinated  to  Pd  via  an  associative  

mechanism.  We  anticipated  that  polymer  films,  which  possess  such  Im-­‐metal  coordination  

as   cross-­‐linker,   would   be   ideal   candidates   for   autonomous   self-­‐healing   materials.   High  

association   affinity   between   imidazole   and   transition   metals   would   make   the   material  

strong  while   promoting   a   fast   healing   for   a   broken   or   partially   deformed   sample.   In   this  

chapter,  synthesis  of  polymethyl  acrylate  (PMA)  with  pendant  imidazole  (VIm)  groups  which  

serve  as  the  cross-­‐linking  sites,  and  preliminary  results  for  self-­‐healing  property  of  p(MA-­‐co-­‐

VIm)-­‐Metal  films  are  reported.        

 

 

 

Figure   1:   Examples   from   recent   literature   for   polymeric   ligands   and   metals   investigated   in   self-­‐

healing  metallosupramolecular  polymers.5,  14-­‐23  

 

 

 

 

ON

NN

NN

O N

NN

NN

x yn

NH

NH

O O NH

NH

O NH

O O O O O

NH

O O NH

NH

O O O1 2 1 4 1 3 1

O4

1 OCNNCO

2N O NO₂

OHHO

3

4

NNN N N N

NHO

OH

HOO

H

NN N

OO ORn m

n m

/Zn²⁺/or/La³⁺ /Fe²⁺/or/Cd²⁺

/Zn²⁺/or/Eu³⁺

O O

OOO

O

/Fe²⁺

a) b) c)

d)

NR₂

NR₂R₂N

R₂N

NNN NO O

NH

R₂N

R₂N

NNO

O

HN

O

NR₂

NR₂

N NO

e)

/Pd²⁺/or/Pt²⁺

Chapter 6

104

Page 112: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Introduction  

Supramolecular   polymer   networks   are   three-­‐dimensional   assemblies   of   macromolecules  

connected  by  non-­‐covalent  bonds.1  These  networks  form  a  useful  class  of  materials  with  the  

potential  utility  caused  by  the  reversibility  of  their  constituent  supramolecular  bonds.2  Such  

bonds  formed  through  hydrogen  bonding  or  transition  metal  complexation  are  quite  strong  

but   under   stress   they   are   broken   more   easily   than   covalent   bonds.3–5   Their   reversibility  

renders   supramolecular  polymer  networks  useful   for  applications  as   self-­‐healing   scaffolds,  

or  in  shape-­‐memory  materials.6,7  

Up  to  date,  self-­‐healing  materials  have  been  developed  using  various  reversible  bonds  such  

as   Diels-­‐Alder   adducts,8   hydrogen   bonding,9,10   ionic   interactions,11   π-­‐π   interactions,12   or  

host-­‐guest   interactions.13   In   addition,   reversible   coordination   bonds   have   been   used   to  

obtain   self-­‐healing   materials   since   metal-­‐ligand   interaction   is   less   sensitive   to   moisture  

compared   to   hydrogen   bonds.   For   the   application   of  metallo-­‐supramolecular   polymers   in  

self-­‐healing  materials,  mostly  nitrogen-­‐based  aromatic  ligands  were  used  (Figure  1).  Metal-­‐

ligand  bond  in  these  materials  was  either  used  for  chain  extension  or  as  a  cross-­‐linker.  5,14–23  

We   have   shown   previously   that   bisimidazole-­‐Pd   complexes   do   not   show   any   significant  

ligand  exchange  between  two  different  Pd  centers  at  room  temperature.  However,  excess  

free   imidazole,   which   is   formed  when   Im-­‐Pd   bonds   break   under   stress,   initiates   a   ligand  

exchange  reaction  and  replaces  one  of  the  imidazoles  coordinated  to  Pd  via  an  associative  

mechanism.  We  anticipated  that  polymer  films,  which  possess  such  Im-­‐metal  coordination  

as   cross-­‐linker,   would   be   ideal   candidates   for   autonomous   self-­‐healing   materials.   High  

association   affinity   between   imidazole   and   transition   metals   would   make   the   material  

strong  while   promoting   a   fast   healing   for   a   broken   or   partially   deformed   sample.   In   this  

chapter,  synthesis  of  polymethyl  acrylate  (PMA)  with  pendant  imidazole  (VIm)  groups  which  

serve  as  the  cross-­‐linking  sites,  and  preliminary  results  for  self-­‐healing  property  of  p(MA-­‐co-­‐

VIm)-­‐Metal  films  are  reported.        

 

 

 

Figure   1:   Examples   from   recent   literature   for   polymeric   ligands   and   metals   investigated   in   self-­‐

healing  metallosupramolecular  polymers.5,  14-­‐23  

 

 

 

 

ON

NN

NN

O N

NN

NN

x yn

NH

NH

O O NH

NH

O NH

O O O O O

NH

O O NH

NH

O O O1 2 1 4 1 3 1

O4

1 OCNNCO

2N O NO₂

OHHO

3

4

NNN N N N

NHO

OH

HOO

H

NN N

OO ORn m

n m

/Zn²⁺/or/La³⁺ /Fe²⁺/or/Cd²⁺

/Zn²⁺/or/Eu³⁺

O O

OOO

O

/Fe²⁺

a) b) c)

d)

NR₂

NR₂R₂N

R₂N

NNN NO O

NH

R₂N

R₂N

NNO

O

HN

O

NR₂

NR₂

N NO

e)

/Pd²⁺/or/Pt²⁺

Supramolecular polymer networks

105

6

Page 113: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  discussions  

 

Scheme  1:  Synthesis  of  copolymer  poly(MA-­‐co-­‐VIm).  

In  order  to  obtain  a  cross-­‐linked  coordination  network,  methyl  acrylate  was  copolymerized  

with   N-­‐vinyl   imidazole   (Scheme   1)   via   Cu   catalyzed   single   electron   transfer   living   radical  

polymerization   (SET-­‐LRP).24  The  mechanism  of  SET-­‐LRP   is  well  established  and   it   results   in  

low  polydispersity  polymers  with  predefined  molecular  weights.  Copolymer  was  synthesized  

with  an  initial  monomer  feed  ratio  in  solution  for  VIm:  MA  of  5:95.    The  fraction  of  VIm  in  

the   final   polymer  was   determined  with   1H  NMR,   using  methyl   peaks   of   initiator   (ethyl   2-­‐

bromo-­‐2-­‐methylpropanoate)  as  internal  reference  (Figure  2).  The  poly(MA-­‐co-­‐VIm)  had  ~4%  

VIm   units,   indicating   similar   reactivity   ratios   for   monomers.25   Molecular   weight   of   the  

copolymer  was  determined  as  Mn  =  25  kDa  by  1H  NMR  and  confirmed  by  MALDI  (Figure  3)  to  

give   an   average   of   n   imidazole   units   per   polymer   chain.   The   molecular   weight   of   the  

polymer  is  also  consistent  with  the  results  previously  reported  the  homo-­‐  polymerizations  of  

MA  by  SET-­‐LRP.  This  shows  that  the  presence  of  VIm  does  not  inhibit  or  slow  down  the  SET-­‐

LRP.        

 

O

OBr

O

O+ +O

N

NOO

Ox

Cu)wireDMSO

RTN

N

1 2 3

y n∗

 

Figure  2:  1H  NMR  spectrum  of  poly(MA-­‐co-­‐VIm)  in  CDCl3    

 

 

Figure  3:  MALDI-­‐TOF  spectrum  of  poly(MA-­‐co-­‐VIm)    

A   3.7   mM   solution   of   copolymer   poly(MA-­‐co-­‐VIm)   in   MeCN   was   mixed   with   either  

PdCl2(CH3CN)2  or  Cu(CH3CN)4PF6  dissolved  in  CH2Cl2.  Upon  mixing,  a  gel  was  formed  in  both  

cases,  as  depicted  in  Figure  4  for  the  Pd  complex.  

Chapter 6

106

Page 114: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Results  and  discussions  

 

Scheme  1:  Synthesis  of  copolymer  poly(MA-­‐co-­‐VIm).  

In  order  to  obtain  a  cross-­‐linked  coordination  network,  methyl  acrylate  was  copolymerized  

with   N-­‐vinyl   imidazole   (Scheme   1)   via   Cu   catalyzed   single   electron   transfer   living   radical  

polymerization   (SET-­‐LRP).24  The  mechanism  of  SET-­‐LRP   is  well  established  and   it   results   in  

low  polydispersity  polymers  with  predefined  molecular  weights.  Copolymer  was  synthesized  

with  an  initial  monomer  feed  ratio  in  solution  for  VIm:  MA  of  5:95.    The  fraction  of  VIm  in  

the   final   polymer  was   determined  with   1H  NMR,   using  methyl   peaks   of   initiator   (ethyl   2-­‐

bromo-­‐2-­‐methylpropanoate)  as  internal  reference  (Figure  2).  The  poly(MA-­‐co-­‐VIm)  had  ~4%  

VIm   units,   indicating   similar   reactivity   ratios   for   monomers.25   Molecular   weight   of   the  

copolymer  was  determined  as  Mn  =  25  kDa  by  1H  NMR  and  confirmed  by  MALDI  (Figure  3)  to  

give   an   average   of   n   imidazole   units   per   polymer   chain.   The   molecular   weight   of   the  

polymer  is  also  consistent  with  the  results  previously  reported  the  homo-­‐  polymerizations  of  

MA  by  SET-­‐LRP.  This  shows  that  the  presence  of  VIm  does  not  inhibit  or  slow  down  the  SET-­‐

LRP.        

 

O

OBr

O

O+ +O

N

NOO

Ox

Cu)wireDMSO

RTN

N

1 2 3

y n∗

 

Figure  2:  1H  NMR  spectrum  of  poly(MA-­‐co-­‐VIm)  in  CDCl3    

 

 

Figure  3:  MALDI-­‐TOF  spectrum  of  poly(MA-­‐co-­‐VIm)    

A   3.7   mM   solution   of   copolymer   poly(MA-­‐co-­‐VIm)   in   MeCN   was   mixed   with   either  

PdCl2(CH3CN)2  or  Cu(CH3CN)4PF6  dissolved  in  CH2Cl2.  Upon  mixing,  a  gel  was  formed  in  both  

cases,  as  depicted  in  Figure  4  for  the  Pd  complex.  

Supramolecular polymer networks

107

6

Page 115: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  4:  Supramolecular  gel  formed  upon  addition  of  PdCl2(MeCN)2  in  MeCN  

 

Figure  5:  Polymer  films  prepared  from  the  gels  by  evaporating  solvent  a)  poly(MA-­‐co-­‐VIm)-­‐PdCl2  and  

b)  poly(MA-­‐co-­‐VIm)-­‐CuPF6  

Polymer  films    (Figure  5)  were  prepared  from  the  gels  by  slowly  evaporating  solvents  on  a  

Teflon  mold.  Dumbbell   shaped  polymer   films  were   subjected   to   tensile   testing  at   a   strain  

rate  of  10-­‐3  s-­‐1.  The  Pd  containing  polymer  has  a  Young’s  modulus  of  2  MPa  that   is  higher  

than  that  of  Cu  containing  polymer  (0.25  MPa).  This  may  be  attributed  to  the  higher  binding  

affinity  of  Pd  for  imidazole.  The  force  induced  bond  scission  has  been  already  shown  to  take  

place  on  the  weakest  bond.3  Therefore;  the  strength  of  the  metal  ligand  coordination  bond  

determines  the  mechanical  properties  for  the  film.    

 

 

Figure   6:   Tensile   tests   for   poly(MA-­‐co-­‐VIm)-­‐PdCl2   and   poly(MA-­‐co-­‐VIm)-­‐CuPF6   before   and   after  

breaking  the  polymer  films.  

Polymer   films   were   tested   for   their   self-­‐healing   properties   at   room   temperature.   Broken  

pieces  were  brought  in  contact  for  2  h  right  after  rupture  (<5  mins).  In  the  time  scale  of  the  

experiment,   Pd   bearing   polymer   film   did   not   heal   significantly   whereas   Cu   containing  

polymer  recovered  approximately  half  of  its  initial  tensile  strength  (Figure  6).  

Conclusions    

A  copolymer  of  methyl  acrylate  (MA)  with  ~4%  vinylimidazole  (VIm)  groups  was  synthesized.  

The  Imidazole  pendant  groups  in  the  polymer  serve  as  metal  coordination  sites  and  give  rise  

to  gelation  when  Pd(II)  or  Cu(I)  salts  are  added  in  solution.  Freestanding  films  were  prepared  

with   these   metallo-­‐supramolecular   polymers.   Metal-­‐ligand   bonds   in   these   polymer   films  

break   upon   crack   formation   and   yield   free   ligand   and   coordinatively   unsaturated   metal  

center.   These   reactive   groups   may   initiate   the   self-­‐healing   property   once   surfaces   are  

brought  in  contact.  Healing  efficiency  of  these  polymers  can  be  increased  with  fine-­‐tuning  of  

the  ligand  metal  ratios  or  healing  time  and  temperatures.    

 

 

 

 

 

 

 

Chapter 6

108

Page 116: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

 

Figure  4:  Supramolecular  gel  formed  upon  addition  of  PdCl2(MeCN)2  in  MeCN  

 

Figure  5:  Polymer  films  prepared  from  the  gels  by  evaporating  solvent  a)  poly(MA-­‐co-­‐VIm)-­‐PdCl2  and  

b)  poly(MA-­‐co-­‐VIm)-­‐CuPF6  

Polymer  films    (Figure  5)  were  prepared  from  the  gels  by  slowly  evaporating  solvents  on  a  

Teflon  mold.  Dumbbell   shaped  polymer   films  were   subjected   to   tensile   testing  at   a   strain  

rate  of  10-­‐3  s-­‐1.  The  Pd  containing  polymer  has  a  Young’s  modulus  of  2  MPa  that   is  higher  

than  that  of  Cu  containing  polymer  (0.25  MPa).  This  may  be  attributed  to  the  higher  binding  

affinity  of  Pd  for  imidazole.  The  force  induced  bond  scission  has  been  already  shown  to  take  

place  on  the  weakest  bond.3  Therefore;  the  strength  of  the  metal  ligand  coordination  bond  

determines  the  mechanical  properties  for  the  film.    

 

 

Figure   6:   Tensile   tests   for   poly(MA-­‐co-­‐VIm)-­‐PdCl2   and   poly(MA-­‐co-­‐VIm)-­‐CuPF6   before   and   after  

breaking  the  polymer  films.  

Polymer   films   were   tested   for   their   self-­‐healing   properties   at   room   temperature.   Broken  

pieces  were  brought  in  contact  for  2  h  right  after  rupture  (<5  mins).  In  the  time  scale  of  the  

experiment,   Pd   bearing   polymer   film   did   not   heal   significantly   whereas   Cu   containing  

polymer  recovered  approximately  half  of  its  initial  tensile  strength  (Figure  6).  

Conclusions    

A  copolymer  of  methyl  acrylate  (MA)  with  ~4%  vinylimidazole  (VIm)  groups  was  synthesized.  

The  Imidazole  pendant  groups  in  the  polymer  serve  as  metal  coordination  sites  and  give  rise  

to  gelation  when  Pd(II)  or  Cu(I)  salts  are  added  in  solution.  Freestanding  films  were  prepared  

with   these   metallo-­‐supramolecular   polymers.   Metal-­‐ligand   bonds   in   these   polymer   films  

break   upon   crack   formation   and   yield   free   ligand   and   coordinatively   unsaturated   metal  

center.   These   reactive   groups   may   initiate   the   self-­‐healing   property   once   surfaces   are  

brought  in  contact.  Healing  efficiency  of  these  polymers  can  be  increased  with  fine-­‐tuning  of  

the  ligand  metal  ratios  or  healing  time  and  temperatures.    

 

 

 

 

 

 

 

Supramolecular polymer networks

109

6

Page 117: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Experimental  

General  

Dry   tetrahydrofuran   (THF,  HPLC  grade)  was  degassed  with   argon  and  purified  by  passage   through  

activated  alumina  solvent  column.  DMSO  was  dried  over  4Å  molecular  sieves  for  at  least  12h  prior  to  

use.  A  Varian  400MR  or  a  Varian  Mercury  400  spectrometer  was  used  to  record  1H  NMR  (400  MHz).  

Chemical  shifts  are  reported  in  ppm  and  referenced  to  tetramethylsilane  or  solvent.  Methyl  acrylate  

and   N-­‐vinyl   imidazole   were   purchased   from   commercial   sources   and   filtered   through   neutral  

Alumina  before  polymerizations.    

Synthesis  of  poly(MA-­‐co-­‐VIm):  

Initiator  (180  mg,  0.10  mmol),  Me6TREN  (46.5  mg,  0.20  mmol),  methyl  acrylate  (5.0  mL)  and  N-­‐vinyl  

imidazole   (0.7  ml)  were   dissolved   in   DMSO   and   degassed   by   three   freeze-­‐pump-­‐thaw   cycles,   and  

purged  with  Ar  prior  to  polymerization.  Me6TREN  was  weighed  into  aluminum  weigh  boat  and  added  

to  a  Schlenk  tube  with  the  boat.  Copper  wire  was  wrapped  around  a  magnetic  stirring  bar  and  used  

as  the  source  of  Cu(0)  catalyst.  The  Schlenk  tube  was  stirred  at  room  temperature  in  a  water  bath  at  

25oC   for   the   period   of   polymerization   (30  mins).   Upon   completion   of   the   reaction,   the   tube   was  

opened  to  air  and  THF  (10  mL)  was  added  to  the  viscous  solution.  The  reaction  was  filtered  through  

a   plug   of   basic   alumina   using   THF   to   remove   Cu(0)   particles   and   added   onto   stirring  MeOH.   The  

polymer   was   centrifuged   out   from   the   turbid   liquid   mixture   and   left   under   vacuum   overnight.  

Molecular  weight  of  polymer  was  determined  by  MALDI-­‐TOF  and  1H  NMR  as  25  kDa.  

 

 

 

 

 

 

 

 

 

References  

(1)     Binder,  W.  H.;   Zirbs,   R.   In  Hydrogen   Bonded   Polymers;   Binder,  W.,   Ed.;   Advances   in   Polymer  

Science;  Springer  Berlin  Heidelberg,  2006;  pp  1–78.  

(2)     Aida,  T.;  Meijer,  E.  W.;  Stupp,  S.  I.  Science  2012,  335  (6070),  813–817.  

(3)     Sijbesma,  R.  P.;  Beijer,  F.  H.;  Brunsveld,  L.;  Folmer,  B.  J.  B.;  Hirschberg,  J.  H.  K.  K.;  Lange,  R.  F.  

M.;  Lowe,  J.  K.  L.;  Meijer,  E.  W.  Science  1997,  278  (5343),  1601–1604.  

(4)     Xu,  D.;  Craig,  S.  L.  Macromolecules  2011,  44  (13),  5465–5472.  

(5)     Yount,  W.  C.;  Juwarker,  H.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2003,  125  (50),  15302–15303.  

(6)     Cordier,  P.;  Tournilhac,  F.;  Soulié-­‐Ziakovic,  C.;  Leibler,  L.  Nature  2008,  451  (7181),  977–980.  

(7)     Murphy,  E.  B.;  Wudl,  F.  Prog.  Polym.  Sci.  2010,  35  (1–2),  223–251.  

(8)     Liu,  Y.-­‐L.;  Chuo,  T.-­‐W.  Polym.  Chem.  2013,  4  (7),  2194–2205.  

(9)     Chen,  Y.;  Kushner,  A.  M.;  Williams,  G.  A.;  Guan,  Z.  Nat.  Chem.  2012,  4  (6),  467–472.  

(10)     Chen,  Y.;  Guan,  Z.  Chem.  Commun.  2014,  50  (74),  10868–10870.  

(11)     Jr,  S.  J.  K.;  Ward,  T.  C.;  Oyetunji,  Z.  Mech.  Adv.  Mater.  Struct.  2007,  14  (5),  391–397.  

(12)     Burattini,   S.;   Colquhoun,   H.  M.;   Fox,   J.   D.;   Friedmann,   D.;   Greenland,   B.  W.;   Harris,   P.   J.   F.;  

Hayes,  W.;  Mackay,  M.  E.;  Rowan,  S.  J.  Chem.  Commun.  2009,  No.  44,  6717–6719.  

(13)     Yang,   X.;   Yu,   H.;  Wang,   L.;   Tong,   R.;   Akram,  M.;   Chen,   Y.;   Zhai,   X.   Soft   Matter   2015,   11   (7),  

1242–1252.  

(14)     Burnworth,  M.;   Tang,   L.;   Kumpfer,   J.   R.;  Duncan,  A.   J.;   Beyer,   F.   L.;   Fiore,  G.   L.;   Rowan,   S.   J.;  

Weder,  C.  Nature  2011,  472  (7343),  334–337.  

(15)     Yuan,   J.;   Fang,   X.;   Zhang,   L.;  Hong,  G.;   Lin,   Y.;   Zheng,  Q.;   Xu,   Y.;   Ruan,   Y.;  Weng,  W.;   Xia,  H.;  

Chen,  G.  J.  Mater.  Chem.  2012,  22  (23),  11515–11522.  

(16)     Hong,  G.;  Zhang,  H.;  Lin,  Y.;  Chen,  Y.;  Xu,  Y.;  Weng,  W.;  Xia,  H.  Macromolecules  2013,  46  (21),  

8649–8656.  

(17)     Bode,  S.;  Zedler,  L.;  Schacher,  F.  H.;  Dietzek,  B.;  Schmitt,  M.;  Popp,  J.;  Hager,  M.  D.;  Schubert,  U.  

S.  Adv.  Mater.  2013,  25  (11),  1634–1638.  

(18)     Mozhdehi,  D.;  Ayala,   S.;   Cromwell,  O.  R.;  Guan,   Z.   J.   Am.   Chem.   Soc.  2014,  136   (46),   16128–

16131.  

(19)     Sandmann,  B.;  Happ,  B.;   Kupfer,   S.;   Schacher,   F.  H.;  Hager,  M.  D.;   Schubert,  U.   S.  Macromol.  

Rapid  Commun.  2015,  36  (7),  604–609.  

(20)     Harrington,   M.   J.;   Masic,   A.;   Holten-­‐Andersen,   N.;   Waite,   J.   H.;   Fratzl,   P.   Science   2010,   328  

(5975),  216–220.  

(21)     Holten-­‐Andersen,  N.;  Harrington,  M.  J.;  Birkedal,  H.;  Lee,  B.  P.;  Messersmith,  P.  B.;  Lee,  K.  Y.  C.;  

Waite,  J.  H.  Proc.  Natl.  Acad.  Sci.  2011,  108  (7),  2651–2655.  

Chapter 6

110

Page 118: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Experimental  

General  

Dry   tetrahydrofuran   (THF,  HPLC  grade)  was  degassed  with   argon  and  purified  by  passage   through  

activated  alumina  solvent  column.  DMSO  was  dried  over  4Å  molecular  sieves  for  at  least  12h  prior  to  

use.  A  Varian  400MR  or  a  Varian  Mercury  400  spectrometer  was  used  to  record  1H  NMR  (400  MHz).  

Chemical  shifts  are  reported  in  ppm  and  referenced  to  tetramethylsilane  or  solvent.  Methyl  acrylate  

and   N-­‐vinyl   imidazole   were   purchased   from   commercial   sources   and   filtered   through   neutral  

Alumina  before  polymerizations.    

Synthesis  of  poly(MA-­‐co-­‐VIm):  

Initiator  (180  mg,  0.10  mmol),  Me6TREN  (46.5  mg,  0.20  mmol),  methyl  acrylate  (5.0  mL)  and  N-­‐vinyl  

imidazole   (0.7  ml)  were   dissolved   in   DMSO   and   degassed   by   three   freeze-­‐pump-­‐thaw   cycles,   and  

purged  with  Ar  prior  to  polymerization.  Me6TREN  was  weighed  into  aluminum  weigh  boat  and  added  

to  a  Schlenk  tube  with  the  boat.  Copper  wire  was  wrapped  around  a  magnetic  stirring  bar  and  used  

as  the  source  of  Cu(0)  catalyst.  The  Schlenk  tube  was  stirred  at  room  temperature  in  a  water  bath  at  

25oC   for   the   period   of   polymerization   (30  mins).   Upon   completion   of   the   reaction,   the   tube   was  

opened  to  air  and  THF  (10  mL)  was  added  to  the  viscous  solution.  The  reaction  was  filtered  through  

a   plug   of   basic   alumina   using   THF   to   remove   Cu(0)   particles   and   added   onto   stirring  MeOH.   The  

polymer   was   centrifuged   out   from   the   turbid   liquid   mixture   and   left   under   vacuum   overnight.  

Molecular  weight  of  polymer  was  determined  by  MALDI-­‐TOF  and  1H  NMR  as  25  kDa.  

 

 

 

 

 

 

 

 

 

References  

(1)     Binder,  W.  H.;   Zirbs,   R.   In  Hydrogen   Bonded   Polymers;   Binder,  W.,   Ed.;   Advances   in   Polymer  

Science;  Springer  Berlin  Heidelberg,  2006;  pp  1–78.  

(2)     Aida,  T.;  Meijer,  E.  W.;  Stupp,  S.  I.  Science  2012,  335  (6070),  813–817.  

(3)     Sijbesma,  R.  P.;  Beijer,  F.  H.;  Brunsveld,  L.;  Folmer,  B.  J.  B.;  Hirschberg,  J.  H.  K.  K.;  Lange,  R.  F.  

M.;  Lowe,  J.  K.  L.;  Meijer,  E.  W.  Science  1997,  278  (5343),  1601–1604.  

(4)     Xu,  D.;  Craig,  S.  L.  Macromolecules  2011,  44  (13),  5465–5472.  

(5)     Yount,  W.  C.;  Juwarker,  H.;  Craig,  S.  L.  J.  Am.  Chem.  Soc.  2003,  125  (50),  15302–15303.  

(6)     Cordier,  P.;  Tournilhac,  F.;  Soulié-­‐Ziakovic,  C.;  Leibler,  L.  Nature  2008,  451  (7181),  977–980.  

(7)     Murphy,  E.  B.;  Wudl,  F.  Prog.  Polym.  Sci.  2010,  35  (1–2),  223–251.  

(8)     Liu,  Y.-­‐L.;  Chuo,  T.-­‐W.  Polym.  Chem.  2013,  4  (7),  2194–2205.  

(9)     Chen,  Y.;  Kushner,  A.  M.;  Williams,  G.  A.;  Guan,  Z.  Nat.  Chem.  2012,  4  (6),  467–472.  

(10)     Chen,  Y.;  Guan,  Z.  Chem.  Commun.  2014,  50  (74),  10868–10870.  

(11)     Jr,  S.  J.  K.;  Ward,  T.  C.;  Oyetunji,  Z.  Mech.  Adv.  Mater.  Struct.  2007,  14  (5),  391–397.  

(12)     Burattini,   S.;   Colquhoun,   H.  M.;   Fox,   J.   D.;   Friedmann,   D.;   Greenland,   B.  W.;   Harris,   P.   J.   F.;  

Hayes,  W.;  Mackay,  M.  E.;  Rowan,  S.  J.  Chem.  Commun.  2009,  No.  44,  6717–6719.  

(13)     Yang,   X.;   Yu,   H.;  Wang,   L.;   Tong,   R.;   Akram,  M.;   Chen,   Y.;   Zhai,   X.   Soft   Matter   2015,   11   (7),  

1242–1252.  

(14)     Burnworth,  M.;   Tang,   L.;   Kumpfer,   J.   R.;  Duncan,  A.   J.;   Beyer,   F.   L.;   Fiore,  G.   L.;   Rowan,   S.   J.;  

Weder,  C.  Nature  2011,  472  (7343),  334–337.  

(15)     Yuan,   J.;   Fang,   X.;   Zhang,   L.;  Hong,  G.;   Lin,   Y.;   Zheng,  Q.;   Xu,   Y.;   Ruan,   Y.;  Weng,  W.;   Xia,  H.;  

Chen,  G.  J.  Mater.  Chem.  2012,  22  (23),  11515–11522.  

(16)     Hong,  G.;  Zhang,  H.;  Lin,  Y.;  Chen,  Y.;  Xu,  Y.;  Weng,  W.;  Xia,  H.  Macromolecules  2013,  46  (21),  

8649–8656.  

(17)     Bode,  S.;  Zedler,  L.;  Schacher,  F.  H.;  Dietzek,  B.;  Schmitt,  M.;  Popp,  J.;  Hager,  M.  D.;  Schubert,  U.  

S.  Adv.  Mater.  2013,  25  (11),  1634–1638.  

(18)     Mozhdehi,  D.;  Ayala,   S.;   Cromwell,  O.  R.;  Guan,   Z.   J.   Am.   Chem.   Soc.  2014,  136   (46),   16128–

16131.  

(19)     Sandmann,  B.;  Happ,  B.;   Kupfer,   S.;   Schacher,   F.  H.;  Hager,  M.  D.;   Schubert,  U.   S.  Macromol.  

Rapid  Commun.  2015,  36  (7),  604–609.  

(20)     Harrington,   M.   J.;   Masic,   A.;   Holten-­‐Andersen,   N.;   Waite,   J.   H.;   Fratzl,   P.   Science   2010,   328  

(5975),  216–220.  

(21)     Holten-­‐Andersen,  N.;  Harrington,  M.  J.;  Birkedal,  H.;  Lee,  B.  P.;  Messersmith,  P.  B.;  Lee,  K.  Y.  C.;  

Waite,  J.  H.  Proc.  Natl.  Acad.  Sci.  2011,  108  (7),  2651–2655.  

Supramolecular polymer networks

111

6

Page 119: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

(22)     Terech,  P.;  Yan,  M.;  Maréchal,  M.;  Royal,  G.;  Galvez,  J.;  Velu,  S.  K.  P.  Phys.  Chem.  Chem.  Phys.  

2013,  15  (19),  7338–7344.  

(23)     Kersey,  F.  R.;  Loveless,  D.  M.;  Craig,  S.  L.  J.  R.  Soc.  Interface  2007,  4  (13),  373–380.  

(24)     Percec,   V.;   Guliashvili,   T.;   Ladislaw,   J.   S.;   Wistrand,   A.;   Stjerndahl,   A.;   Sienkowska,   M.   J.;  

Monteiro,  M.  J.;  Sahoo,  S.  J.  Am.  Chem.  Soc.  2006,  128  (43),  14156–14165.  

(25)     Pekel,  N.;  Rzaev,  Z.  M.  O.;  Güven,  O.  Macromol.  Chem.  Phys.  2004,  205  (8),  1088–1095.  

 

Summary  

Mechanochemical  Scission  of  Transition  Metal-­‐Ligand  Bonds  in  Coordination  

Polymers  

Mechanical  activation  of  chemical  bonds  in  polymers  offers  opportunities  for  a  broad  range  

of   uses   such   as   mechanically   activated   catalytic   activity,   and   the   formation   of   block  

copolymers.  It  has  been  shown  that  mechanical  work  done  by  an  external  force  lowers  the  

energy  barrier  for  bond  dissociation  to  such  an  extent  that  thermal  fluctuations  can  exceed  

this   barrier   at   room   temperature.   One   of   the   most   efficient   ways   to   exert   force   on   a  

molecule  in  solution  is  sonication.  Polymers  with  sufficient  molecular  weight  undergo  chain  

scission   when   they   are   sonicated   in   solution   due   to   elongational   stresses   experienced  

around  collapsing  bubbles.  As  a  consequence  of  the  nature  of  the  stress  field,  chain  scission  

occurs  preferentially  around  the  chain  midpoint.  

The  main  aim  of  this  thesis  is  to  gain  a  better  understanding  of  the  fundamental  processes  

and  mechanisms  underlying  mechanochemical  chain  scission   in  organometallic  complexes.  

Therefore,   the   use   of   mechanical   force   to   break   coordination   bonds   between   transition  

metals  and  ligands  was  investigated.  

Ultrasound  induced  chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  and  Pt(NHC-­‐pTHF)2Cl2,  coordination  

complexes   of   PdII   and   PtII   with   polytetrahydrofuran   functionalized  N-­‐heterocyclic   carbene  

(NHC)  ligands,  was  investigated  in  Chapter  2.  Application  of  force  in  solution  by  the  use  of  

ultrasound   resulted   in   selective   chain   scission   at   the   metal–ligand   coordination   bond.  

Scission   in   coordination   bond   is   reversible;   however,   scission   products  were   trapped   and  

monitored   by   NMR   and   GPC.   Sonicating   a   series   of   polymer   complexes   with   different  

molecular  weights  (MW),  it  was  found  that  the  chain  scission  rate  is  directly  proportional  to  

the  MW  of   the  polymers  with  above   the   limiting  molecular  weight.  Comparing   scission  of  

Palladium   and   Platinum   containing   polymers   showed   the   influence   of   ligand   dissociation  

energy  on  mechanochemical   response  of   the  coordination  polymers.  Pd-­‐NHC  (195  kJ/mol)  

and   Pt-­‐NHC   (245   kJ/mol)   have   limiting   MWs   as   20   and   22   kDa   respectively.   The   force  

required   to   break   the   bond,   Fmax   ,   was   calculated   using   COGEF   method.   Two   important  

Chapter 6

112

Page 120: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

(22)     Terech,  P.;  Yan,  M.;  Maréchal,  M.;  Royal,  G.;  Galvez,  J.;  Velu,  S.  K.  P.  Phys.  Chem.  Chem.  Phys.  

2013,  15  (19),  7338–7344.  

(23)     Kersey,  F.  R.;  Loveless,  D.  M.;  Craig,  S.  L.  J.  R.  Soc.  Interface  2007,  4  (13),  373–380.  

(24)     Percec,   V.;   Guliashvili,   T.;   Ladislaw,   J.   S.;   Wistrand,   A.;   Stjerndahl,   A.;   Sienkowska,   M.   J.;  

Monteiro,  M.  J.;  Sahoo,  S.  J.  Am.  Chem.  Soc.  2006,  128  (43),  14156–14165.  

(25)     Pekel,  N.;  Rzaev,  Z.  M.  O.;  Güven,  O.  Macromol.  Chem.  Phys.  2004,  205  (8),  1088–1095.  

 

Summary  

Mechanochemical  Scission  of  Transition  Metal-­‐Ligand  Bonds  in  Coordination  

Polymers  

Mechanical  activation  of  chemical  bonds  in  polymers  offers  opportunities  for  a  broad  range  

of   uses   such   as   mechanically   activated   catalytic   activity,   and   the   formation   of   block  

copolymers.  It  has  been  shown  that  mechanical  work  done  by  an  external  force  lowers  the  

energy  barrier  for  bond  dissociation  to  such  an  extent  that  thermal  fluctuations  can  exceed  

this   barrier   at   room   temperature.   One   of   the   most   efficient   ways   to   exert   force   on   a  

molecule  in  solution  is  sonication.  Polymers  with  sufficient  molecular  weight  undergo  chain  

scission   when   they   are   sonicated   in   solution   due   to   elongational   stresses   experienced  

around  collapsing  bubbles.  As  a  consequence  of  the  nature  of  the  stress  field,  chain  scission  

occurs  preferentially  around  the  chain  midpoint.  

The  main  aim  of  this  thesis  is  to  gain  a  better  understanding  of  the  fundamental  processes  

and  mechanisms  underlying  mechanochemical  chain  scission   in  organometallic  complexes.  

Therefore,   the   use   of   mechanical   force   to   break   coordination   bonds   between   transition  

metals  and  ligands  was  investigated.  

Ultrasound  induced  chain  scission  in  Pd(NHC-­‐pTHF)2Cl2  and  Pt(NHC-­‐pTHF)2Cl2,  coordination  

complexes   of   PdII   and   PtII   with   polytetrahydrofuran   functionalized  N-­‐heterocyclic   carbene  

(NHC)  ligands,  was  investigated  in  Chapter  2.  Application  of  force  in  solution  by  the  use  of  

ultrasound   resulted   in   selective   chain   scission   at   the   metal–ligand   coordination   bond.  

Scission   in   coordination   bond   is   reversible;   however,   scission   products  were   trapped   and  

monitored   by   NMR   and   GPC.   Sonicating   a   series   of   polymer   complexes   with   different  

molecular  weights  (MW),  it  was  found  that  the  chain  scission  rate  is  directly  proportional  to  

the  MW  of   the  polymers  with  above   the   limiting  molecular  weight.  Comparing   scission  of  

Palladium   and   Platinum   containing   polymers   showed   the   influence   of   ligand   dissociation  

energy  on  mechanochemical   response  of   the  coordination  polymers.  Pd-­‐NHC  (195  kJ/mol)  

and   Pt-­‐NHC   (245   kJ/mol)   have   limiting   MWs   as   20   and   22   kDa   respectively.   The   force  

required   to   break   the   bond,   Fmax   ,   was   calculated   using   COGEF   method.   Two   important  

113

Page 121: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

conclusions  can  be  drawn  from  the  results  of  these  calculations:  (i)  Mlim  scales  with  Fmax  and  

(ii)  direction  of  pulling  and  the  choice  of  attachment  points  are  important  when  determining  

the  value  of  Fmax.  

Free   NHC   released   during   sonication   was   used   to   induce   chemiluminescence   via   proton  

abstraction   from  a  2-­‐coumaranone  derivative,  which  decomposes   via   a   chemiluminescent  

pathway   in   the   presence   of   oxygen.   Using   the   chemiluminescent   response,   the   rate   of  

ultrasound   induced   scission   and   molecular   weight   threshold   (Mlim)   for   mechanochemical  

chain  scission  were  determined  in  Chapter  3.  The  rate  constants  were  also  simulated  using  

GEPASI,   which   showed   that   the   rate-­‐determining   step   is   the   decomposition   of  

coumaranone.    

The  MW  dependence  of  chain  scission  rates  proves  that  the  bond  scission  observed  during  

sonication   is   mechanically   induced.   However,   it   is   still   arguable   that   trapping   agents  

promote  the  mechanochemical  chain  scission  by  decreasing  the  mechanical  stability  of  the  

ligand-­‐metal   coordination   bond.   In   Chapter   4   the   rates   of   mechanically   induced   ligand  

exchange   reactions   were   determined.   It   was   established   by   NMR   that   the   ligands   of   Pd-­‐

Imidazole  complexes  exchange  via  an  associative  pathway,  and  stoichiometric  complexes  of  

Pd-­‐Imidazole  do  not  show  ligand  exchange  because  free  ligands  are  absent.  However,  free  

ligands   released  by  ultrasound   initiate   ligand  exchange   in  a  mixture  of  polymeric  and   low  

molecular  weight   Pd(Im)2   complex.   The   rate   of   consumption   of   polymer   complex,   Pd(Im-­‐

pTHF)2Cl2,  was  monitored  by  GPC  and  kinetic  simulations  showed  that  the  ligand  association  

is  the  rate  determining  step.    

In  Chapter  5,  mechanochemically  induced  ligand  exchange  of  Pd(II)  complexes  was  used  to  

direct   the   formation   of   heterocomplexes.   Symmetric   complexes   with   high   and   low  

molecular   weight   polymer-­‐attached   ligands   were  mixed   in   solution   and   sonicated.  When  

one  of  the  complexes  has  a  molecular  weight  higher  than  Mlim  for  mechanochemical  chain  

scission,   while   the   other   is   smaller,   sonication   leads   to   the   directed   formation   of   a  

heterocomplex   with   two   different   ligands.   Two   Pd(Im)2Cl2   complexes   with   different  

molecular  weights  were  mixed  in  toluene  and  the  solution  was  subjected  to  ultrasound.  GPC  

traces   showed   that   the   main   polymer   fraction   after   sonication   was   the   heterocomplex.  

Furthermore,   pTHF   attached   NHC-­‐Pd   complexes,   Pd(NHC-­‐pTHF)2Cl2,   and  

poly(methylacrylate)  attached  pyridine-­‐Pd  complexes,  Pd(PyPMA)2Cl2,    were  also  mixed  with  

Pd(Im-­‐pTHF)2Cl2  and  sonicated.  In  all  cases,  ultrasound-­‐induced  formation  of  heterocomplex  

was  observed.    

In   Chapter   6,   preliminary   results   for   the   self-­‐healing   properties   of   polymer   films   of  

poly(methyl  acrylate)-­‐vinyl  imidazole  copolymers,  cross-­‐linked  by  ligand  metal  coordination  

are  described.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

Summary

114

Page 122: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

conclusions  can  be  drawn  from  the  results  of  these  calculations:  (i)  Mlim  scales  with  Fmax  and  

(ii)  direction  of  pulling  and  the  choice  of  attachment  points  are  important  when  determining  

the  value  of  Fmax.  

Free   NHC   released   during   sonication   was   used   to   induce   chemiluminescence   via   proton  

abstraction   from  a  2-­‐coumaranone  derivative,  which  decomposes   via   a   chemiluminescent  

pathway   in   the   presence   of   oxygen.   Using   the   chemiluminescent   response,   the   rate   of  

ultrasound   induced   scission   and   molecular   weight   threshold   (Mlim)   for   mechanochemical  

chain  scission  were  determined  in  Chapter  3.  The  rate  constants  were  also  simulated  using  

GEPASI,   which   showed   that   the   rate-­‐determining   step   is   the   decomposition   of  

coumaranone.    

The  MW  dependence  of  chain  scission  rates  proves  that  the  bond  scission  observed  during  

sonication   is   mechanically   induced.   However,   it   is   still   arguable   that   trapping   agents  

promote  the  mechanochemical  chain  scission  by  decreasing  the  mechanical  stability  of  the  

ligand-­‐metal   coordination   bond.   In   Chapter   4   the   rates   of   mechanically   induced   ligand  

exchange   reactions   were   determined.   It   was   established   by   NMR   that   the   ligands   of   Pd-­‐

Imidazole  complexes  exchange  via  an  associative  pathway,  and  stoichiometric  complexes  of  

Pd-­‐Imidazole  do  not  show  ligand  exchange  because  free  ligands  are  absent.  However,  free  

ligands   released  by  ultrasound   initiate   ligand  exchange   in  a  mixture  of  polymeric  and   low  

molecular  weight   Pd(Im)2   complex.   The   rate   of   consumption   of   polymer   complex,   Pd(Im-­‐

pTHF)2Cl2,  was  monitored  by  GPC  and  kinetic  simulations  showed  that  the  ligand  association  

is  the  rate  determining  step.    

In  Chapter  5,  mechanochemically  induced  ligand  exchange  of  Pd(II)  complexes  was  used  to  

direct   the   formation   of   heterocomplexes.   Symmetric   complexes   with   high   and   low  

molecular   weight   polymer-­‐attached   ligands   were  mixed   in   solution   and   sonicated.  When  

one  of  the  complexes  has  a  molecular  weight  higher  than  Mlim  for  mechanochemical  chain  

scission,   while   the   other   is   smaller,   sonication   leads   to   the   directed   formation   of   a  

heterocomplex   with   two   different   ligands.   Two   Pd(Im)2Cl2   complexes   with   different  

molecular  weights  were  mixed  in  toluene  and  the  solution  was  subjected  to  ultrasound.  GPC  

traces   showed   that   the   main   polymer   fraction   after   sonication   was   the   heterocomplex.  

Furthermore,   pTHF   attached   NHC-­‐Pd   complexes,   Pd(NHC-­‐pTHF)2Cl2,   and  

poly(methylacrylate)  attached  pyridine-­‐Pd  complexes,  Pd(PyPMA)2Cl2,    were  also  mixed  with  

Pd(Im-­‐pTHF)2Cl2  and  sonicated.  In  all  cases,  ultrasound-­‐induced  formation  of  heterocomplex  

was  observed.    

In   Chapter   6,   preliminary   results   for   the   self-­‐healing   properties   of   polymer   films   of  

poly(methyl  acrylate)-­‐vinyl  imidazole  copolymers,  cross-­‐linked  by  ligand  metal  coordination  

are  described.    

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

Summary

115

Page 123: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

    Curriculum  Vitae  

 

Abidin   Balan   was   born   on   August   27th,   1983   in   Malatya,  Turkey.   He   received   his   BSc   in   chemistry   from   Bilkent  University  in  2007.  He  then  joined  the  group  of  Prof.  Levent  Toppare   at   Middle   East   Technical   University   (METU)   to  pursue  an  MSc   in  chemistry.   In  his  MSc  studies,  he  focused  on   the   synthesis   of   the   Donor–Acceptor   type   conjugated  polymers,   mainly   benzotriazole   derivatives   for   the  application  in  electrochromic  devices  and  organic  solar  cells.  His  master   thesis  was  awarded  as   the   thesis  of   the   year   in  2009  (METU).  In  2010,  he  started  his  PhD  in  Macromolecular  and  Organic  Chemistry  Group  at  Eindhoven  University  of  Technology,  under  supervision  of  Prof.  Rint  Sijbesma.  The  most  important  results  of  this  research  are  described  in  this  thesis.  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 124: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

    Curriculum  Vitae  

 

Abidin   Balan   was   born   on   August   27th,   1983   in   Malatya,  Turkey.   He   received   his   BSc   in   chemistry   from   Bilkent  University  in  2007.  He  then  joined  the  group  of  Prof.  Levent  Toppare   at   Middle   East   Technical   University   (METU)   to  pursue  an  MSc   in  chemistry.   In  his  MSc  studies,  he  focused  on   the   synthesis   of   the   Donor–Acceptor   type   conjugated  polymers,   mainly   benzotriazole   derivatives   for   the  application  in  electrochromic  devices  and  organic  solar  cells.  His  master   thesis  was  awarded  as   the   thesis  of   the   year   in  2009  (METU).  In  2010,  he  started  his  PhD  in  Macromolecular  and  Organic  Chemistry  Group  at  Eindhoven  University  of  Technology,  under  supervision  of  Prof.  Rint  Sijbesma.  The  most  important  results  of  this  research  are  described  in  this  thesis.  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

117

Page 125: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

     

Acknowledgements  

First   and   foremost,   I   would   like   to   express   my   deepest   gratitude,   profound   respect   and  sincere   thanks   to  my   supervisor   prof.   dr.   Rint   Sijbesma   for   giving  me   the   opportunity   to  perform  my  PhD   study   in   his   group.   It  was   a   great   pleasure   and  privilege   for  me   to   be   a  member  of  your  group.  Your  scientific  guidance  always  helped  me  to  find  my  way  out  during  my  PhD.    

I  would  like  to  thank  my  co-­‐supervisor  prof.  dr.  Bert  Meijer  and  members  of  my  promotion  committee  prof.  dr.  Giancarlo  Cravotto,  prof.  dr.  Joost  Reek,  prof.  dr.   ir.  Emiel  Hensen  and  prof.  dr.  Albert  Schenning  for  their  valuable  comments  and  suggestions  on  my  thesis.    

I  would  like  to  thank  dr.  Bartek  Szyja  for  theoretical  calculations  reported  in  Chapter  2  and  Serge  Söntjens  for  helping  me  with  the  500  MHz  NMR  used  for  the  experiments  reported  in  Chapter  4.    

Ralph  Bovee  and  Xianwen  Lou,   I  would   like  to  thank  both  of  you  for  the  help  and  support  you  provided  in  the  analytical   lab.  I  appreciate  the  great  discussions  we  had  over  scientific  and  non-­‐scientific  topics.      

Ramon  and  Sascha,  thank  you  very  much  for  being  by  my  side  as  my  paranymphs.  Ramon,  I  always   enjoyed   your   companionship   in   and   outside   the   lab.   Having   thought   provoking  discussions  with  you  was   inspiring   for  me.  Sascha,   thank  you  for  being  such  a  good  friend  and  sharing  the  most  memorable  moments  of  my   life.   It  was  a   lot  of   fun  to  enjoy  Turkish  cuisine  with  you  and  have  you  in  the  big  Turkish  community  in  Eindhoven.          

Dear  Bob,  thank  you  for  being  the  one  who  provided  me  with  the  most  generous  support  during  my  PhD.  Having  you  in  the  next  office  made  me  feel  more  confident  as  I  always  knew  that  you  would  be  ready  to  help  whenever  I  needed.  I  think  we  should  keep  our  quarterly  meetings  going  together  with  Ramon.  Being  friend  with  you  guys  means  a  lot  to  me.      

My  former  office  mate,  Yulan  Chen;  you  are  a  very  good  friend  and  an  excellent  scientist.  Sometimes  I  miss  our  conversations  in  the  office  over  many  different  topics.  Thank  you  for  introducing  me  different  aspects  of  Chinese  culture.  I  hope  to  visit  you  in  China  soon.  Jody  Lugger;  you  have  taken  Yulan’s  seat  after  she  left  and  I  should  say  I  enjoyed  a   lot  being  in  the  same  office  with  you.  You  are  very  enthusiastic  and  always  eager  to  learn  person  and  I  wish  you  a  lot  of  success  for  the  rest  of  your  PhD.    

 

 

Page 126: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

     

Acknowledgements  

First   and   foremost,   I   would   like   to   express   my   deepest   gratitude,   profound   respect   and  sincere   thanks   to  my   supervisor   prof.   dr.   Rint   Sijbesma   for   giving  me   the   opportunity   to  perform  my  PhD   study   in   his   group.   It  was   a   great   pleasure   and  privilege   for  me   to   be   a  member  of  your  group.  Your  scientific  guidance  always  helped  me  to  find  my  way  out  during  my  PhD.    

I  would  like  to  thank  my  co-­‐supervisor  prof.  dr.  Bert  Meijer  and  members  of  my  promotion  committee  prof.  dr.  Giancarlo  Cravotto,  prof.  dr.  Joost  Reek,  prof.  dr.   ir.  Emiel  Hensen  and  prof.  dr.  Albert  Schenning  for  their  valuable  comments  and  suggestions  on  my  thesis.    

I  would  like  to  thank  dr.  Bartek  Szyja  for  theoretical  calculations  reported  in  Chapter  2  and  Serge  Söntjens  for  helping  me  with  the  500  MHz  NMR  used  for  the  experiments  reported  in  Chapter  4.    

Ralph  Bovee  and  Xianwen  Lou,   I  would   like  to  thank  both  of  you  for  the  help  and  support  you  provided  in  the  analytical   lab.  I  appreciate  the  great  discussions  we  had  over  scientific  and  non-­‐scientific  topics.      

Ramon  and  Sascha,  thank  you  very  much  for  being  by  my  side  as  my  paranymphs.  Ramon,  I  always   enjoyed   your   companionship   in   and   outside   the   lab.   Having   thought   provoking  discussions  with  you  was   inspiring   for  me.  Sascha,   thank  you  for  being  such  a  good  friend  and  sharing  the  most  memorable  moments  of  my   life.   It  was  a   lot  of   fun  to  enjoy  Turkish  cuisine  with  you  and  have  you  in  the  big  Turkish  community  in  Eindhoven.          

Dear  Bob,  thank  you  for  being  the  one  who  provided  me  with  the  most  generous  support  during  my  PhD.  Having  you  in  the  next  office  made  me  feel  more  confident  as  I  always  knew  that  you  would  be  ready  to  help  whenever  I  needed.  I  think  we  should  keep  our  quarterly  meetings  going  together  with  Ramon.  Being  friend  with  you  guys  means  a  lot  to  me.      

My  former  office  mate,  Yulan  Chen;  you  are  a  very  good  friend  and  an  excellent  scientist.  Sometimes  I  miss  our  conversations  in  the  office  over  many  different  topics.  Thank  you  for  introducing  me  different  aspects  of  Chinese  culture.  I  hope  to  visit  you  in  China  soon.  Jody  Lugger;  you  have  taken  Yulan’s  seat  after  she  left  and  I  should  say  I  enjoyed  a   lot  being  in  the  same  office  with  you.  You  are  very  enthusiastic  and  always  eager  to  learn  person  and  I  wish  you  a  lot  of  success  for  the  rest  of  your  PhD.    

 

 

119

Page 127: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Pauline,   Lech,  Katja   (and  Alexandra  Sophie  J),   Samanneh,  Erik,  Marta,  Veronique,  Marcel  K.,  Dana,  Berry,   Jessica,  Xiao,  Remco,  Marcel   S.,   Jurgen,  Olga,  Ralph,  Thuur,  Asish,  Abhijit,  Gajanan;   I  would   like  to  thank  all  of  you  for  being  so  friendly.  Having  such  great  scientists  like  you  around  me  made  my  life  in  MST  perfectly  pleasant.    

I  would   like   to   thank   Joke   Rediker,  Marjo   van  Hoof,   Jolanda   Spiering,   Bas   de  Waal,   Hans  Damen  and  Henk  Eding   for   their  great  help  during  my  PhD.  My   life  would  be  much  more  difficult  without  them.  

Now  comes  the  Turkish  part.    

Tezin   en   zor   bölümü   teşekkür   bölümüymüş.   Ne   çok   anı,   ne   güzel   dostluklar   biriktirmişiz.  Düşündükçe   daha   da   yoğunlaşıyor   duygular.   Keşke   şu   anda   hissettiklerim   kendiliğinden  dökülse  bu  sayfaya.    

Gönüllerin   muhtarı   Barış   Yağcı   ve   sevgili   başkan   Seda   Cantekin’le   başlamalıyım   sanırım.  Eindhoven’a   ilk   geldiğim   zamanlar   hissettiğim   sudan   çıkmış   balık   şaşkınlığını,   o   zamandan  bugüne  hiç  eksilmeyen  muhteşem  dostluğunuz  sayesinde  aştım.  Vefalı  arkadaş  tabiri  sanırım  en  çok   ikinize  uyar.     İkinizi  de  çok  seviyorum.  Barış  hocam,  hem  Hollanda’daki  arkadaşlığın  hem  de   İstanbul’daki  misafirperverliğin   için  ayrı  ayrı   teşekkür  ederim.  Şu  an   sana  sarılmak  istiyorum   J.   Sedacım,   seninle   vakit   geçirmek   benim   için   hep   çok   keyifli   oldu.   Nereye  gittiğimizi   bilmediğimiz   araba   yolculukları,   bilimsel   tartışmalar   veya   zaman   zaman   senden  sopa  yemek…  Samimiyetin  için  çok  teşekkürler.  

Başar   efendi…   Tartışmalar,   sohbetler,   geziler,   belgeseller,   filmler,   şarkılar,   türküler,   çaylar  kahveler,   çeşitli   içmeler…   Sana   bu   paylaşımlarımızın   her   biri   için   tek   tek   teşekkür   ederim.  Dostluğun   benim   için   çok   değerli.   Şimdi   bu   teşekkürden   sonra   benden   kurtulacağını  düşünüyorsan   yanılıyorsun   J.   Umuyorum   ki   çok   uzun   yıllar   daha   görüşmeye   devam  edeceğiz  sevgili  dostum.    

Ali   Can   kardeşim,   hayatı   sen   yaşıyorsun  J.   Sohbetlerimiz   ve   tartışmalarımız   bana   çok   şey  öğretti.   Doktoramın   en   sıkıntılı   zamanlarını   aydınlatıcı   belgesel   seanslarımız   ve   içtiğimiz  kırmızı   şaraplar   renklendirdi.   Bütün   bunların   yeri   ayrı   bende   ama   sana   en   çok   kalbinde  taşıdığın  özgürlük  ateşi  ve  her  türlü  otoriteyi  sorgulama  kararlılığı  için  teşekkür  ederim.        

Gökhan   hocam,   etrafına   yaydığın   neşe   için,   ne   olursa   olsun   pozitif   düşünüp   umut   dolu  olabildiğin   için   çok   teşekkür   ederim.   Umarım   ileride   sahip   olmayı   düşündüğüm   çocuk(lar)  senin  gibi  bir  akademisyenden  eğitim  alma  şansını  yakalar(lar).  Bu  satırlar  yazıya  dökülürken  hayatını  birleştirmek  üzere  olduğun  eşine  ve  sana  çok  mutlu  bir  hayat  dilerim.  

 

 

Kamiiiil…   Hocam   seninle   ilgili   anılarımızı   düşünürken,   çok   yağmur   yağdığı   bir   gün   evde  atomların  yapısı  hakkında  yaptığımız  uzun  ve  çok  keyifli  bir  sohbete  takıldı  aklım.  En  son  bir  su   hortumunun   iki   ucundan   karşılıklı   tutup   sallayarak   elektronların   hareketlerini   ve  orbitalleri   anlamaya   çalıştığımızı   hatırlıyorum   J.   Pek   çok   ufuk   açan,   algı   zorlayan  tartışmalarımız  oldu  seninle.  Bunların  yanında  hayattan  keyif  almak  için  de  neredeyse  hiçbir  fırsatı  kaçırmadık.  Tüm  bu  anılar  için  sana  çok  teşekkür  ederim.    

Teyzemin  evladı  Can  Nemlioğlu.  Gerçek  bir  sincan  delikanlısı  J.  Büyük  projelerin  insanı.  Her  ne   kadar   politik   tartışmalarda   ayrı   görüşleri   savunsak   da   seninle   tartışmaktan   her   zaman  büyük  keyif  aldım.  Seni  tanıdığım  için  çok  mutluyum.    

Egelstraat  11a  ve  bu  adresin  hayatıma  kattığı  güzel   insanlar.  Başar,  Kamil,  Can,  Barış,  güzel  sohbetiyle,   titizliğiyle,   çektiği   birbirinden   güzel   fotoğraflarıyla   İlkin,   bütün   bir   günü   tek  cümleyle   geçirebilecek   kadar   az   ve   öz   konuşan   Cengiz,   heyecanı   cümlelerinde   devrilen  Gözde,  hikayeleriyle  hepimizi   gözyaşlarına  boğan,   acıların   çocuğu  Alios,   tanıdığım  en   sakin  insan  olan  Öncü,  kavgaların  eşiğinde,  direnişçi  Kıvılcım...  Hepinizi  çok  seviyorum.      

Ceylan,  Sinan,  Banu,  Oktay,  Demet,  Ümit,  Roos,  Atilla,  Sena,  Bahadır,  Seda  L.,  Kubilay,  Tekin  Hocam,  Tayfun  abi  ve  ETÜD  sakinleri,  hepinize  Hollanda’daki  hayatıma  anlam  kattığınız  için  teşekkür  ederim.  Benim  için  çok  kıymetli  insanlarsınız.  Daha  pek  çok  partilerde,  kamplarda,  gezilerde,  toplantılarda,  gösteri-­‐protesto  ve  forumlarda  buluşmak  dileğiyle.    

Sevgili  Sema,  hayatımın  ciddi  bir  şekilde  değişmeye  başladığı,  pek  çok  önemli  adımın  atıldığı  bir   dönemde   tam   da   ihtiyacım   olan   arkadaşlığı   sunduğun   ve   en   unutulmaz   anlarda   tüm  mesafelere  rağmen  yanımda  olmayı  tercih  ettiğin  için  teşekkür  ederim.    

Hollanda   Gezi   Dayanışması   üyeleri,   Osman   Hocam,   Kıvılcım,   Hande,   Maral,   Leyla,   Yaşar,  Burhan,  Çiğdem,  Hakan,  Nil,  Levent,  Elif,  Eylem,  Derya  ve  adını  yazmadığım  herkese  sonsuz  teşekkürler.  Umut  dolu,  samimi,  sıcacık  insanlarsınız  ve  iyi  ki  varsınız.      

Annem,  babam  ve  Ceren...  Sizin  gibi  bir  ailem  olduğu  için  çok  şanslıyım.  Sevginizi,  desteğinizi  ve  güveninizi  benden  hiçbir  zaman  esirgemediğiniz  için  size  ne  kadar  teşekkür  etsem  azdır.    

ve   Gizem...     En   iyi   arkadaşım,   diğer   yarım...   Bu   tezin   içindeki   her   deney   hasretinle,   sana  kavuşmanın  hayaliyle  yapıldı.  Çok  bekledik,  zorlandık,  kızdık,  küstük  ama  demek  ki  en  çok  da  sevdik.  Yoksa  aşılır  mıydı  onca  sınır?  Ceylan’ın  da  dediği  gibi  beklediğime  değdi  J.    

 

Abidin

Acknowledgements

120

Page 128: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction

Pauline,   Lech,  Katja   (and  Alexandra  Sophie  J),   Samanneh,  Erik,  Marta,  Veronique,  Marcel  K.,  Dana,  Berry,   Jessica,  Xiao,  Remco,  Marcel   S.,   Jurgen,  Olga,  Ralph,  Thuur,  Asish,  Abhijit,  Gajanan;   I  would   like  to  thank  all  of  you  for  being  so  friendly.  Having  such  great  scientists  like  you  around  me  made  my  life  in  MST  perfectly  pleasant.    

I  would   like   to   thank   Joke   Rediker,  Marjo   van  Hoof,   Jolanda   Spiering,   Bas   de  Waal,   Hans  Damen  and  Henk  Eding   for   their  great  help  during  my  PhD.  My   life  would  be  much  more  difficult  without  them.  

Now  comes  the  Turkish  part.    

Tezin   en   zor   bölümü   teşekkür   bölümüymüş.   Ne   çok   anı,   ne   güzel   dostluklar   biriktirmişiz.  Düşündükçe   daha   da   yoğunlaşıyor   duygular.   Keşke   şu   anda   hissettiklerim   kendiliğinden  dökülse  bu  sayfaya.    

Gönüllerin   muhtarı   Barış   Yağcı   ve   sevgili   başkan   Seda   Cantekin’le   başlamalıyım   sanırım.  Eindhoven’a   ilk   geldiğim   zamanlar   hissettiğim   sudan   çıkmış   balık   şaşkınlığını,   o   zamandan  bugüne  hiç  eksilmeyen  muhteşem  dostluğunuz  sayesinde  aştım.  Vefalı  arkadaş  tabiri  sanırım  en  çok   ikinize  uyar.     İkinizi  de  çok  seviyorum.  Barış  hocam,  hem  Hollanda’daki  arkadaşlığın  hem  de   İstanbul’daki  misafirperverliğin   için  ayrı  ayrı   teşekkür  ederim.  Şu  an   sana  sarılmak  istiyorum   J.   Sedacım,   seninle   vakit   geçirmek   benim   için   hep   çok   keyifli   oldu.   Nereye  gittiğimizi   bilmediğimiz   araba   yolculukları,   bilimsel   tartışmalar   veya   zaman   zaman   senden  sopa  yemek…  Samimiyetin  için  çok  teşekkürler.  

Başar   efendi…   Tartışmalar,   sohbetler,   geziler,   belgeseller,   filmler,   şarkılar,   türküler,   çaylar  kahveler,   çeşitli   içmeler…   Sana   bu   paylaşımlarımızın   her   biri   için   tek   tek   teşekkür   ederim.  Dostluğun   benim   için   çok   değerli.   Şimdi   bu   teşekkürden   sonra   benden   kurtulacağını  düşünüyorsan   yanılıyorsun   J.   Umuyorum   ki   çok   uzun   yıllar   daha   görüşmeye   devam  edeceğiz  sevgili  dostum.    

Ali   Can   kardeşim,   hayatı   sen   yaşıyorsun  J.   Sohbetlerimiz   ve   tartışmalarımız   bana   çok   şey  öğretti.   Doktoramın   en   sıkıntılı   zamanlarını   aydınlatıcı   belgesel   seanslarımız   ve   içtiğimiz  kırmızı   şaraplar   renklendirdi.   Bütün   bunların   yeri   ayrı   bende   ama   sana   en   çok   kalbinde  taşıdığın  özgürlük  ateşi  ve  her  türlü  otoriteyi  sorgulama  kararlılığı  için  teşekkür  ederim.        

Gökhan   hocam,   etrafına   yaydığın   neşe   için,   ne   olursa   olsun   pozitif   düşünüp   umut   dolu  olabildiğin   için   çok   teşekkür   ederim.   Umarım   ileride   sahip   olmayı   düşündüğüm   çocuk(lar)  senin  gibi  bir  akademisyenden  eğitim  alma  şansını  yakalar(lar).  Bu  satırlar  yazıya  dökülürken  hayatını  birleştirmek  üzere  olduğun  eşine  ve  sana  çok  mutlu  bir  hayat  dilerim.  

 

 

Kamiiiil…   Hocam   seninle   ilgili   anılarımızı   düşünürken,   çok   yağmur   yağdığı   bir   gün   evde  atomların  yapısı  hakkında  yaptığımız  uzun  ve  çok  keyifli  bir  sohbete  takıldı  aklım.  En  son  bir  su   hortumunun   iki   ucundan   karşılıklı   tutup   sallayarak   elektronların   hareketlerini   ve  orbitalleri   anlamaya   çalıştığımızı   hatırlıyorum   J.   Pek   çok   ufuk   açan,   algı   zorlayan  tartışmalarımız  oldu  seninle.  Bunların  yanında  hayattan  keyif  almak  için  de  neredeyse  hiçbir  fırsatı  kaçırmadık.  Tüm  bu  anılar  için  sana  çok  teşekkür  ederim.    

Teyzemin  evladı  Can  Nemlioğlu.  Gerçek  bir  sincan  delikanlısı  J.  Büyük  projelerin  insanı.  Her  ne   kadar   politik   tartışmalarda   ayrı   görüşleri   savunsak   da   seninle   tartışmaktan   her   zaman  büyük  keyif  aldım.  Seni  tanıdığım  için  çok  mutluyum.    

Egelstraat  11a  ve  bu  adresin  hayatıma  kattığı  güzel   insanlar.  Başar,  Kamil,  Can,  Barış,  güzel  sohbetiyle,   titizliğiyle,   çektiği   birbirinden   güzel   fotoğraflarıyla   İlkin,   bütün   bir   günü   tek  cümleyle   geçirebilecek   kadar   az   ve   öz   konuşan   Cengiz,   heyecanı   cümlelerinde   devrilen  Gözde,  hikayeleriyle  hepimizi   gözyaşlarına  boğan,   acıların   çocuğu  Alios,   tanıdığım  en   sakin  insan  olan  Öncü,  kavgaların  eşiğinde,  direnişçi  Kıvılcım...  Hepinizi  çok  seviyorum.      

Ceylan,  Sinan,  Banu,  Oktay,  Demet,  Ümit,  Roos,  Atilla,  Sena,  Bahadır,  Seda  L.,  Kubilay,  Tekin  Hocam,  Tayfun  abi  ve  ETÜD  sakinleri,  hepinize  Hollanda’daki  hayatıma  anlam  kattığınız  için  teşekkür  ederim.  Benim  için  çok  kıymetli  insanlarsınız.  Daha  pek  çok  partilerde,  kamplarda,  gezilerde,  toplantılarda,  gösteri-­‐protesto  ve  forumlarda  buluşmak  dileğiyle.    

Sevgili  Sema,  hayatımın  ciddi  bir  şekilde  değişmeye  başladığı,  pek  çok  önemli  adımın  atıldığı  bir   dönemde   tam   da   ihtiyacım   olan   arkadaşlığı   sunduğun   ve   en   unutulmaz   anlarda   tüm  mesafelere  rağmen  yanımda  olmayı  tercih  ettiğin  için  teşekkür  ederim.    

Hollanda   Gezi   Dayanışması   üyeleri,   Osman   Hocam,   Kıvılcım,   Hande,   Maral,   Leyla,   Yaşar,  Burhan,  Çiğdem,  Hakan,  Nil,  Levent,  Elif,  Eylem,  Derya  ve  adını  yazmadığım  herkese  sonsuz  teşekkürler.  Umut  dolu,  samimi,  sıcacık  insanlarsınız  ve  iyi  ki  varsınız.      

Annem,  babam  ve  Ceren...  Sizin  gibi  bir  ailem  olduğu  için  çok  şanslıyım.  Sevginizi,  desteğinizi  ve  güveninizi  benden  hiçbir  zaman  esirgemediğiniz  için  size  ne  kadar  teşekkür  etsem  azdır.    

ve   Gizem...     En   iyi   arkadaşım,   diğer   yarım...   Bu   tezin   içindeki   her   deney   hasretinle,   sana  kavuşmanın  hayaliyle  yapıldı.  Çok  bekledik,  zorlandık,  kızdık,  küstük  ama  demek  ki  en  çok  da  sevdik.  Yoksa  aşılır  mıydı  onca  sınır?  Ceylan’ın  da  dediği  gibi  beklediğime  değdi  J.    

 

Abidin

Acknowledgements

121

Page 129: Mechanochemical scission of transition metal-ligand bonds ... · 11/11/2015  · useful chemical transformations. Current developments in mechanochemistry involve the introduction