ORCA labs 6 Frontier orbitals - University of...

20
PART II – ADVANCED AND SPECIAL SUBJECTS 1 PART II ADVANCED AND SPECIAL SUBJECTS

Transcript of ORCA labs 6 Frontier orbitals - University of...

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS   1  

PART  II      

ADVANCED AND SPECIAL SUBJECTS

 

 

 

 

 

 

 

 

 

 

 

 

     

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     2  

1 Computer   Experiment   7:   Interpretation   of   Structure,   Bonding  and  Reactivity  Using  Orbitals  

1.1  Background  

1.1.1 Hückel  MO  Theory  The  Hückel  theory  is  the  simplest  semi-­‐empirical  method  for  the  description  of  

molecular  systems.  It  was  developed  in  1931  by  Erich  Hückel  for  the  calculation  of  

planar  conjugated  hydrocarbons.  Later  Hoffmann  generalized  this  very  successful  

concept  to  general  molecules  and  applied  it  with  great  success  to  many  areas  of  

chemistry.  He  termed  his  method  the  “extended  Hückel  theory”  (EHT).  The  appeal  of  

the  Hückel  and  EHT  methods  is  that  they  can  provide  many  qualitative  insights  into  

the  behaviour  of  molecules,  in  particular  of  molecules  that  are  closely  related.1  It  

does  not,  however,  provide  reliable  numbers.  For  this  purpose  ab  initio  or  DFT  

methods  must  be  consulted.2    

 

A.  Extended  Hückel  MO  theory  

EHT  theory  might  be  thought  of  as  an  essentially  very  simple,  semiempirical  

approximation  of  Hartree-­‐Fock  theory.  In  Hartree-­‐Fock  theory,  a  single  Slater  

determinant  is  taken  as  an  Ansatz  for  the  description  of  the  N-­‐electron  system.  The  

single  determinant  is  composed  of  one-­‐particle  molecular  orbitals  (MOs)  that  are  

found  as  solutions  to  a  pseudo-­‐eigenvalue  problem  of  the  form:  

  F!i= "

i!

i                  

  (  1)  

Where   F  is  the  „Fock  operator“,  which  describes  the  motion  of  a  single  electron  in  

the  field  of  the  nuclei  and  the  remaining  electrons.The  MOs  are  written  as  a  linear  

combination  of  atomic  orbitals  (LCAO-­‐Ansatz):  

                                                                                                               1  Recommended   literature  to  this  chapter   is:   Ian  Fleming,  Grenzorbitale  und  Reaktionen  organischer  Verbindungen,  VCH  1990  2  Even  today  it  might  not  be  a  bad  idea  to  do  a  few  simple  EHT  calculations  at  the  beginning  of  a  project  if  one  has  no  familiarity  with  the  systems  being  investigated  and  to  use  the  results  to  gain  some  feeling  for  the  factors  that  might  be  worthwhile  to  examine  with  more  rigorous  electronic  structure  methods.  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     3  

  !

ir( ) = c

µi"

µr( )

µ!                

  (  2)  After  which  the  calculation  boils  down  to  the  solution  of  a  generalized  eigenvalue  

problem  for  the  determination  of  the  unknown  MO  coefficients   c

µi  and  the  orbital  

energies   !i  

Fc = !Sc                  

  (  3)  

Where   !  is  a  diagonal  matrix  with  orbital  energies  and  the  matrix  elements  of  the  

Fock  matrix  F  and  the  overlap  matrix  S  are  defined  by:  

  F

µ!= "

µF "

!                

  (  4)    

S

µ!= "

µ"!                

  (  5)  As  pointed  out  in  the  introduction,  Hartree-­‐Fock  calculations  require  the  use  of  large  

basis  sets  and  iterative  cycles  for  the  optimization  of  the  MOs.  In  EHT  theory,  one  

focuses  on  the  qualitative  shape  of  the  valence  orbitals  and  usually  pays  attention  to  

those  orbitals  that  are  located  near  the  HOMO-­‐LUMO  gap  (the  “frontier  orbitals”).  

According  to  frontier  molecular  orbital  theory  (described  below  in  chapter  1.1.3)  

these  are  the  most  important  ones  for  the  reactivity  of  the  system.  In  order  to  get  an  

impression  of  how  these  orbitals  may  look  like  it  is  not  necessary  to  solve  the  rather  

laborious  HF  equations  –  in  fact  this  was  a  major  challenge  for  all  but  the  smallest  

molecules  in  the  1960s  when  EHT  theory  was  suggested  –  it  is  enough  to  replace  the  

Fock  operator  by  a  semiempirical  effective  one-­‐electron  operator   heff  and  to  only  

consider  the  “minimal”  chemical  set  of  valence  orbitals3  as   !{ } .  It  is  not  necessary  

to  specify  the  operator   heff  precisely.  In  order  to  solve  eq  (3),  it  is  only  necessary  to  

specify  the  matrix  elements  of  this  operator.  In  EHT  theory  they  are  given  by:  

                                                                                                               3  For  example,  for  a  carbon  atom  one  includes  2s,  2px,  2py  and  2pz  and  for  titanium  one  would  use  3dxy,  3dx2-­‐y2,  3dxz,  3dyz,  2dz2-­‐r2,  4s  and  perhaps  also  4px,  4poy  and  4pz.  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     4  

 

hµ!eff = "

µheff "

!=

if µ = !

$µ!

if µ ! !

"#$$$

%$$$

  (  6)  

The  diagonal  elements  ! µ  represent  the  “energy  of  the  atomic  orbital  !µ  in  the  free  

atom”.  It  is  approximated  by  the  so-­‐called  “valence  shell  ionization  energy”  (VOIP)  

which  can  be  determined  from  atomic  spectroscopy.4  These  numbers  are  negative  

and  represent  the  average  energy  required  to  remove  an  electron  from  the  given  

orbital  (2s,  2p,  3d,  …)  The  lower  these  energies,  the  higher  the  electron  attracting  

ability  of  the  atom.  The  off-­‐diagonal  matrix  elements  are  called  “resonance  

integrals”.  They  measure  the  strength  of  the  interaction  between  two  atomic  

orbitals.  Intuitively,  it  is  reasonable  to  expect  that  these  integrals  are  related  to  the  

overlap  integrals.  Thus,  in  EHT  they  are  given  by:5  

  !µ" =

12 KSµ" # µ +#"( )              

  (  7)  

The  constant  K  ~  1.75  is  an  empirical  constant  to  adjust  the  resonance  integrals  to  

more  reasonable  values  given  that  the  form  of  eq  (7)  is  an  oversimplification.  In  

order  to  provide  a  feel  for  the  values  of  the  parameters  we  give  them  for  elements  

H-­‐Ne  in  Table  1:.  Table  1:  VOIPs  for  elements  H-­‐Ne  to  be  used  together  with  EHT  calculations.  Note,  that  the  negative  of  the  values  listed  below  is  to  be  used.  

Element   αs  (eV)   αp  (eV)  H   13.61   -­‐  He   24.59   -­‐  Li   5.39   3.60  Be   9.32   6.00  B   12.93   8.30  C   16.59   11.26  N   20.33   14.53  

                                                                                                               4  The   values   of   these   parameters   are   tabulated   in  multiple   places.   Since   there   are  many  ways   to   determine   these  parameters  from  the  experimental  data,  no  consensus  has  been  reached  in  the  community  about  a  “universal”  set  of  VOIPs.    5  There  are  many  modifications  of  the  formulas  for  the  off-­‐diagonal.   It   is  questionable   if  any  of  these   is  really  to  be  preferred  over  another  one  and  in  this  situation  “Ockham’s  razor”  principle  applies  which  essentially  states  that  the  simplest  solution  is  just  as  good  as  any.  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     5  

O   28.48   13.62  F   37.85   17.42  Ne   48.47   21.61    

Following  the  solution  of  the  generalized  eigenvalue  problem  in  eq  (3)  and  with  the  

approximation  described  above,  the  resulting  MOs  are  filled  in  order  of  increasing  

energy  with  the  available  electrons  in  keeping  with  the  “Aufbau  principle”  in  order  

to  find  the  electronic  ground  state.  Since  the  approximations  are  so  simple,  the  total  

energy  in  the  EHT  method  is  simply  the  sum  of  orbital  energies:  

  E EHT = ni! i

i"                   (  8)  

Where   ni  is  the  occupation  number  of  the  i’th  orbital  (=0,1  or  2).  In  connection  with  

Walsh’s  rules,  we  will  see  below  that  the  variations  of  the  orbital  energies  with  

geometry  can  be  used  to  obtain  quite  important  and  general  insights  about  the  

geometric  structures  of  molecules.  In  fact,  many  areas  of  chemistry  have  profited  

from  performing  such  qualitative  calculations.    

B.  Hückel  theory  for  π  electron  systems  (HMO)  

Although  it  existed  prior  to  EHT  theory,  the  HMO  theory  of  aromatic  π-­‐systems  

might  be  thought  of  as  a  further  simplification  of  the  EHT  method.  In  this  case,  one  

only  considers  the  π-­‐electrons  of  the  investigated  molecules  and  only  keeps  the  pz  

orbitals  of  the  atoms  involved  in  the  π-­‐system.  Substituent  effects  are  included  in  an  

approximate  manner  by  modifying  the  α-­‐parameters  of  the  atoms  to  which  the  

substituents  are  attached.  Furthermore,  in  HMO  theory,  the  off-­‐diagonal  elements  of  

the  overlap  matrix  are  neglected  such  that  the  final  eigenvalue  problem  to  be  solved  

is:  

  heff c = !c                  

  (  9)  

Finally,  one  only  keeps  resonance  integrals  between  atoms  that  are  nearest  

neighbours  and  assigns  a  constant  value  to  them.  Thus,  the  calculations  also  become  

geometry  independent  and  the  only  thing  that  is  required  in  order  to  perform  a  

HMO  calculation  is  a  molecular  connectivity.  Although  these  approximations  clearly  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     6  

represent  a  gross  oversimplification  of  the  molecular  electronic  structure  problem,  

it  can  hardly  be  overemphasized  how  much  HMO  theory  has  shaped  chemical  

thinking  about  aromatic  molecules  and  their  properties.6  However,  we  will  not  

pursue  HMO  theory  in  this  course  but  stay  in  the  framework  of  the  EHT  method  

below.  

1.1.2 Walsh’s  Rules7  Walsh  diagrams  are  the  graphical  representation  of  the  MO  energies  in  dependence  

of  a  geometrical  parameter,  most  commonly  a  bond  angle.  For  example,  consider  the  

case  of  the  H3  molecule  in  its  linear  and  triangular  forms  as  shown  in  Figure  1.  It  is  

observed  that  upon  bending,  the  energy  of  the  nonbonding  σu  orbital  (it  correlates  

with  a  b2  orbital  in  C2v  symmetry)  strongly  increases  and  finally  correlates  with  one  

component  of  the  antibonding  e-­‐orbital  in  the  final  D3h  structure.  Since  there  are  

three  electrons  to  be  filled  into  the  three  MOs,  it  is  predicted  that  H3  should  be  linear    

((1σg)2(1σu)1  configuration)  while  H3+  should  be  triangular  ((a1)2  configuration)  

since  the  a1  orbital  is  stabilized  upon  bending.    

 Figure  1:  Walsh  diagram  for  the  distortion  of  the  H3  system.  

The  variation  of  a  geometrical  degree  of  freedom  may  cause  a  crossing  of  different  

MO-­‐levels.  In  general,  such  a  crossing  is  allowed  if  the  two  MOs  transform  under  

different  irreducible  representations  and  is  avoided  otherwise.(the  famous  non-­‐

crossing  rule;  compare  Figure  2).  The  non-­‐crossing  rule  is  particularly  important  for  

the  interpretation  of  photochemical  reactions.  The  occurrence  of  a  HOMO  /  LUMO  

                                                                                                               6  The  classic  text  on  HMO  theory  is  E.  Heilbronner,  H.  Bock  “Das  HMO  Modell  und  seine  Anwendung”.  Verlag  Chemie,  Weinheim/Bergstrasse,  1968  7  For  a  detailed  ab  initio  perspective  on  Walsh’s  rules  see:  Buenker,  R.J.M;  Peyerimhoff,  S.D.  Chem.  Rev.,  1974,  74,  127  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     7  

crossing  with  identical  symmetry  implies  that  the  reaction  is  symmetry  forbidden,  

due  to  the  non-­‐conservation  of  the  orbital  symmetry  of  the  occupied  MOs.  Therefore  

a  high  activation  energy  is  expected.  

 

 Figure  2:  Crossing  and  avoided  crossing  of  two  energy  levels.  

In  Figure  3  the  Walsh  diagram  for  the  bending  mode  of  a  AH2  molecule  is  shown.  

The  energy  of  the  lowest  valence  MOs  decrease  upon  bending  as  three  center  

bonding  between  s-­‐AOs  becomes  more  favourable  in  the  bent  form.  The  1  σu-­‐MO  is  

strongly  bonding  in  the  linear  case,  but  upon  reducing  the  angle,  the  1b2-­‐MO  

becomes  only  weakly  bonding,  because  in  the  linear  case  the  2py-­‐AOs  can  interact  

more  strongly  with  the  AOs  of  the  H-­‐atoms.  The  1  πu-­‐MO  separates  in  two  

components,  1b1  and  3a1.  The  energy  of  the  1b1  MO  is  almost  constant  while  the  3a1-­‐

MO  becomes  a  bonding  MO  upon  bending.  The  overlap  of  the  2pz-­‐AOs  from  A  with  

1s-­‐AOs  from  the  H-­‐atoms  is  zero  in  the  linear  case,  but  reducing  the  angle  results  in  

a  three-­‐center  bonding.  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     8  

Figure  3:  Walsh  diagram  for  a  AH2  molecule.8  

 

1.1.3 Approximate  Intermolecular  Interactions  In  this  section  we  will  indicate  how  the  energies  and  shapes  of  molecular  orbitals  

can  be  related  to  chemical  reactivity.  One  of  the  key  features  of  chemical  reactions  is  

the  activation  energy  that  has  to  be  overcome  by  the  reactants.  On  the  basis  of  

perturbation  theory  and  EHT,  Klopman  and  Salem  [G.  Klopman,  JACS,  90,  223  

(1968);  S.  Salem,  JACS,  90,  543  (1968)]  developed  a  simple    but  highly  useful  

decomposition  scheme  for  the  interaction  energy  ΔE  of  two  approaching  molecules  

A  and  B  into  contributions  from  occupied  and  unoccupied  orbitals.  All  quantities  in  

(10)  are  obtained  for  the  separate,  non-­‐interacting  molecules.  

!E =" (qµ

+q!)"

µ!µ!

A,B

# Sµ!

+Q

IQ

J

#RIJI<J

#

+2( c

aµc

q!"

µ!)2

µ!#

Ea"E

qq

unocc

#a

occ

# +2( c

bµc

p!"

µ!)2

µ!#

Eb"E

pp

unocc

#b

occ

#

 

  (  10)  This  expression  –  although  somewhat  lengthy  –  is  already  simplified  (the  approach  

does  not  take  electron-­‐electron  interaction  into  account).  So  let  us  discuss  the  terms  

step  by  step.  The  first  term  is  a  double  sum  over  all  atomic  orbitals  (AOs)  μ  of  

                                                                                                               8Kutzelnigg,  W.,  Einführung  in  die  Theoretische  Chemie  Band  2,  VCH:  Weinheim,  1994  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     9  

molecule  A  and  ν  of  molecule  B.  The  q’s  are  the  AO  occupation  numbers  of  the  non-­‐

interacting  molecules  (in  the  original  paper  called  charge  densities),  β  is  similar  but  

not  identical  to  the  resonance  integral  of  Hückel  theory  (it  includes  only  the  

electron-­‐nuclear  attraction  terms),  and  S  is  the  overlap  integral  mentioned  

previously.  Since  β  is  always  negative,  this  term  is  positive  and  is  basically  

responsible  for  the  occurrence  of  activation  barriers  within  this  model.  Its  origin  is  

the  repulsive  interaction  between  doubly  occupied  molecular  orbitals  (MOs)  of  A  

and  B  which  arises  from  the  non-­‐symmetric  level  splitting  (Figure  4).  

 Figure  4:  Interaction  of  Two  Occupied  Molecular  Orbitals  

The  second  term,  a  sum  over  all  atoms  I  of  molecule  A  and  atoms  J  of  molecule  B,  

describes  the  Coulomb  interaction  between  pairs  of  atoms  with  net  charge  Q  at  

distance  R  (ε  is  the  dielectric  constant).  The  two  last  terms  include  sums  over  all  

occupied  MOs  a  of  A  and  unoccupied  MOs  q  of  B  and  vice  versa.  For  each  pair  of  

orbitals  the  term  consists  of  sums  of  corresponding  MO  coefficients  c  weighted  with  

β.  The  dominator  is  the  MO  energy  difference.  This  part  describes  the  stabilization  

due  to  formation  of  orbitals  combining  occupied  A  MOs  with  unoccupied  B  MOs  or  

vice  versa  in  the  complex  [AB]*.  

 Figure  5:  Interaction  of  an  Occupied  and  an  Unoccupied  Orbital.  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     10  

 

1.1.4 Frontier  orbital  picture  It  is  exactly  this  last  effect  which  is  widely  used  in  further  simplified  qualitative  

approaches  to  determine  chemical  reactivity.  If  the  two  reacting  molecules  are  

similar,  the  smallest  energy  differences  are  those  between  the  highest  occupied  MOs  

(HOMOs)  and  the  lowest  unoccupied  MOs  (LUMOs).  If  it  is  assumed  that  the  other  

contributions  to  the  last  two  terms  of  (1)  are  always  of  similar  magnitude,  the  

HOMO-­‐LUMO  contributions  will  dominate.  If  it  is  further  assumed  that  the  first  term  

is  roughly  independent  from  the  intermolecular  orientation,  only  two  main  

contributions  are  left  for  the  determination  of  the  feasibility  and,  sometimes  even  

more  important,  the  regioselectivity  of  chemical  reactions:  the  Coulomb  term  and  

the  MO  coefficients  of  the  frontier  orbitals,  HOMO  and  LUMO.    

This  concept  is  used  e.g.  to  estimate  the  relative  reactivity  of  different  compounds  C  

towards  a  given  reactant  R  via  calculation  of  E(HOMO,C)-­‐E(LUMO,R)+E(HOMO,R)-­‐

E(LUMO,C).  

The  frontier  orbital  approach  is  also  applied  to  ionic  reactions  in  organic  chemistry.  

In  electrophilic  or  nucleophilic  reactions  the  inorganic  ions  are  classified  as  hard  or  

soft.  A  hard  nucleophilic  has  a  low  HOMO  energy,  a  hard  electrophilic  has  a  high  

LUMO  energy,  and  vice  versa.  In  nucleophilic  reactions,  a  hard  anion  (e.g.  OH-­‐,  NH2-­‐)  

will  attack  preferentially  the  atom  with  the  highest  positive  charge  Q  of  an  organic  

molecule.  In  the  frontier  orbital  picture  this  is  due  to  the  large  energetic  difference  

between  the  anion’s  HOMO  and  the  molecule’s  LUMO  making  the  last  term  in  the  

above  equation  small  so  that  the  Coulomb  part  prevails.  On  the  other  hand,  the  

preferred  site  for  a  soft  anion  (CN-­‐,  (CH2,Ph)-­‐C-­‐O-­‐)  is  the  atom  with  the  largest  

contribution  to  the  molecule’s  LUMO.  

1.1.5 The  Woodward-­‐Hoffman  Rules  In  reaction  theory  synchronous  organic  reactions  such  as  pericyclic  reactions  would  

in  general  take  place  under  preservation  of  the  orbital  symmetry.  Each  occupied  MO  

of  the  reactants  will  result  in  an  occupied  MO  of  the  products  with  the  same  

symmetry.  Experimentally,  there  are  two  ways  of  activating  such  reactions:  

1. Thermal  activation  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     11  

2. Photochemical  activation  

For  the  theoretical  description,  the  main  difference  between  these  two  methods  lies  

in  the  MOs  that  are  involved  in  these  processes.  The  photochemical  activation  will  

result  in  the  excitation  of  an  electron  from  a  bonding  (or  nonbonding)  into  an  anti-­‐

bonding  orbital.  The  resulting  state  may  have  a  different  symmetry  than  the  ground  

state  and  will  therefore  also  have  a  distinct  reactivity  (Figure  6).  Essential  difference  

between  photochemical  and  thermal  reactions  must  therefore  be  expected  and  are  

also  observed  in  practice.    

 Figure  6:  The  orbitals  that  are  involved  in  the  different  activation  cases.9  

A.  Electro-­‐cyclic  reactions  

Electro-­‐cyclic  reactions  may  be  described  as  an  isomerisation  of  open-­‐chain  polyens  

to  ring  isomers  which  takes  place  under  thermal  or  photochemical  activation.  In  

order  to  close  the  ring  an  additional  σ-­‐bond  has  to  created  at  the  expense  of  a  π-­‐

bond.  In  EHT  language,  the  π-­‐MOs  on  one  fragment  have  to  overlap  with  correct  

phases  with  the  orbitals  of  another  fragment  in  order  to  form  a  σ-­‐bond.  The  

example  of  1,3-­‐Butadiene  is  shown  above  in  (Figure  6).  In  fact,  it  can  be  determined  

in  advance,  when  a  suitable  constructive  overlap  of  the  π-­‐orbitals  is  possible  –  it                                                                                                                  9Woodward,  R.  B.;  Hoffmann  R.  Angew.  Chem.  81.  Jahrgang  1969,  21,  797  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     12  

simply  depends  on  the  number  of  π-­‐electrons.  As  long  as  both  fragments  stay  in  

their  electronic  ground  state  4n  (n=1,2,...)  π-­‐electrons  in  the  molecule  lead  ti  in-­‐

phase  arrangement  of  the  terminal  π-­‐orbitals,  whereas  4n+2  (n=1,2,...)  π-­‐electrons  

result  in  an  out-­‐of-­‐phase  arrangement.  In  order  to  build  up  a  new  σ-­‐bond,  the    π-­‐

orbitals  in  the  molecule  with  4n  π-­‐electrons  have  to  rotate  in  the  same  direction  

(conrotatory)  in  order  to  result  in  a  bonding  orbital,  while  the  molecules  containing  

4n+2  π-­‐electrons  are  obliged  to  rotate  in  opposite  directions  (disrotatory)  (compare  

Figure  7)  For  photochemical  reactions  the  requirements  for  constructive  overlap  

are  exactly  the  reverse  of  the  one  described  above.  Therefore,  a  system  with  4n  π-­‐

electrons  has  to  rotate  in  opposite  directions,  whereas  a  system  containing  4n+2  π-­‐

electrons  needs  to  rotate  in  the  same  direction.  Since  the  two  possibilities  lead  to  

different  stereoisomers  of  the  products  that  are  formed  (Figure  7),  one  can  readily  

predict  whether  one  has  to  activate  a  given  reaction  photochemically  or  thermally  in  

order  to  obtain  the  desired  product.  

 

 Figure  7:  Example  for  con-­‐  and  dis-­‐rotatory  rotation  of  the  terminal  pz  orbitals    in  case  of  thermal  activation  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     13  

Table  2:  Summary  of  allowed  and  forbidden  electrocyclic  reactions  based  on  the  Woodward-­‐Hoffman  rules.  

m+n-electro-cyclic reactions Thermal activation

ground state

Photochemical activation

excited state

forbidden allowed allowed forbidden

M+n = 4q q=1, 2,... disrotatory   conrotatory   disrotatory   conrotatory  M+n = 4q + 2 q=1, 2,... conrotatory   disrotatory   conrotatory   disrotatory  

 

B.  Cycloadditions  

In  concerted  cycloadditions  two  new  σ-­‐bonds  are  built  up  due  to  overlap  of  the  π-­‐

orbitals  of  both  reactants.  The  location  of  the  σ-­‐bond  is  arranged  either  on  the  same  

side  of  the  reacting  π-­‐system  or  on  the  opposite  side.  The  first  case  is  referred  to  as  

a  “suprafacial”  reaction  and  the  second  case  as  a  “antarafacial”  process.  The  

alignment  of  the  “responsive”  π-­‐orbitals  is  again  dependent  on  the  number  of  π-­‐

electrons  and  on  the  form  of  activation  as  summarized  in  Figure  8,  Figure  9  and  

Table  3.  

 Figure  8:  Orbital  symmetry  of  a  photochemical  cyclic  addition.  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     14  

 Figure  9:  Orbital  symmetry  of  a  thermal  cyclic  addition.  (a)  anti-­‐binding  (b)  binding,  but  geometrically  difficult  

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     15  

Table  3:  Summary  of  the  Woodward-­‐Hoffman  rules.  

m+n-cyclic addition Thermal activation

ground state

Photochemical activation

excited state

forbidden allowed allowed forbidden

m+n = 4q q=1, 2,... Πma  +  Πna    antara,  antara  

Πms  +  Πna    supra,  antara  

Πms  +  Πns    supra,  supra  

Πma  +  Πns    antara,  supra  

m+n = 4q + 2 q=1, 2,... Πma  +  Πns    antara,  supra  

Πms  +  Πns    supra,  supra  

Πms  +  Πna    supra,  antara  

Πma  +  Πna    antara,  antara  

 

1.1.6 Limitations  of  the  Frontier  Molecular  Orbital  Approach  Although  the  above  scheme  has  been  proven  to  give  qualitatively  correct  

predictions  in  many  cases,  as  we  will  see  below,  it  is  of  course  limited  due  to  its  

approximate  nature.  From  the  quantum-­‐chemical  viewpoint,  it  is  mainly  the  use  of  

atomic  charges  and  LUMO  energy  and  shape  that  makes  the  simple  approach  

questionable.  Atomic  charges  are  no  observables  and  strongly  depend  on  the  

analysis  and  the  basis  set.  But  also  the  shape  of  the  orbitals  is  method  and  basis  set  

dependent.  The  MOs  employed  in  the  above  analysis  are  obtained  with  minimal  

basis  sets  in  the  extended  Hückel  framework.  Nowadays,  one  of  course  wants  to  go  

beyond  this  simple  level  of  theory  and  use  HF  or  DFT  calculations  as  basis  for  

analysis  of  chemical  reactivity.  And  here  arises  a  fundamental  problem:  For  a  given  

method,  the  HOMO  converges  to  something  definite  while  the  unoccupied  orbitals  

do  not.  Remember  that  only  the  occupied  orbitals,  either  included  in  the  Slater  

determinant  in  HF  theory  or  used  to  calculate  the  electron  density  in  DFT,  are  

optimized  by  the  variation  procedure.  The  only  criterion  for  unoccupied  orbitals  is  

their  orthogonality  to  the  occupied  space.  In  the  limit  of  a  complete  basis  set,  the  

unoccupied  space  forms  a  continuum  and  the  LUMO  is  not  really  defined.  We  will  

demonstrate  this  in  an  example  in  this  experiment.  In  HF  theory  even  the  energies  

Ep,  Eq  of  the  unoccupied  MOs  are  fundamentally  wrong  because  the  fictitious  

(virtual)  electrons  in  the  UMOs  experience  the  repulsive  field  of  all  N  electrons  of  

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     16  

the  system,  whereas  the  occupied  orbitals  experience  the  correct  N-­‐1  electron  

potential.  Their  energy  can  be  related  to  physical  properties  (ionization  energies)  

via  the  Koopmans  theorem.  The  situation  in  DFT  is  not  much  better.  Here  the  virtual  

MOs  experience  the  correct  N-­‐1  electron  repulsive  field,  but  the  energies  Ea,  Eb  of  the  

occupied  MOs  are  incorrect  due  to  artificial  self-­‐interaction  of  the  electrons  which  is  

caused  by  the  approximate  description  of  electron  exchange  in  DFT.  As  a  

workaround  of  the  above  mentioned  shortcomings  of  HF  and  DFT,  hybrid  methods  

such  as  B3LYP  have  been  developed  where  the  electron  exchange  is  described  by  a  

mixture  of  the  exact  one-­‐determinant  expression  provided  by  HF  theory  and  a  

density  functional.  The  mixing  coefficient  is  usually  treated  as  an  empirical  

parameter.  Its  value  is  determined  by  optimal  reproduction  of  experimental  

properties  (including  ionization  energies  and  optical  spectra).  In  this  way  one  uses  

the  cancellation  of  the  inherent  errors  of  both  approaches.  

Having  this  in  mind,  it  is  not  surprising  that  the  frontier  orbital  approach  fails  in  

some  cases.  Nevertheless,  as  a  qualitative  tool  to  understand  fundamental  reasons  

for  the  course  of  chemical  reactions,  it  is  still  worth  to  discuss.  

1.2  Description  of  the  Experiment  

1.2.1 Frontier  Molecular  Orbital  Theory  As  an  example  we  study  the  reactivity  of  the  pyridinium  cation  shown  below:    

 Figure  10:  The  pyridinium  cation  studied  in  this  experiment.  

Use  the  Gaussian  and/or  ORCA  programs  and  one  of  the  methods  RHF,  B3LYP,  or  

BP86  to  calculate  the  most  reactive  sites  of  the  pyridinium  cation  in  a  reaction  with  

a  hard  and  with  a  soft  nucleophilic,  respectively,  based  on  an  analysis  of  the  

Mulliken  net  charges  and  the  LUMO  coefficients.      

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     17  

Create  the  molecular  structure  as  described  in  previous  experiments,  perform  a  

structure  optimization  and  analyse  the  wavefunction  of  the  final  structure.  

Check  if  the  Mulliken  net  charges  and  the  shape  of  the  LUMO  change  if  you  increase  

the  basis  set  from  SVP  to  TZVP.  Will  the  qualitative  results  of  the  frontier  orbital  

analysis  change?  Hint:  instead  of  analysing  the  MO  coefficients  it  is  sufficient  to  

draw  an  orbital  picture  using  MOLDEN  or  MOLEKEL.  In  any  case  the  keyword  

POP=FULL  is  needed  in  Gaussian  (ORCA  prints  a  Loewdin  analysis  of  each  MO  by  

default).  

Experimentally  it  is  known  that  hard  nucleophilics  (OH-­‐,  NH2-­‐)  prefer  C(2),  and  soft  

nucleophilics  (CN-­‐,  (CH2,Ph)-­‐C-­‐O-­‐)  prefer  C(4).  

Verify  the  relative  reactivity  of  the  two  sites  versus  OH-­‐  and  CN-­‐  by  calculating  the  

energy  difference  of  the  four  transition  structures  CN(2),  CN(4),  OH(2)  and  OH(4)  as  

shown  below.  In  this  case  it  is  sufficient  to  fully  optimize  the  structures,  no  TS  

search  is  necessary.    

     

       

Figure  11:  Structures  of  the  Reaction  Products  to  be  Optimized  

 

1.2.2 The  Woodward-­‐Hoffman  Rules  

- Dissociation  of  formaldehyde  

Use  the  ORCA  program  to  optimize  the  following  structures  employing  HF/6-­‐311G*        

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     18  

and  generate  .cube  files  for  each  MO  using  orca_plot.  Alternatively,  you  can  use  the  

Gaussian  program  employing  HF/6-­‐311G*  and  generate  .chk  files.    

A.    H2C=O  

B.  CO  

C.  H2  Classify  the  HF  valence  orbitals  and  the  lowest  unoccupied  orbitals  according  to  

their  irreducible  representation  (irrep).  Investigate  the  orbital  symmetry  

concerning  the  symmetry  elements  of  the  point  group  C2v.  

•  Identity  (E)  

•  two-­‐fold  rotation  axis  (C2)  

•  mirror-­‐plane  in  the  xz-­‐plane  (σxz)  

•  mirror-­‐plane  in  the  yz-­‐plane  (σyz)  

 

 Figure  12:  Coordinate  system  used  in  this  experiment.  

Table  4:  The  character  table  of  the  C2v  point  group.  

C2v E C2 σxz σyz

A1 1 1 1   1 z x2, y2, z2

A2 1 1 -1 -1 Rz xy

B1 1 -1 1 -1 x, Ry xz

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     19  

B2 1 -1 -1 1 y, Rx

 

• Draw  a  MO  diagram  for  H2C=O  on  the  one  side  and  for  CO  +  H2  on  the  other  side.  Connect  orbitals  which  belong  to  the  same  irrep  with  each  other.  

• Repeat   these   steps   in   Cs   symmetry   with   the   symmetry   elements   E   and   σ.  Remember  that  there  are  two  possibilities  for  the  orientation  of  formaldehyde  in  this  point  group.  What  should  attract  your  attention?  

1.2.3 Walsh’s  Rules  Perform  calculations  on  the  following  molecules  with  the  Hartree-­‐Fock  method  and  the  SVP  basis  set  and  the  ORCA  program:  

• LiH2+  (distance=  1.7  Angström)  

• BeH2  (distance=  1.34  Angström)  

• CH2  (S=0;  RKS  calculation)  (distance=1.078  Angström)  

• CH2  (S=1;  ROHF  calculation;  keyword    !  ROHF)  (distance  1.078  Angström)  

• H2O  (distance=0.956  Angström)  

First  look  at  the  variation  of  the  total  energy  as  a  function  of  angle  by  performing  a  

rigid  scan  of  that  angle  and  plot  the  results  with  xmgrace.  An  input  is:  

                   A  summary  of  the  total  energy  as  a  function  of  angle  will  be  printed  at  the  bottom  of  

the  output  file.    

Now  look  at  the  orbitals  and  their  energies  at  the  bent  geometry  and  the  linear  

geometry.  Plot  them.  Discuss  the  variation  of  the  orbital  energies  and  the  total  

energy  as  a  function  of  bond  angle.  Do  you  find  Walsh’s  results  to  be  consistent  with  

the  ab  initio  results?    

ADDITIONAL  CALCULATIONS  (VOLUNTARY)  

# Scan the angle ! RHF SVP %paras Ang= 180,90,17 end * int 0 1 O 0 0 0 0 0 0 H 1 0 0 0.956 0 0 H 1 2 0 0.956 {Ang} 0 *

PART  II  –  ADVANCED  AND  SPECIAL  SUBJECTS     20  

Plot  the  variation  of  the  nuclear  repulsion  energy,  the  one-­‐electron  energy,  the  two-­‐

electron  energy  as  well  as  the  kinetic  energy  of  the  system  as  a  function  of  angle.  

Also  plot  the  sum  of  the  occupied  MO  energies  weighted  by  their  occupancy  as  a  

function  of  bond  angle.  These  data  must  be  extracted  from  the  ORCA  output  file.  

What  do  you  find?  What  is  the  physical  reason  for  obtaining  a  bent  geometry  versus  

a  linear  geometry?.