37
1 Computer Experiment 15: Computational Coordination Chemistry 1.1 Background 1.1.1 Introduction The rich chemistry of transition metals gives rise to the large and important fields of coordination chemistry, organometallic chemistry and branches of solid state chemistry. In fact, transition metal complexes play an important role in biology and chemical catalysis since they are able to catalyze a very large variety of reactions, many of them of fundamental relevance to biological and industrial processes. The high reactivity of transition metal complexes is intimately linked to their electronic structure and the nature of the metalligand bonds. In fact, these bonds are typically stronger than the weak interactions studied in Computer Experiment number 13 (section Error! Reference source not found., page Error! Bookmark not defined.) but are weaker than the typical covalent bonds formed in organic chemistry. Consequently, the metalligand bonds are easier to break and reform which gives rise to lowenergy pathways in chemical reactions. Furthermore, due to the partially filled dshells most of the transition metals are redox active and can exist in different oxidation states (e.g. Mn can assume oxidation states from +VII to 0 or even I). The potential at which the oneelectron steps occur are a sensitive function of the nature of the ligands and the coordination geometry and therefore be easily tuned by suitable ligand design. Furthermore, transition metals typically contain unpaired electrons in their partially filled dshells which gives rise to several accessible spin states. These spin states may show considerably different reactivities and this leads to added complexity in studying their reactions. Frequently, the systems even pass from one spin state to another surface during a given catalytic cycle which opens many fascinating routes to unique reactions. Last but not least, the partially filled dshells and the variable metalligand bonds are associated with a huge variety of fascinating spectroscopic phenomena. Thus, very detailed insight into the geometric and electronic structure of the transition metal

1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

Embed Size (px)

Citation preview

Page 1: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

1 Computer   Experiment   15:   Computational   Coordination  Chemistry  

1.1  Background  

1.1.1 Introduction  The  rich  chemistry  of  transition  metals  gives  rise  to  the  large  and  important  fields  of  

coordination  chemistry,  organometallic  chemistry  and  branches  of  solid  state  

chemistry.  In  fact,  transition  metal  complexes  play  an  important  role  in  biology  and  

chemical  catalysis  since  they  are  able  to  catalyze  a  very  large  variety  of  reactions,  

many  of  them  of  fundamental  relevance  to  biological  and  industrial  processes.  The  

high  reactivity  of  transition  metal  complexes  is  intimately  linked  to  their  electronic  

structure  and  the  nature  of  the  metal-­‐ligand  bonds.  In  fact,  these  bonds  are  typically  

stronger  than  the  weak  interactions  studied  in  Computer  Experiment  number  13  

(section  Error!  Reference  source  not  found.,  page  Error!  Bookmark  not  

defined.)  but  are  weaker  than  the  typical  covalent  bonds  formed  in  organic  

chemistry.  Consequently,  the  metal-­‐ligand  bonds  are  easier  to  break  and  reform  

which  gives  rise  to  low-­‐energy  pathways  in  chemical  reactions.  Furthermore,  due  to  

the  partially  filled  d-­‐shells  most  of  the  transition  metals  are  redox  active  and  can  

exist  in  different  oxidation  states  (e.g.  Mn  can  assume  oxidation  states  from  +VII  to  0  

or  even  -­‐I).  The  potential  at  which  the  one-­‐electron  steps  occur  are  a  sensitive  

function  of  the  nature  of  the  ligands  and  the  coordination  geometry  and  therefore  be  

easily  tuned  by  suitable  ligand  design.  Furthermore,  transition  metals  typically  

contain  unpaired  electrons  in  their  partially  filled  d-­‐shells  which  gives  rise  to  

several  accessible  spin  states.  These  spin  states  may  show  considerably  different  

reactivities  and  this  leads  to  added  complexity  in  studying  their  reactions.  

Frequently,  the  systems  even  pass  from  one  spin  state  to  another  surface  during  a  

given  catalytic  cycle  which  opens  many  fascinating  routes  to  unique  reactions.  Last  

but  not  least,  the  partially  filled  d-­‐shells  and  the  variable  metal-­‐ligand  bonds  are  

associated  with  a  huge  variety  of  fascinating  spectroscopic  phenomena.  Thus,  very  

detailed  insight  into  the  geometric  and  electronic  structure  of  the  transition  metal  

Page 2: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

sites  can  be  obtained  from  spectroscopic  studies  and  frequently  these  studies  yield  

the  key  to  understanding  the  unique  reactivities  displayed  by  transition  metal  

complexes.  In  this  Computer  Experiment  we  will  study  some  basic  aspects  of  

theoretical  coordination  chemistry.  We  will  stress  the  use  of  the  model  of  ligand-­‐

field  theory  (LFT)  as  a  conceptual  framework  in  which  the  results  of  actual  MO  

calculations  can  be  readily  and  successfully  interpreted.  

1.1.2 Ligand  Field  Theory  Beyond  all  advanced  theoretical  considerations,  the  major  driving  force  for  the  

formation  of  coordination  complexes  is  the  electrostatic  attraction  of  a  positively  

charged  metal  ion  and  a  negatively  charged  ligand  or  a  ligand  with  functional  groups  

that  bear  a  partial  negative  charge  –  this  is  nicely  seen  from  the  computed  

electrostatic  potential  of  the  typical  ligand    ehthylenediamine  (Figure  1).  

 Figure  1:  Principle  of  the  formation  of  coordination  compounds.  

Traditionally,  the  properties  of  such  coordination  compounds  are  interepreted  in  

the  framework  of  a  theory  that  was  worked  out  in  the  1950s  by  physicists  (mainly  

Hans  Bethe)  and  that  is  called  “crystal  field  theory”.  Slightly  more  chemistry  

oriented  modifications  are  known  collectively  as  “ligand  field  theory”  but  we  will  

not  distinguish  between  these  terms.  The  essential  idea  is  to  focus  the  interest  on  

the  open  d-­‐shells  of  the  metals  and  study  their  interaction  with  their  environment  

which  is  assumed  to  somehow  provide  an  electrostatic  field  in  which  the  d-­‐electrons  

have  to  move.  Thus,  LFT  takes  no  detailed  account  of  chemical  bonding  but  is  

enormously  successful  in  its  domain  of  applicability.  In  fact,  using  LFT  many  

properties  of  large  classes  of  complexes  can  be  understood  in  a  unified  and  simple  

language.  In  particular,  the  concept  of  a  “dN-­‐metal”  (N=number  of  d-­‐electrons)  that  

is  so  central  to  inorganic  chemistry  would  not  be  possible  without  ligand  field  

Page 3: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

theory.  The  concepts  are  usually  worked  out  on  the  example  of  octahedral  

coordination  complexes.  While  perfectly  octahedral  complexes  may  be  rather  rare,  

approximately  octahedral  complexes  are  highly  abundant  and  the  conclusions  

drawn  from  studying  the  octahedral  case  largely  remain  valid.    

The  shapes  of  the  d-­‐orbitals  are  well  known  and  are  shown  again  in  Figure  2.  The  

ligands  are  assumed  to  be  placed  along  the  axes  of  the  coordinate  system  such  that  

an  octahedral  field  of  negative  point  charges  arises.  This  set  of  negative  point  

charges  now  starts  to  interaction  with  a  “probe”  electron  that  is  placed  into  each  of  

the  five  d-­‐orbitals  in  turn.  Upon  evaluating  the  energy  of  interaction  of  the  probe  

electron  with  the  negative  point  charges1    it  is  found  that  the  energy  is  different  for  

the  different  orbitals  as  shown  in  Figure  3.  

 Figure   2:   The   shapes   of   the   five   d-­‐orbitals   together   with   their   common   designation   in   terms   of   real   spherical  harmonics.  The  combination  z2-­‐r2  is  usually  only  abbreviated  as  “z2”.  

Thus,  as  a  consequence  from  lowering  the  symmetry  of  the  surrounding  of  the  ion  

from  spherical  (in  the  free  ion)  to  octahedral  (in  the  complex),  the  degeneracy  of  the  

five  d-­‐orbitals  is  lifted.  In  fact,  it  is  plausible  from    Figure  3  that  an  electron  that  

occupied  the  dx2-­‐y2  or  the  dz2  orbital  is  –  on  average  –  more  strongly  repelled  by  the  

negatively  charged    ligand  than  an  electron  that  occupied  the  dxy,  dxz  or  dyz  orbitals.  

A  quantitative  treatment  (and  in  fact  already  the  group  theoretical  treatment)  

                                                                                                               1  This  will  not  be  done  here.  The  classic  and  rigorous  mathematical  text  is  due  to  J.S.  Griffith  The  Theory  of  Transition-­‐Metal   Ions,   Cambridge   University   Press,   Cambridge,   1964.   It   may   appear   that   this   is   an   outdated   book   and   an  outdated   theory   –   but   its   not!   You   will   find   many   insights   that   rigorously   carry   over   to   the   modern   methods   of  quantum  chemistry.  A  thorough  understanding  of  ligand  field  theory  teaches  you  an  enormous  amount  about  the  N-­‐electron  problem  in  quantum  chemistry.  The  intrinsic  value  of  this  understanding  can  hardly  be  overemphasized.  

Page 4: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

shows,  that  the  set  of  five  d-­‐orbitals  fall  into  a  t2g  subshell  at  lower  energy  and  an  eg  

shell  at  higher  energy.2  

 Figure  3:  d-­‐Orbital  Splittings  induced  by  an  octahedral  ligand  field.  

Following  the  rules  of  how  to  assign  the  many  electron  term  symbol  from  the  one-­‐

electron  MO  scheme,  it  is  evident  that  the  configuration  (t2g)1(eg)0  gives  rise  to  a  2T2g  

term3  while  to  higher  energy  we  expect  the  configuration  (t2g)0(eg)1  which  gives  rise  

to  a  2Eg  term  in  Oh  symmetry.  Since  the  ‘T’  irreducible  representation  is  three-­‐

dimensional  and  ‘E’  is  two-­‐dimensional  we  say  that  2T2g  is  threefold  orbitally  

degenerate  and  2Eg  is  two-­‐fold  orbitally  degenerate.  The  total  degeneracy  of  the  

state  in  the  absence  of  spin-­‐orbit  coupling  (SOC)  and  magnetic  fields  is  (2S+1)  times  

the  orbital  degeneracy  which  amounts  to  six  for  2T2g  and  four  for  2Eg  in  the  present  

case.  All  of  these  ten  states  are  observable  and  can  be  probed  by  spectroscopic  

experiments.  One  of  the  simplest  experiments  is  to  induce  a  2T2g→2Eg  transition,  by  

application  of  a  visible  photon  of  suitable  energy.4  The  experiment  provides  a  

quantitative  measure  of  the  d-­‐orbital  splitting  Δ0  which  is  typically  in  the  range  1-­‐4  

eV  (8000-­‐32000  cm-­‐1).  Already  the  example  of  [Ti(H2O)6]3+  in  Figure  4  shows  that  this  

                                                                                                               2  Note  carefully,  that  these  energies  are  NOT  drawn  to  scale.  Above  all,  the  repulsion  leads  to  a  strong  increase  of  the  energies  of  all  five  d-­‐orbitals  and  the  relatively  small  difference  in  the  energy  increase  is  the  ligand-­‐field  splitting.  Note  also,   that   the   final   result   does   not   depend   on   how  we  place   the   coordinate   system.   Thus,   the   result   is   said   to   be  rotationally  invariant.  What  does  change  if  we  rotate  the  coordinate  system  are  only  the  designation  of  the  orbitals.    3  Note:  capital  letters  refer  to  many  electron  states  and  lowercase  letters  to  one-­‐electron  orbitals.  Furthermore,  the  term  symbol  carries  as  a  left  superscript  the  total  multiplicity  of  the  state  which  is  2S+1  where  S   is  the  total  spin  of  this  state.  4  This  is  only  true  on  a  superficial  level.  In  fact,  the  optical  spectroscopy  of  transition  metal  complexes  becomes  very  complicated  and  very  subtle  rather  quickly  as  one  enters  into  more  detail  –  try  it,  its  worthwhile!  

Page 5: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

is  a  slight  oversimplification  since  in  reality  we  observe  a  broad  and  asymmetric  

absorption  band  rather  than  a  single  line  like  in  atomic  spectroscopy.    

 Figure  4:  Example  of  the  measurement  of  the  ligand  field  splitting  in  [Ti(H2O)6]

3+  

However,  we  will  not  go  into  any  further  detail  of  the  interpretation  of  the  spectrum  

here  but  merely  pretend  that  it  is  possible  extract  reasonable  values  of  the  ligand-­‐

field  splitting  Δ0  from  the  study  of  the  absorption  spectra  of  the  complexes.  Since  the  

magnitude  of  the  splitting  depends  on  the  nature  of  the  ligand,  the  ability  of  the  

ligand  to  induce  d-­‐orbital  splittings  may  be  classified  in  the  spectrochemical  series  

shown  below.  

 Figure  5:  The  spectrochemical  series.  

On  this  basis  CN-­‐  is  known  to  be  a  “strong-­‐field  ligand”  since  it  gives  rise  to  large  

ligand  field  splittings  and  I-­‐  is  known  to  be  a  “weak  field  ligand”  since  it  gives  rise  to  

small  ligand  field  splittings.5  

A  nice  experimental  confirmation  that  this  way  of  thinking  about  the  electronic  

structure  of  transiton  metal  complexes  makes  sense  is  shown  in  Figure  6  where  the  

experimental  enthalpies  of  formation  of  [M(H2O)6]2+  complexes  is  shown  (M=first  

row  transition  metal).                                                                                                                  5  Please   observe   some   irregularities   here:   hydroxide   is   a   weaker   field   ligand   than   H2O   despite   the   fact   that   it   is  actually  negatively  charged.  This  should  not  be  the  case  if  our  reasoning  about  the  origin  of  the  ligand  field  splitting  above  would  be  correct.  We  will  come  back  to  this  later.    

Page 6: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 Figure  6:  Thermochemical  confirmation  of  ligand  field  splittings:  the  ligand  field  stabilization  energy  (LFSE).  Plotted  are   the  enthalpies  of   formation  of   the  hexaquo  complexes   for   the  divalent   first-­‐row  transition  metal   ions.  Open  circles  –  raw  experimental  data.  Full  circles  –  after  subtraction  of  the  LSFE.  

The  interpretation  of  the  results  is  as  follows:  The  enthalpy  of  formation  is  

increasing  along  the  series.  Thus,  the  later  transiton  metals  hold  the  water  ligands  

more  strongly  than  the  early  transition  metals.  This  is  first  of  all  an  effect  of  the  

effective  nuclear  charge  of  the  metal.  Since  the  d-­‐electrons  shield  the  increasing  

nuclear  charge  with  increasing  atomic  number  less  effectively,  the  ligands  “see”  an  

increased  positive  charge  from  the  metal  which  leads  to  the  slope  of  the  line.  

However,  more  interestingly  is  the  “double-­‐bowl”  behavior  of  the  experimental  data.  

This  readily  find  an  explanation  from  considering  the  energy  of  a  given  dN  

configuration  relative  to  the  hypothetical  case  where  all  d-­‐orbitals  have  the  same  

energy:  thus  occupation  of  a  t2g  level  leads,  on  average,  to  an  energy  decrease  of  -­‐

2/5Δ0  while  occupation  of  an  eg  level  leads  to  a  3/5Δ0  increase.  Knowing  the  Δ0-­‐

values  from  optical  spectroscopy,  the  LFSE  of  a  given  dN  configuration  can  be  

subtracted  from  the  experimental  data  to  arrive  at  the  full  circles  which  now  

perfectly  fall  on  a  straight  line.  Thus,  in  this  way  one  has  found  a  simple  explanation  

for  a  seemingly  puzzling  and  irregular  experimental  result.    

1.1.3 Electron-­‐Electron  Repulsion  Effects  A  little  contemplation  shows  that  our  discussion  of  the  energies  of  given  dN  

configurations  is  somewhat  oversimplified  since  up  to  now  we  have  not  taken  

account  of  the  basic  fact  that  electrons  are  negatively  charged  and  actually  repel  

Page 7: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

each  other.6  This  leads  to  a  number  of  interesting  effects  in  coordination  chemistry.  

First  of  all,  two  rules-­‐of-­‐thumb  should  be  mentioned:  (a)  electrons  that  occupied  the  

same  orbital  repel  each  other  most  strongly,  (b)  Electrons  of  opposite  spin  repel  

each  other  more  strongly  than  electrons  of  the  same  spin.  Thus,  the  system  will  

want  to  assume  the  configuration  where  –  as  much  as  possible  –  the  electrons  have  

the  same  spin  and  occupy  different  orbitals.  This  leads  to  interesting  multiplet  

splittings  in  the  dN  configurations  where  states  of  different  multiplicity  but  the  same  

electron  configuration  have  different  energies.    

 Figure  7:  Principles  of  electron-­‐electron  repulsion  in  LFT.  

In  LFT,  one  has  the  fortunate  situation  that  electron-­‐electron  repulsion  can  be  

described  to  a  good  approximation  with  a  single  parameter:  the  “Racah  B-­‐value”.  

Thus,  besides  the  ligand-­‐field  splitting  parameter  Δ0,  the  second  important  

parameter  of  LFT  is  the  B-­‐value.  It  is  of  the  order  of  ~1000  cm-­‐1  and  state  splittings  

are  typically  multiples  of  B.  The  example  in  Figure  7  is  the  1T1g-­‐3T2g  splitting  (both  

from  the  (t2g)2(eg)0  configuration!)  is  4B  and  experimentally  the  splitting  between  

these  states  is  3000-­‐4000  cm-­‐1.    

The  electron-­‐electron  repulsion  has  a  very  interesting  consequence:  for  several  of  

the  dN  configurations  (actually  d4-­‐d7)  it  is  not  a  priori  clear  which  electron  

configuration  will  be  of  lowest  energy  since  there  is  a  competition  between  ligand-­‐

field  splitting  and  electron-­‐electron  repulsion  (Figure  8).  For  example,  for  the  case  of  

                                                                                                               6  As  mentioned  in  the  introduction,  the  electron-­‐electron  repulsion  is  the  biggest  challenge  for  theoretical  chemistry.  Its   accurate   calculation   from   first   physical   principles   (the   “electron   correlation   problem”)   is   very   difficult   and  represents  a  major  branch  of  research  in  quantum  chemistry.  

Page 8: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

the  d5  configuration,  the  high-­‐spin  state  ((t2g)3(eg)2;  6A1g)  minimizes  the  electron-­‐

electron  repulsion  according  to  our  rules  of  thumb  but  at  the  expense  of  promoting  

two  electrons  into  the  higher-­‐energy  eg-­‐orbitals.  This  can  be  avoided  by  forming  the  

alternative  low-­‐spin  ((t2g)5(eg)0;  2T2g)  configuration  which  maximizes  the  LSFE  but  at  

the  expense  of  a  larger  electron-­‐electron  repulsion.  Thus,  depending  on  the  ligand  

environment  either  one-­‐  or  the  other  alternative  will  be  preferred:  strong  field  

ligands  lead  to  low-­‐spin  complexes  and  weak  field  ligands  to  high-­‐spin  complexes.  

Experimentally,  this  is  found  to  be  true:  [Fe(CN)6]3-­‐  has  a  2T2g  ground  state  and  

[FeCl6]3-­‐  has  a  6A1g  ground  state.    

 Figure  8:  Competition  between  electron-­‐electron  repulsion  and  ligand  field  splitting  to  produce  either  high-­‐spin  or  low-­‐spin  complexes.  

Everything  that  has  been  said  about  the  nature  of  the  two  dominating  influences:  

ligand  field  splitting  and  electron-­‐electron  repulsion  can  be  nicely  summarized  and  

systematized  in  the  so-­‐called  Tanabe-­‐Sugano  diagrams  (Figure  9).  What  is  plotted  

there  on  the  X-­‐axis  is  the  ligand  field  strength  in  units  of  B  while  on  the  y-­‐axis  the  

energy  of  a  given  state  relative  to  the  lowest  state  of  the  system  is  plotted.    

Page 9: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 Figure  9:  General  organization  of  a  Tanabe-­‐Sugano  diagram.  

For  the  d4-­‐d7  electron  configurations,  the  Tanabe-­‐Sugano  diagrams  have  two  parts:  

a  weak  field  part,  where  the  high-­‐spin  states  is  the  ground  state  and  a  strong  field  

part,  where  the  low-­‐spin  state  is  the  ground  state.  At  some  critical  ligand  field  

strength,  a  discontinous  change  from  one  ground  state  to  the  other  occurs.  

Experimentally,  this  is  known  to  be  possible:  complexes  with  ligands  that  lead  to  

ligand  field  strengths  close  to  the  crossover  point  are  know  to  be  switchable  from  

one  ground  state  to  the  other  by  external  perturbation  such  as  pressure.  This  leads  

to  the  interesting  field  of  “spin-­‐crossover”  chemistry  which  is  a  promising  approach  

to  the  design  of  molecular  switches.    

Secondly,  the  Tanabe-­‐Sugano  diagrams  are  excellent  guides  to  the  assignment  of  the  

ligand-­‐field  absorption  spectra  of  transiton  metal  complexes.  Given,  that  the  

selection  rule  ΔS=0  for  an  electric  dipole  allowed  transition  between  two  states  is  

fairly  strong,  one  simply  has  to  search  in  the  Tanabe-­‐Sugano  diagram  for  states  of  

the  same  multiplicity  as  the  ground  state  in  order  to  find  out  how  many  ligand-­‐field  

bands  are  expected  in  the  absorption  spectrum.7  The  only  thing  that  is  needed  then  

in  order  to  determine  Δ0  is  a  ruler  and  a  reasonable  estimate  of  B.  Two  examples  of  

such  diagrams  together  with  actual  absorption  spectra  are  shown  in  Figure  10.  

                                                                                                               7  Actually,  even  the  spin-­‐allowed  ligand  field  bands  are  weak  since  they  are  parity  forbidden  in  octahedral  complexes.  For  example,  in  tetrahedral  complexes  there  is  no  center  of  inversion  and  the  observed  ligand  field  bands  are  much  stronger   than   in   octahedral   complexes.   Spin-­‐allowed   ligand   field   bands  may   have   ε’s   of   several   hundred  M-­‐1  cm-­‐1  while  spin-­‐forbidden  ones  typically  have  ε<1  M-­‐1  cm-­‐1    

Page 10: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 Figure   10:   Example   of   a   ligand   field   analysis   for   the   d2   and   d3configurations   on   the   basis   of   Tanabe-­‐Sugano  diagrams.    

A  summary  of  the  many  electron  terms  together  with  the  leading  one-­‐electron  

configurations  is  shown  in  Table  1.  Table  1:  Ground-­‐  and  Ligand  Field  Excited  States  of  Octahedral  dN  Complexes.  

dN   Ground  Term   Spin-­‐Allowed  Transitions   Spin-­‐Forbidden  Transitions     Oh   Config.      d9   2Eg   t2g6eg3   2T2g(t2g5eg4)   None  

d8   3A2g   t2g6eg2   3T2g(t2g5eg3)  3T1g(t2g5eg3)  3T1g(t2g4eg4)    1Eg(t2g6eg2)  1T2g(t2g5eg3)  

d7   4T1g   t2g5eg2   4T2g(t2g4eg3)  4T1g(t2g4eg3)  4A2g(t2g3eg4)    2Eg(t2g6eg1)  2A2g(t2g5eg2)  2T2g(t2g5eg2)    

d6   5T2g   t2g4eg2   5Eg(t2g3eg3)    3T2g(t2g5eg1)  

d5   6A1g   t2g3eg2   None    4T1g,  4T2g  (t2g4eg1)  4A1g,4Eg,4T2g(t2g3eg2)  

d4   5Eg   t2g3eg1   5T2g(t2g2eg2)    3T1g(t2g2eg2)    

d3   4A2g   t2g3eg0   4T2g(t2g2eg1)  4T1g(t2g2eg1)  4T1g(t2g1eg2)    2Eg(t2g3)  2T1g(t2g3)  

d2   3T1g   t2g2eg0   3T2g(t2g1eg1)  3T1g(t2g1eg1)    1Eg(t2g2)  1T2g(t2g2)  1A1g(t2g2)  

d1   2T2g   t2g1eg0   2Eg(t2g0eg1)   None  

Page 11: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 

1.1.4 Other  Coordination  Geometries  The  entire  chain  of  arguments  developed  so  far  is  only  valid  for  octahedral  

complexes.  However,  in  reality,  coordination  compounds  exist  in  many  different  

effective  geometries  with  coordination  numbers  ranging  from  2-­‐8.  Each  distinct  

coordination  geometry  induces  a  different  orbital  splitting  pattern  and  therefore  

also  a  different  set  of  molecular  multiplets.  For  some  of  the  more  important  

coordination  geometries  the  orbital  splitting  patterns  are  shown  in  Figure  11.  

 Figure  11:  Orbital  splitting  patterns  in  different  coordination  geometries.  

While  each  of  these  geometries  would  require  an  elaborate  analysis,  it  is  clear  that  

the  absorption  spectra  of  coordination  complexes  react  very  sensitively  to  the  

coordination  geometry  and  consequently,  the  coordination  geometry  can  be  

inferred  from  a  proper  analysis  of  the  absorption  spectra  (or  other  optical  

techniques  that  probe  the  same  transitions).  We  will  not  go  into  more  detail  here.  

 

1.1.5 Inadequacy   of   Ligand   Field   Theory,   Covalency,   Nephelauxetic        Effect  

Below  you  will  find  a  quote  from  one  of  the  father’s  of  ligand  field  theory  and  the  

dominating  spectroscopist  of  his  time:  

Page 12: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 What  Jörgensen  refers  to  so  elaborately  as  “electrostatic  romantics”  is  the  crystal  

field  model  that  we  have  just  discussed  in  the  previous  section.  Indeed,  there  are  

very  good  evidences  from  spectroscopic  measurements  that  LFT  is  an  

oversimplification.  For  example,  consider  the  following  EPR  experiment:  

 Figure   12:   The   EPR   spectrim   of   [Cu(imidazole)4]

2+   (left).   On   the   right   is   the   singly   occupied   orbital   according   to  ligand  field  theory.  On  this  basis,  the  sharp  features  in  the  middle  of  the  spectrum  can  not  be  explained.  

What  you  see  there  is  the  EPR  spectrum  of  [Cu(imidazole)4]2+  -­‐  a  typical  

coordination  complex.  It  is  a  typical  square  planar  d9  complex  like  many  others  in  

coordination  chemistry.  According  to  Figure  11  the  unpaired  electron  resides  in  a  

metal  dx2-­‐y2  orbital.  The  interesting  sharp  features  in  the  EPR  spectrum  actually  

represent  hyperfine  structure  that  arise  from  coupling  of  the  unpaired  electron  to  

four  equivalent  nitrogen  nuclei  of  the  imidazole  ligand.8  Since  we  learned  in  

experiment  number  9  (section  Error!  Reference  source  not  found.,  page  Error!  

Bookmark  not  defined.)  that  the  hyperfine  structure  is  a  measure  for  the  

probability  to  meet  the  unpaired  electron  at  the  nucleus  in  question  there  is  a  

                                                                                                               8  It  is  an  interesting  question  to  ask  to  the  chemist  how  one  can  proof  the  origin  of  these  signals?  The  answer  is  the  following:  nitrogen  has  two  accessible  isotopes  14N  (the  naturally  most  abundant)  and  15N.  14N  has  a  nuclear  spin  of  I=1  while  15N  has  I=1/2.  Furthermore  the  nuclear  g-­‐value  of  14N  is  twice  as  large  as  that  of  15N.  Thus,  synthesis  of  the  imidazole  ligand  with  14N  and  15N  will  lead  to  precisely  predictable  changes  in  the  EPR  spectrum  and  this  is  how  the  nature  of  the  signals  can  be  experimentally  verified  without  any  doubt.    

Page 13: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

mismatch  between  the  predictions  of  LFT  and  the  experimental  observations:  

According  to  LFT  the  SOMO  is  a  pure  metal  orbital  and  there  is  no  probability  to  find  

the  electron  at  the  nitrogen  nucleus  –  yet,  the  experiment  proofs  the  contrary.    

In  fact,  this  experimental  finding  is  readily  explained  on  the  basis  of  MO  theory.  The  

shape  of  the  SOMO  is  shown  in  Figure  13.  It  is  evident  that  the  SOMO  indeed  has  dx2-­‐y2  

character  on  the  central  copper.  However,  the  orbital  is  σ-­‐antibonding  with  the  

coordinating  imidazole-­‐nitrogens.  Thus,  part  of  the  unpaired  spin-­‐density  is  

delocalized  into  these  ligand  atoms  and  this  accounts  for  the  experimental  

observations.  In  fact,  from  the  experimental  data  one  is  able  to  make  a  semi-­‐

quantiative  estimate  of  how  much  spin  is  required  to  be  on  the  nitrogens  in  order  to  

explain  the  experimental  data.  In  the  present  case,  this  amount  to  ~60-­‐65%  spin  

population  on  the  copper  with  some  ~9-­‐10%  on  each  nitrogen.  

 Figure  13:  Interpretation  of  the  EPR  spectrum  of  [Cu(imidazole)4]

2+  according  to  MO  theory.  

The  “dilution”  of  the  metal  orbitals  with  ligand  character  has  in  fact  also  been  noted  

by  the  optical  spectroscopists  interested  in  LFT.  They  needed  to  account  for  

electron-­‐electron  repulsions  that  were  smaller  than  expected  from  atomic  

spectroscopy.  They  explained  this  finding  by  an  increase  of  the  size  of  the  metal  

orbitals  and  called  the  effect  “nephelauxetic  effect”  (cloud  expanding  effect).  

However,  it  finds  a  much  more  natural  explanation  in  terms  of  MO  theory.  In  fact,  

the  mixing  of  ligand  character  into  the  metal  orbitals  is  referred  to  as  metal-­‐ligand  

covalency.  The  more  covalent  the  metal-­‐ligand  bond,  the  more  ligand  character  is  

mixed  into  the  formally  metal  based  MOs.    

Page 14: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

1.1.6 Molecular  Orbitals  of  Octahedral  Complexes  The  discussion  above  raises  the  question  of  how  the  MO  theory  of  coordination  

complexes  works  in  general.  This  has  been  worked  out  in  great  detail  by  many  

authors  and  it  can  be  found  in  many  textbooks.  Therefore,  we  will  be  very  brief  in  

this  respect.    

A  typical  MO  scheme  of  a  ML6  complex  as  it  could  be  derived  from  extended  Hückel  

calculations9.    

Thus,  one  considers  that  ionization  potentials  of  transition  metals  are  typically  a  few  

eV  lower  than  those  of  main  group  atoms.  Consequently,  one  place  the  metal  

orbitals  nd  (n+1)s  and  (n+1)p  to  0th  order  at  higher  energy  than  the  ligand  p-­‐

orbitals.  Perturbation  theory  then  tells  us,  that  upon  forming  molecular  orbitals  

from  fragment  orbitals  of  correct  symmetry,  that  the  higher  lying  orbitals  become  

destabilized  and  antibonding  and  the  lower  lying  orbitals  become  stabilized  and  

bonding.  Furthermore,  the  smaller  the  energy  gap  between  the  mixing  fragment  

orbitals  and  the  stronger  their  interaction10,  the  larger  the  admixture.  Thus,  the  

metal  d-­‐orbitals  of  ligand  field  theory  are  the  partially  occupied  antibonding  orbitals  

that  can  be  found  in  MO  calculations  on  coordination  compounds.  Some  typical  

bonding  and  antibonding  components  are  shown  in  Figure  14.  It  is  evident  that  the  

metal-­‐based  t2g-­‐orbitals  correspond  to  π-­‐antibonding  MOs  while  the  metal-­‐based  eg-­‐

orbitals  correspond  to  σ-­‐antibonding  MOs.  Since  π-­‐interactions  are  typically  weaker  

than  σ-­‐interactions  this  accounts  for  the  energetic  ordering  of  the  two-­‐sets  of  

orbitals  that  was  so  luckily  correctly  predicted  by  LFT.  

                                                                                                               9  An  interesting  and  subtle  point   is  that  this  MO  scheme  could  not  be  derived  from  Hartree-­‐Fock  calculations.   In  HF  calculations,  the  metal  orbitals  typically  fall  energetically  below  the  ligand  orbitals  but  are  still  the  first  to  be  ionized.  This  arises  because  the  electronic  relaxation  energy  upon  ionization  is  typically  small  for  the  ligand  orbitals  and  very  large  for  the  metal  orbitals.  Consequently,  the  extended  Hückel  like  arguments  give  a  somewhat  correct  picture  of  the  ionization  events  in  complexes  but  the  quantitative  theory  is  more  complicated.  10  In   extended  Hückel   and  other   semiempirical   theories,   this   interaction   is   taken   to  be  proportional   to   the  overlap  between   the   metal   and   ligand   fragment   orbitals.   (on   also   refers   to   these   interaction   parameters   as   “resonance  integrals”)  

Page 15: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 Figure  14:  MOs  of  octahedral  ML6  complexes.  The  red  box  shows  the  five  metal-­‐based  MOs  that  are  the  ones  taken  into  account  in  ligand  field  theory.  

It  is  also  clear  that  since  the  M-­‐L  bonding  orbitals  are  always  fully  occupied  and  the  

metal  based-­‐d-­‐orbitals  are  only  partially  occupied  that  there  is  a  net  transfer  of  

charge  from  the  ligand  to  the  metal.  Thus,  in  MO  theories  one  never  finds  partial  

charges  that  come  close  the  formal  charges  as  measured  by  the  oxidation  state.  

However,  as  long  as  the  mixing  between  metal-­‐  and  ligand-­‐orbitals  is  not  extremely  

excessive11  the  concept  of  a  dN  configuration  that  is  so  central  to  LFT  still  remains  

valid  –  one  simply  has  to  realize  that  the  orbitals  that  one  talks  about  are  partially  

delocalized  onto  the  ligands.  From  this  discussion  it  also  becomes  evident  that  it  

does  not  make  sense  to  try  to  compare  the  partial  charges  from  a  MO  calculation  of  

whatever  kind  to  the  charge  implied  by  the  formal  oxidation  state.  

1.1.7 Pi-­‐Bonding  and  Pi-­‐Backbonding  There  is  one  important  class  of  ligands  that  do  not  perfectly  correspond  to  the  

discussion  given  above.  These  are  ligands  that  have  low-­‐lying  empty  π*-­‐orbitals.  

Such  orbitals  can  also  interact  with  the  metal-­‐based  t2g  orbitals  to  form  π-­‐bonds.  

                                                                                                               11  For  example  exceeding  30-­‐40%  ligand  character  in  the  metal  orbitals  

Page 16: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

However,  in  this  case,  the  nature  of  the  donor  and  acceptor  orbitals  is  reversed  

compared  to  the  standard  π-­‐donor  ligands  discussed  above.  Thus,  the  interaction  

stabilizes  the  metal  t2g  orbitals  that  now  become  bonding  rather  than  antibonding.  

Furthermore  occupation  of  these  orbitals  leads  to  a  backflow  of  charge  from  the  

metal  to  the  ligand  –  this  type  of  bonding  is  therefore  called  backbonding.  

 Figure  15:  orbitals  in  π -­‐donor  and  π -­‐acceptor  complexes.  

 The  effect  is  exemplified  in    Figure  15  –  the  standard  π-­‐donor  complex  [FeCl6]3-­‐  

(high-­‐spin  d5)  features  metal  t2g  based  orbitals  that  are  π-­‐antibonding  while  those  of  

[Cr(CO)6]  (low-­‐spin  d6)  are  π-­‐bonding  with  CO-­‐π*  orbitals.  

The  effect  of  π-­‐backbonding  naturally  has  a  very  pronounced  effect  on  the  ligand-­‐

field  splitting.  Thus,  since  the  metal  based  t2g-­‐orbitals  are  stabilized  through  

backbonding  while  the  metal-­‐eg  based  orbitals  are  destabilized  due  to  σ-­‐bonding,  the  

net  effect  is  an  increase  in  the  ligand  field  splitting.  The  opposite  effect  is  of  course  

true  for  standard  π-­‐donor  complexes  –  the  better  the  π-­‐bond  between  the  metal-­‐  

and  the  ligand  the  smaller  the  ligand  field  splitting.  On  this  basis  one  comes  now  to  a  

much  more  appealing  interpretation  of  the  spectrochemical  series  (Figure  16):  strong  

field  ligands  are  π-­‐acceptor  ligands  while  weak  field  ligands  are  good  π-­‐donor  

ligands.  In  the  middle  of  the  spectrochemical  series  one  simply  finds  ligands  that  

have  no  π-­‐orbitals  for  either  bonding  or  backbonding  available  (for  example  NH3  or  

neutral  aliphatic  amines).    

Page 17: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 Figure  16:  Interpretation  of  the  spectrochemical  series  on  the  basis  of  MO  theory.  

1.2 Description  of  the  Experiment  The  area  of  theoretical  coordination  chemistry  is  very  large.  Although  this  section  is  

somewhat  lengthy,  it  is  not  possible  to  do  more  than  to  scratch  the  surface  of  it  in  

this  course.  Below  you  will  find  some  problems  in  theoretical  coordination  

chemistry  that  are  more  or  less  representative.  It  will  not  be  possible  to  work  out  all  

of  these  problems.  It  is  suggested  to  work  out  the  first  problem  (1.2.1)  and  pick  two  

of  the  larger  remaining  ones.  Problems  1.2.2  and  1.2.5  are  small  and  require  less  

work.  If  these  are  chosen,  at  least  one  additional  larger  problem  should  be  

investigated.  

In  this  section  we  will  use  very  small  basis  sets  (SV  for  the  metal  and  the  oxygens  

and  STO-­‐3G  for  the  hydrogens)  in  order  to  keep  the  calculations  fast.  In  practice  it  is  

advisable  to  use  at  least  polarized  triple-­‐ζ  basis  sets  for  the  metals  and  at  least  

polarized  double-­‐ζ  basis  sets  for  the  ligands  in  order  to  obtain  reliable  results.    

As  a  generic  “template”  for  our  calculations,  we  will  use  approximately  octahedral  

fragments  with  water  ligands.  Appropriate  metal-­‐ligand  bond  distances  are  shown  

in  Table  2  for  M(II)  and  M(III)  ions  of  the  first  transition  row.  The  complexes  studied  

will  be  of  the  form  [M(H2O)5X]n+  where  M  is  a  first-­‐row  transition  ion  and  X  is  an  

additional  ligand  that  will  be  varied  in  the  course  of  the  experiment.  For  X=H2O,  the  

Page 18: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

highest  possible  symmetry  that  such  complexes  can  assume  is  Th  which  is  a  genuine  

cubic  group.    

Table  2:  Bond  distances  (Angström)  for  divalent  and  trivalent  hexaquo  ions  of  the  first  transiton  row.  

  Divalent   Trivalent  Ti   -­‐   2.028  V   2.128   1.992  Cr   2.052   1.959  Mn  

2.192   1.991  

Fe   2.114   1.995  Co   2.106   1.873  Ni   2.061   -­‐  Cu   1.964   -­‐    A  Z-­‐matrix  for  cubic  symmetry  is  shown  below  in  ORCA  format  (the  example  chosen  

is  that  of  [Ni(H2O)6]2+  with  a  S=1  ground  state).  

Page 19: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 

1.2.1 Some  aspects  of  Ligand  Field  Theory  with  DFT  Application  of  DFT  to  transition  metal  complexes  is  often  successful  but  not  without  

its  pitfalls.  One  problem  in  the  present  case,  where  we  deal  with  highly  symmetric  

complexes  is  that  DFT  only  applies  to  orbitally  nondegenerate  ground  states.  For  

degenerate  ground  states,  symmetry  breakings  occur  which  may  lead  to  artifacts  in  

the  calculations.12  Nevertheless,  in  this  course,  we  will  focus  on  dN  configurations  

with  orbitally  nondegenerate  ground  states.  

Do  a  calculation  on  the  complex  [Cr(H2O)6]3+.  The  setup  of  the  job  can  be  done  as  

follows:  

 

                                                                                                               12  In   practice   this   is   usually   not   such   a   strong   limitation   since  most   systems   of   practical   interest   are   not   orbitally  degenerate.   Even   those   that   are   orbitally   degenerate   are   Jahn-­‐Teller   active   and   become   orbitally   nondegenerate  upon  distortion.    

%paras R_ML 2.061 # the metal-water-oxygen distance R_OH 0.9817 # the water O-H distance A_HOM 126.153 # the angle H-O-Metal R_MX 2.061 # the distance of the metal and the # external ligand. Here another water end * int 2 3 # The central metal: here nickel Ni 0 0 0 0.0 0.0 0.0 # the five coording water oxygens O 1 0 0 {R_ML} 0.0 0.0 O 1 2 0 {R_ML} 180.0 0.0 O 1 2 3 {R_ML} 90.0 0.0 O 1 2 4 {R_ML} 90.0 180.0 O 1 2 4 {R_ML} 90.0 270.0 # this is the coordinating atom of the external ligand X O 1 2 4 {R_MX} 90.0 90.0 # These are the hydrogens of the water ligands of M(H2O)5 H 2 1 3 {R_OH} {A_HOM} 90.0 H 2 1 3 {R_OH} {A_HOM} 270.0 H 3 1 2 {R_OH} {A_HOM} 90.0 H 3 1 2 {R_OH} {A_HOM} 270.0 H 4 1 2 {R_OH} {A_HOM} 0.0 H 4 1 2 {R_OH} {A_HOM} 180.0 H 5 1 2 {R_OH} {A_HOM} 180.0 H 5 1 2 {R_OH} {A_HOM} 0.0 H 6 1 2 {R_OH} {A_HOM} 90.0 H 6 1 2 {R_OH} {A_HOM} 270.0 # here come the remaining atoms of the external ligand X H 7 1 2 {R_OH} {A_HOM} 90.0 H 7 1 2 {R_OH} {A_HOM} 270.0 *

Page 20: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 One  common  observation  with  calculations  on  transition  metal  complexes  is  that  

the  SCF  calculations  do  not  converge  well  and  some  precautions  are  necessary.13  

Now  look  at  the  orbitals.  In  spin-­‐unrestricted  calculations  we  have  separate  spin-­‐up  

and  spin-­‐down  MOs.  In  the  present  case  we  have  a  d3  system  and  expect  that  the  

three  spin-­‐up  HOMOs  are  mainly  metal  t2g  in  character  and  the  two  spin-­‐up  LUMOs  

will  have  metal  eg  character  –  let  us  verify  that  this  is  so!  Thus,  we  look  at  the  spin-­‐

up  HOMO/LUMO  gap:   39 1.0000 -0.759449 -20.6651 40 1.0000 -0.759446 -20.6651 41 1.0000 -0.759445 -20.6650 42 0.0000 -0.655032 -17.8239 43 0.0000 -0.655029 -17.8238

We  observe  that  the  three  highest  occupied  and  two  lowest  unoccupied  orbitals  are  

degenerate  (as  expected).  Now  let  us  look  at  their  composition:   36 37 38 39 40 41 -0.82938 -0.82938 -0.82938 -0.75945 -0.75945 -0.75945 1.00000 1.00000 1.00000 1.00000 1.00000 1.00000 -------- -------- -------- -------- -------- -------- 0 Cr dxz 0.0 0.0 0.0 12.9 56.4 0.2 0 Cr dyz 0.0 0.0 0.0 0.2 0.1 69.2 0 Cr dxy 0.0 0.0 0.0 56.4 13.0 0.1 -0.65503 -0.65503 -0.49107 -0.29188 -0.29187 -0.29187 0.00000 0.00000 0.00000 0.00000 0.00000 0.00000 -------- -------- -------- -------- -------- -------- 0 Cr s 0.0 0.0 61.6 0.0 0.0 0.0 0 Cr dz2 5.2 71.0 0.0 0.0 0.0 0.0 0 Cr dx2y2 71.0 5.2 0.0 0.0 0.0 0.0

 

                                                                                                               13  Above  we  have  chosen  a  recipe  that  is  typical  for  many  calculations  on  transition  metal  complexes  with  ORCA  (The  keyword  “SlowConv”  indicates  that  you  expect  slow  convergence).  In  general  obtaining  convergence  requires  some  experimentation.  One  commonly  used  trick  is  to  use  level  shifting  (like  in  the  example  above).      

! UKS BP RI SlowConv NoFinalGrid SV(P) %scf shift shift 0.3 erroff 0 end end # geometric parameters %paras R_ML = 1.959 # the metal-water-oxygen distance R_OH = 0.9817 # the water O-H distance A_HOM = 126.153 # the angle H-O-Metal R_MX = 1.959 # the distance of the metal and the # external ligand. Here another water end

Page 21: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

Indeed,  the  result  is  the  one  that  we  wanted  as  we  see  from  the  compositions  of  the  

orbitals.14  Likewise,  the  five  spin-­‐down  LUMOs  are  expected  to  represent  the  spin-­‐

down  counterparts  of  the  metal  t2g  and  metal  eg  orbitals  –  Let’s  verify.      

The  first  unoccupied  spin-­‐down  orbitals:   39 0.0000 -0.652195 -17.7467 40 0.0000 -0.652191 -17.7466 41 0.0000 -0.652190 -17.7466 42 0.0000 -0.605269 -16.4698 43 0.0000 -0.605267 -16.4697

And  their  compositions:   36 37 38 39 40 41 -0.82555 -0.82555 -0.82554 -0.65219 -0.65219 -0.65219 1.00000 1.00000 1.00000 0.00000 0.00000 0.00000 -------- -------- -------- -------- -------- -------- 0 Cr dxz 0.0 0.0 0.0 16.2 62.3 11.6 0 Cr dyz 0.0 0.0 0.0 0.8 11.2 78.1 0 Cr dxy 0.0 0.0 0.0 73.1 16.6 0.4 42 43 44 45 46 47 -0.60527 -0.60527 -0.48831 -0.29023 -0.29022 -0.29022 0.00000 0.00000 0.00000 0.00000 0.00000 0.00000 -------- -------- -------- -------- -------- -------- 0 Cr s 0.0 0.0 62.3 0.0 0.0 0.0 0 Cr dz2 6.9 73.8 0.0 0.0 0.0 0.0 0 Cr dx2y2 73.8 6.9 0.0 0.0 0.0 0.0

 Interestingly,  the  orbital  splittings  in  the  spin-­‐up  and  spin-­‐down  levels  is  quite  

different.  In  general  in  DFT  without  admixture  of  Hartree-­‐Fock  exchange  (as  is  the  

case  with  the  BP  functional  that  we  use  here),  the  orbital  energy  difference  between  

occupied  and  virtual  orbitals  is  a  0th  order  approximation  to  the  excitation  energy.  

In  the  present  case,  we  would  predict  a  ligand  field  transiton  around  2.8  eV  

(~23000  cm-­‐1).  Let’s  check  by  running  a  TD-­‐DFT  calculation.  We  calculate  just  six  

excited  states  corresponding  to  the  six  ways  of  taking  an  electron  out  of  a  t2g  orbital  

and  placing  it  into  the  eg  orbital.  In  order  to  do  this  calculation  include:  

 The  result  is:  ------------------------------------------------------------------------------ ABSORPTION SPECTRUM ------------------------------------------------------------------------------ State Energy Wavelength fosc T2 TX TY TZ

                                                                                                               14  However,  they  are  not  „pure“  dxy,dxz  and  dyz  orbitals.  This  is  because  degenerate  orbitals  can  arbitrarily  mix  among  themselves  without  changing  anything  in  the  physics.  These  three  orbitals  simply  span  the  three  components  of  the  t2g  representation.    

%tddft nroots 6 end

Page 22: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

(cm-1) (nm) (D**2) (D) (D) (D) ------------------------------------------------------------------------------ 1 20591.4 485.6 0.000000000 0.00000 0.00000 -0.00002 0.00000 2 20591.5 485.6 0.000000000 0.00000 -0.00000 0.00000 0.00002 3 20591.8 485.6 0.000000000 0.00000 0.00000 -0.00003 0.00001 4 23600.1 423.7 0.000000000 0.00000 0.00001 -0.00000 0.00001 5 23600.7 423.7 0.000000000 0.00000 0.00001 -0.00001 -0.00001 6 23601.3 423.7 0.000000000 0.00000 0.00000 0.00002 0.00001

 We  see  two  excited  states  that  are  triply  degenerate  each  and  correspond  to  the  

singly  excited  4T1g  and  4T2g  terms  expected  from  Table  1.15  Despite  the  much  too  

small  basis  set,  the  result  does  not  even  compare  poorly  with  experiment,  where  

transitions  are  observed  at  17,400  and  24,700  cm-­‐1.16  Furthermore  the  average  of  

the  two  transition  energies  correspond  well  with  the  orbital  energy  difference.  The  

splitting  between  the  two-­‐states  is  one  of  the  characterisitics  of  electron-­‐electron  

repulsion  leading  to  different  multiplets.  The  compositions  of  these  transition  shows  

that  they  indeed  correspond  to  the  expected  excitations:  STATE 1: E= 0.093822 au 2.553 eV 20591.4 cm**-1 39a -> 43a : 0.040541 (c= -0.20134677) 40a -> 42a : 0.235055 (c= -0.48482442) 40a -> 43a : 0.091500 (c= 0.30248889) 41a -> 42a : 0.033109 (c= 0.18195871) 41a -> 43a : 0.591178 (c= 0.76888093) STATE 2: E= 0.093822 au 2.553 eV 20591.5 cm**-1 39a -> 43a : 0.102372 (c= 0.31995577) 40a -> 42a : 0.340978 (c= 0.58393339) 40a -> 43a : 0.178600 (c= -0.42261116) 41a -> 42a : 0.028003 (c= 0.16734110) 41a -> 43a : 0.340662 (c= 0.58366233) STATE 3: E= 0.093823 au 2.553 eV 20591.8 cm**-1 39a -> 42a : 0.835399 (c= -0.91400145) 40a -> 42a : 0.010739 (c= -0.10363150) 40a -> 43a : 0.138077 (c= -0.37158743) STATE 4: E= 0.107530 au 2.926 eV 23600.1 cm**-1 41a -> 42a : 0.921691 (c= -0.96004745) 41a -> 43a : 0.057615 (c= 0.24003050) STATE 5: E= 0.107533 au 2.926 eV 23600.7 cm**-1

                                                                                                               15  Hint:  you  can  find  out  which  is  which  by  using  a  group  theoretical  table  and  the  output  for  the  CD  spectrum.  The  magnetic  dipole  components  RX,  RY,  RZ  transform  as  the  molecular  rotations.  Thus,  you  have  to  find  out  whether  a  4A2g→

4T1g  or  a  4A2g→

4T2g  transition  is  magnetic  dipole  allowed!  16  However,  this   is  not  always  the  case  for  TD-­‐DFT  calculations.  There   is  every  reason  to  be  critical  about  them  and  compare  very  careful  to  experiment!  In  many  cases  artifacts  arise.  Charge  transfer  states  tend  to  appear  MUCH  too  low  in  the  spectrum,  double  excitations  are  entirely  missing,  neutral-­‐to-­‐ionic  transitions  are  poorly  predicted,  spin-­‐flip  transitions   can   not   be   predicted   etc.   Some   of   this   is   discussed   in   Neese,   F.   (2006)   A   Critical   Evaluation   of   DFT,  including   Time-­‐Dependent   DFT,   Applied   to   Bioinorganic   Chemistry.   J.   Biol.   Inorg.   Chem.,   11,   702-­‐711.   Ab   initio  methods  are  covered  in  some  detail  in:  Neese  F.;  Petrenko,  T.;  Ganyushin,  D.;  Olbrich,  G.  (2006)  Advanced  Aspects  of  ab  initio  Theoretical  Spectroscopy  of  Open-­‐Shell  Transition  Metal  Ions.  Coord.  Chem.  Rev.,  205,  288-­‐327  

Page 23: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

39a -> 42a : 0.122088 (c= 0.34941155) 39a -> 43a : 0.011888 (c= 0.10903325) 40a -> 42a : 0.341659 (c= -0.58451589) 40a -> 43a : 0.505530 (c= -0.71100658) STATE 6: E= 0.107536 au 2.926 eV 23601.3 cm**-1 39a -> 42a : 0.029369 (c= -0.17137471) 39a -> 43a : 0.828386 (c= 0.91015695) 40a -> 42a : 0.062645 (c= -0.25029038) 40a -> 43a : 0.070858 (c= 0.26619100)

 

In  LFT,  the  energies  of  the  two  transitions  are  given  by:  

  4A2g→4T2g   :  ΔE=  Δ0  

  4A2g→4T1g   :  ΔE=  Δ0+12B  

Thus,  from  the  calculations  you  can  easily  determined  the  values  of  Δ0  and  B.  In  the  

free  Cr(III)  ion,  the  B-­‐value  is  known  to  be  950  cm-­‐1.  The  ration  of  

B(complex)/B(free  ion)  is  known  as  the  nephelauxetic  ratio.  Is  it  larger  or  smaller  

than  unity?  Why  is  it  larger  or  smaller  than  unity?    

JOB:  • Run   similar   calculations   for   other   d3   systems:   [V(H2O)6]2+   and   the  

hypothetical   [Ti(H2O)6]+   as  well   as   the  d8   system   [Ni(H2O)]2+  as  well   as   the  

hypothetical  [Cu(H2O)6]3+    

• Look  at  the  results  of  the  population  analysis.  How  does  the  partial  charge  on  

the  metal  change.  How  does  the  spin-­‐population  at  the  metal  change.  Look  at  

the  occupancies  of  the  metal  d-­‐orbitals.  What  does  that  say  about  σ-­‐  and  π-­‐

donation   of   the   water   ligands.   How   does   it   compare   with   the   formal  

occupations.   Compare   the   spin-­‐populations   of   the   d3   and   the   d8   systems.  

What  do  you  observe?    

• Calculate  the  first  six  ligand  field  excited  states  and  determine  Δ0  and  B  from  

the   computational   results.     Do   the   trends   make   sense   in   terms   of   your  

chemical  intuition?17  Calculate  the  nephelauxetic  ratios  –  what  do  you  include  

                                                                                                               17  The  Racah-­‐parameters  of  the  free  ions  are:  Ti2+=720  cm-­‐1,  Ti+=680  cm-­‐1,  V2+=765  cm-­‐1,  V3+=860  cm-­‐1,  Ni2+=1080  cm-­‐1,  Ni+=   1040   cm-­‐1,   Cu2+=   1240   cm-­‐1,   Cu+=1220   cm-­‐1,   Cu3+=1260   cm-­‐1.   In   the  d8   configuration,   the   first   two   ligand   field  excited   states   (3A2g→

3T2g,3T1g)   are   predicted   to   occur   at  Δ0   and  Δ0+12B   in   a   similar   way   as   found   for   the   the   d3  

configuration.   For   more   extensive   tabular   material   see   Haberditzl,   W.   Quantenchemie   4.   Komplexverbindungen,  Hüthig,  Heidelberg,  1979.    

Page 24: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

with  respect  to  the  ability  of  TD-­‐DFT  to  model  these  values  correctly.18  Will  

the  nephelauxetic  ratios  depend  on  the  nature  of  the  metal,  its  oxidation  state  

and  the  ligand?  Please  explain  your  answer.  

• Study   the   dependence   of   Δ0   (and   B)   on   the   metal-­‐ligand   distance   of  

[Cr(H2O)6]3+:  Does  the  result  match  your  expectation?  With  which  power  of  

the  metal-­‐ligand  distance  R  do  you  expect  Δ0  to  vary  based  on  conventional  

ligand  field  arguments?    

1.2.2 Inverted  Bonding19  In  many  cases,  you  will  observe  that  it  is  difficult  to  unambiguously  identify  the  

metal  based  MOs.  This  is  particularly  so  for  later  metals,  higher  oxidation  states  and  

increasing  numbers  of  unpaired  electrons.  The  reason  is  that  higher  oxidation  states  

tend  to  stabilize  the  d-­‐orbitals  that  consequently  move  closer  to  the  ligand  orbitals.  

In  addition,  large  number  of  singly  occupied  orbitals  that  are  mainly  centered  on  the  

metal  lead  to  a  very  large  exchange  stabilization  which  further  suppresses  the  

energies  of  the  occupied  spin-­‐up  orbitals.  As  a  consequence,  these  orbitals  move  

down  in  energy  into  the  ligand  based  orbitals.  This  effect  is  more  pronounuced  the  

more  Hartree-­‐Fock  exchange  you  have  in  your  density  functional  and  is  extremely  

strong  for  the  Hartree-­‐Fock  method  itself.  Yet,  the  ligand  field  picture  is  not  invalid  –  

it  is  simply  more  difficult  to  recover  it  from  the  MO  calculations.    

• One  useful  way   is   to   run   the   orbitals   through   a   localization  procedure   and  

look  at  the  localized  orbitals  of  mainly  metal  character.  

• Another  useful  way  to  analyze  the  calculations  is  to  look  at  the  set  of  orbitals  

that   we   call   “quasi-­‐restricted   orbitals”. 20  The   singly   occupied   ones   will  

correspond  well  with  your  ligand  field  expectations.    

JOB:  

                                                                                                               18  The  nephelauxetic  ration  are  found  to  be  0.79  for  [CrIII(H2O)6]

3+  and  0.77  for  [NiII(H2O)6]2+.    

19  A  nice  discussion  is  found  in:   I,  J.;  Noodleman,  L.;  Case,  D.A.   (2000)   In:  Solomon,  E.I.;  Lever,  A.B.P  (Eds.)   Inorganic  Electronic  Structure  and  Spectroscopy,  Vol.  1,  John  Wiley  &  sons,  New  York,  pp  661ff  20  For  their  definition  see:  Neese,  F.  (2006)   Importance  of  Direct  Spin-­‐Spin  Coupling  and  Spin-­‐Flip  Excitations  for  the  Zero-­‐Field  Splittings  of  Transition  Metal  Complexes:  A  Case  Study,  J.  Am.  Chem.  Soc.,  128,  10213-­‐10222  

Page 25: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

• Do  a  calculation  on   [Fe(H2O)6]3+.  Look  at   the  canonical  orbitals   in  both,   the  

spin-­‐up  and  the  spin-­‐down  sets.  Where  are  the  metal  orbitals,  what   is   their  

composition,  what  are  their  energies,  how  do  you  interpret  the  result?  

• Localize  the  orbitals  using  a  %loc end  block.  Visualize  the  orbitals.  Analyze  

their   composition   by   re-­‐reading   them   into   another   calculation   with   the   !

NoIter  Keyword.  

• Obtain   a   set   of   QRO’s   by   including   the   ! UNO   keyword.   Visualize   them,  

analyze  them.    

What  does  your  result  imply  in  terms  of  frontier  molecular  orbital  theory?  Where  do  

you  think  will  the  electron  go  upon  reduction  of  the  complex  to  the  ferrous  form?    

1.2.3 Covalencies  of  Metal  Ligand  Bonds  As  mentioned  in  the  introduction  to  this  chapter,  the  coalvency  of  the  metal  ligand  

bonds  is  an  important  factor  in  determining  the  physical  and  reactive  properties  of  

transition  metal  complexes.  We  wish  to  understand  covalency  here  as  the  ability  of  

the  metal  and  the  ligand  to  share  electrons.  A  quantitative  measure  for  this  is  how  

much  ligand  character  is  mixed  into  the  metal  d-­‐based  orbitals  and  vice  versa.21    

Let  us  look  at  [M(H2O)6(X)]n+  complexes  and  vary  X  for  fixed  M  and  then  M  for  fixed  

X.    

JOB:  

• Choose  X  to  be:  CO,  CN-­‐,  NH3,  OH2,  O2-­‐,  S2-­‐,  F-­‐,  Cl-­‐,  Br-­‐  and  I-­‐.    

• For   the   hypothetical   complexes   [CrIII(H2O)5(X)]3+/[CrIII(H2O)5(X-­‐

)]2+/[CrIII(H2O)5(X2-­‐)]2+   optimize   the   geometry   of   the   complexes   using   RI-­‐

BP86   and   the   small   basis   set   defined   above.   Calculate   the   energies   of   the  

system   using   the   B3LYP   functional   in   single   point   calculations   and   include  

the   COSMO(water)   solvation   model.   Look   at   the   metal-­‐t2g   and   metal-­‐eg  

                                                                                                               21  Please  recall  what  we  have  said  about  the  unitary  invariance  of  the  wavefunction  (or  Kohn-­‐Sham  determinant)  with  respect   to   rotations   among   occupied   orbitals.   Thus,   there   is   no   “absolute   truth”   in   such   numbers   and   they   are   of  course   not   observables;   yet,   it   is   highly   suggestive   and   chemically   sensible   to   study   the   variations   in  metal-­‐ligand  bonding  by  looking  at  the  metal-­‐ligand  bonding  and  antibonding  orbitals.  

Page 26: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

derived   orbitals.   Localize   the   occupied   orbitals   and   look   at   the   LMO’s.22  

Which   conclusions   can   you   draw   about   the   nature   of   the   bond?  Discuss  π-­‐

bonding   versus   π-­‐backbonding.   How   does   it   vary   with   the   ligand?   Which  

changes  do  you  observe  in  the  populations  of  the  metal  d-­‐orbitals?  

• From   the   total   energy   calculations   that   you  did   you   can   compare   the  bond  

energy  of  the  ligand  X  relative  to  the  hexaquo  complex.  To  this  end  you  have  

to  run  a  geometry  optimization  plus  energy  for  the  free  ligands  X.  How  does  

the  bond  energy  vary  with  the  ligand  and  the  oxidation  state  of  the  metal?23  

Is  it  in  accord  with  your  chemical  intuition?  

• Repeat   the   calculations   for   fixed   ligands   with   a   lower-­‐valent   metal,   for  

example,   the   low-­‐spin  configuration  (S=0)  of  Fe(II).  How  does   the  ability  of  

the  metal  for  backbonding  change?    

1.2.4 Organometallics  The  field  or  organometallic  chemistry  is  of  growing  importance  due  to  the  fact  that  

versatile  and  powerful  catalysts  can  be  designed  on  the  basis  of  such  complexes.  

Here,  we  take  a  brief  look  at  the  “father”  of  all  organometallic  compounds  –  Fe(cp)2.  

Below  you  find  a  set  of  reasonable  coordinates  for  this  complex:  

JOB:  

• Run  a  single  point  calculation  using  the  BP86  functional  and  the  small  basis  

set  used  throughout  this  computer  experiment.  

                                                                                                               22  Recall   that   the   localized  orbitals  never   show  a  bond  where   there   is  none.  Thus,   if   a  bonding  and  an  antibonding  partner   of   a   bond   are   both   fully   occupied,   the   localization   procedure   simply   separates   the   two   constituents   into  atomic  or  fragment  orbitals.  A  bond  only  remains  if  the  antibonding  partner  is  in  the  virtual  space.  Thus,  localization  and  orbital  inspection  is  a  good  way  to  discuss  the  net  bonding  effects.  You  can  visualize  these  localized  orbitals  with  the   orca_plot   program   and   you   can   analyze   their   composition   by   running   a   calculation  which   have   the  myjob.loc  orbitals  as  initial  guess  and  ! noiter.    23  Alternatively,  if  you  want  to  have  an  absolute  value  for  the  bond  energy,  optimize  the  geometry  of    [Cr(H2O)5]

3+/2+  and   calculate   its   energy   in   the   same  way;   run  a   geometry  optimization  of   the   free   ligand  and   calculate   its   energy.  Subtract  the  energies  of  reactants  and  products.  However,  in  order  to  get  more  reliable  values  you  also  have  to  taken  into  account  the  basis  set  superposition  error  (BSSE)  as  explained  in  chapter  Error!  Reference  source  not   found.  on  page  144).  The  basis  sets  we  use  here  are  so  small  that  these  undesired  effects  are  likely  to  be  large.  Therefore,  for  the  qualitative  purposes  of  this  section  it  is  probably  better  to  simply  compare  the  bond  energies  using  the  hexaquo  complex  as  a  reference.    

Page 27: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

• Analyze   the   electronic   structure   of   the   complex   using   orbitals   and  

populations.  What  is  the  nature  of  the  metal  ligand  bond?  

• Do   a   calculation   on   the   free   cyclopentadienyl   ligand   and   determine   the  

fragment   orbitals   that   contribute   most   to   the   bonding.   Draw   a   fragment  

orbital  correlation  diagram.  

• Use   the   BHLYP   functional   to   calculate   the   first   15   electronically   excited  

states.   What   is   their   nature?   Compare   the   transition   energies   with   d-­‐d  

transition   energies  of   the  hexaquo   complexes   that   you  have   studied   above.  

What  do  you  observe?  What  is  your  interpretation?  Determine  the  dominant  

donor   and   acceptor   orbital   pairs.   Work   out   the   symmetry   of   the   final  

electronic   states.   Determine   the   selection   rules   for   the   electric   dipole   and  

magnetic  dipole  transitions.  

Structure  of  Fe(cp)2  used  in  this  experiment  (charge=0,  multiplicity=1;  closed  shell)24  

 

                                                                                                               24  To  the  interested  student:  work  out  a  strategy  for  computing  the  bond  energy  of  the  Fe(II)-­‐cp-­‐  bond  and  think  about  a  suitable  Werner  type  Fe(II)  complex  with  which  you  can  compare  your  results.  A  challenge   is   that  you  separate  a  neutral  molecule  into  charged  fragments;  therefore  you  should  use  the  charge  compensation  offered  by  the  COSMO  model  and  you  may  have  to  do  it  via  a  ligand  exchange  reaction.    

Fe 0 0 0 0.000000 0.000 0.000 C 1 0 0 2.035741 0.000 0.000 C 2 1 0 1.399741 69.943 0.000 C 3 2 1 1.399822 108.010 59.725 C 4 3 2 1.399734 108.009 0.000 C 5 4 3 1.399880 107.981 359.931 H 1 2 3 2.949216 112.347 67.069 H 7 1 2 2.691289 62.854 207.853 H 8 7 1 2.691150 107.996 314.868 H 9 8 7 2.690913 108.001 0.000 H 10 9 8 2.690897 108.005 0.000 C 7 1 2 1.118895 28.137 142.474 C 8 7 1 1.118881 54.749 328.336 C 9 8 7 1.118864 54.753 13.469 C 10 9 8 1.118884 54.757 13.468 C 11 10 9 1.118850 54.762 13.464 H 2 1 3 1.100415 137.025 239.400 H 3 2 1 1.100553 125.205 225.905 H 4 3 2 1.100560 125.189 166.179 H 5 4 3 1.100431 125.231 166.185 H 6 5 4 1.119068 125.203 166.363

Page 28: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

1.2.5 Conformation.  A  Famous  Example  The  complex  ion  [CuCl4]2-­‐  is  known  to  exist  in  different  crystals  with  varying  

counterions.  While  Cu(II)  normally  prefers  a  square  planar  coordination,  this  ion  is  

found  in  various  forms  ranging  from  square-­‐planar  to  almost  tetrahedral.  Here  we  

calculate  the  energy  associated  with  this  distortion.  You  can  use  the  following  input  

file:  

 This  input  will  lead  the  ORCA  program  to  perform  a  “rigid  scan”  that  will  lead  from  a  

square  planar  to  a  tetrahedral  structure  by  defining  a  distortion  angle  θ  (the  final  

angle  should  (180-­‐109.4712)/2;  you  can  also  use  a  larger  number  of  steps  in  order  

to  obtain  a  more  smooth  looking  surface).    

JOB  • Do   the   calculation,   plot   the   resulting   potential   energy   surface,   convert   the  

energies  to  relative  energies  in  kcal/mol.  

• Where  do  you  find  the  minimum?  

• Does  it  require  a  lot  of  energy  to  distort  the  molecule?    

• What   is   the   change   in   electronic   structure?   Examine   two   or   three   points  along  the  surface.  Look  at  the  orbitals  and  populations;  calculate  the  first  four  ligand   field   excited   states.   Do   you   think   that   spectroscopy  will   be   a   useful  indicator  of  the  coordination  geometry?25    

 The  latter  calculation  represents  one  of  the  cases  where  TD-­‐DFT  with  normal  GGA  

functionals  like  BP86  fails  badly  and  predicts  LMCT  states  where  there  should  be  d-­‐                                                                                                                25  A  very  nice  analysis  of  this  problem  can  be  found  in:  Solomon,  E.I.  Comments  Inorg.  Chem.  (1984),  3,  227  

! UKS BP RI SV DeMon/J SlowConv NoFinalGrid %scf shift shift 0.3 erroff 0 end end # geometric parameters %paras R_ML = 2.25 # the metal-water-oxygen distance Theta = 0,35,17 end * xyz -2 2 cu 0 0 0 cl { R_ML*cos(Theta) } 0 { R_ML*sin(Theta) } cl {-R_ML*cos(Theta) } 0 { R_ML*sin(Theta) } cl 0 { R_ML*cos(Theta) } { -R_ML*sin(Theta) } cl 0 {-R_ML*cos(Theta) } { -R_ML*sin(Theta) } *

Page 29: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

d  excitations  (in  the  region  below  ~15000  cm-­‐1).  You  should  therefore  do  the  latter  

calculation  with  a  functional  that  includes  more  HF  exchange  (BHLYP  is  

recommended  which  introduces  as  much  as  50%  HF  exchange).  You  may  also  need  

to  calculate  more  than  the  desired  four  ligand  field  states  in  order  to  “catch”  the  

states  of  interest  since  they  do  not  come  in  order  of  increasing  orbital  energy  

difference.  With  25  states  you  are  on  the  safe  side  –  please  observe  the  assignment  

of  the  transitions  and  see  where  the  occupied  metal  orbitals  are  in  the  spin  down  set  

(draw  a  MO  diagram).  

1.2.6 Reactivity:  Powerful  Oxidants.  The  interesting  species  [FeIVO(H2O)5]2+  has  can  be  made26  and  will  be  studied  in  this  

computer  experiment.  It  is  a  simple  model  for  related  FeIV-­‐oxo  species  that  are  of  

much  relevance  to  chemistry  and  biochemistry  since  they  are  powerful  oxidants.27  

Here  we  will  study  some  aspects  of  structure,  bonding  and  energetics  of  such  

species.  First,  of  all  –  which  spin  state  is  the  lowest  one  for  this  system  –  S  =  1  or  S  =  

2.  Try  to  find  out  in  a  computer  experiment  and  optimize  the  structure  of  the  

species:  

                                                                                                               26  O.  Pestovsky,  S.  Stoian,  E.L.  Bominaar,  X.P.  Shan,  E.  Münck,  L.  Que  Jr.,  A.  Bakac,  Angew.  Chem.   Int.  Ed.  44   (2005)  6871  27  The  most  intensely  studied  system  is  Cytochrome  P450  which  features  a  very  complicated  electronic  structure.  The  active  species  (compound  I)  involves  an  iron-­‐oxo  group  coupled  to  a  porphyrin  radical  .  For  recent  theoretical  studies  see:  Shaik,  S.;  de  Visser,  S.  P.;  Ogliaro,  F.;  Schwarz,  H.;  Schröder,  D.  Curr.  Opin.  Chem.  Biol.  (2002),  6,  556;  S.  Shaik,  D.  Kumar,  S.P.  de  Visser,  A.  Altun,  W.  Thiel,  Chem.  Rev.  105  (2005)  2279;  Schöneboom,  J.;  Neese,  F.;  Thiel,  W.  (2005)  J.  Am.  Chem.  Soc.,  127,  5840-­‐5853;    

Page 30: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

   

• Repeat   the   optimization   for   the   low-­‐spin   state.   Analyze   the   electronic  

structure   of   both   species.  What   are   the   electron   configurations   in   the   two  

states?  How  did  the  structure  change?  How  did  the  changes  in  the  structure  

reflect   the   different   electronic   configurations?28  Is   the   oxo-­‐group   strongly  

affected  by  the  change  in  spin  state?    

• Recalculate   the   energies   at   the   optimized   geometries   using   the   B3LYP  

functional.29  What   is   the   energy  difference  between   the   two   states?30  Could  

either  be  the  ground  state?    

                                                                                                               28  More  interestingly  for  the  synthetically  oriented  chemists:  How  do  have  to  choose  your  ligand  in  order  to  force  the  system  to  adopt  either  one  or  the  other  ground  state?  29  In   “real   life”   you   would   probably   do   it   similarly   but   with   much   larger   basis   sets;   for   example,   the   geometry  optimization   with   TZVP   and   the   final   energy   calculation   with   TZVPP.   We   do   not   do   that   in   order   to   keep   the  computation   times   short.   In   a   serious   study   you   would   also   want   to   evaluate   the   zero-­‐point   energy   and   thermal  contributions  to  the  energy  differences  just  as  done  in  chapter  Error!  Reference  source  not  found.  on  page  51.  30  Compare  for  example:  Neese,  F.  (2006)  J.  Inorg.  Biochem.  100,  716-­‐726.  The  entire  issue  deals  with  Fe(IV)  and  you  will  find  many  other  interesting  articles  there  dealing  with  the  structure,  chemistry,  spectroscopy  and  theory  of  such  compounds.  

! UKS BP RI SlowConv NoFinalGrid TightSCF Opt %scf shift shift 0.3 erroff 0 end end %basis basis sv newgto h "STO-3G" end end %paras R_ML = 1.995 R_OH = 0.9817 A_HOM = 126.153 R_MX = 1.64 end * int 2 5 Fe 0 0 0 0.0 0.0 0.0 O 1 0 0 {R_ML} 0.0 0.0 O 1 2 0 {R_ML} 180.0 0.0 O 1 2 3 {R_ML} 90.0 0.0 O 1 2 4 {R_ML} 90.0 180.0 O 1 2 4 {R_ML} 90.0 270.0 O 1 2 4 {R_MX} 90.0 90.0 H 2 1 3 {R_OH} {A_HOM} 90.0 H 2 1 3 {R_OH} {A_HOM} 270.0 H 3 1 2 {R_OH} {A_HOM} 90.0 H 3 1 2 {R_OH} {A_HOM} 270.0 H 4 1 2 {R_OH} {A_HOM} 0.0 H 4 1 2 {R_OH} {A_HOM} 180.0 H 5 1 2 {R_OH} {A_HOM} 180.0 H 5 1 2 {R_OH} {A_HOM} 0.0 H 6 1 2 {R_OH} {A_HOM} 90.0 H 6 1 2 {R_OH} {A_HOM} 270.0 *

Page 31: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

Now  let  us  study  the  hydrogen  atom  abstraction  reaction:  

  [FeO(H2O)5]2+    +  CH4  →    [Fe(OH)(H2O)5]2+  +  CH3•  

JOB:  

• Optimize   the   structures   of   the   reactants   and   products   (assume   S=5/231)  

using  the  RI-­‐BP86  method  as  above.    

• Recalculate  the  energies  using  B3LYP  single  point  calculations.  

• Is  the  reaction  predicted  to  be  feasible?  If  not,  would  this  molecule  be  able  to  

oxidize  weaker  C-­‐H  bonds  such  as  that  of  toluene?  

• Analyze  the  structure  of  the  reaction  product.  What  are  the  major  geometric  

changes?  What  are  the  electronic  changes  brought  about  by  protonating  the  

oxo  group  and  reducing  the  metal?  What  happens  to  the  Fe-­‐O  bond,  what  to  

its  covalency?    

• Compare   the   capability   of   this   iron-­‐oxo   species   to   two   other   systems:   OH•  

radical   and   the   well-­‐known   species   [VO(H2O)5]2+.   The   latter   calculation  

requires  another  set  of  geometry  optimizations.  

1.2.7 Unusual  Bonds  and  Coordinated  Radicals  –  Iron-­‐Nitrosyls  Questions  of  structure  and  bonding  can  become  quite  complicated  when  the  

oxidation  state  of  the  ligand  becomes  ambiguous.  A  famous  example  are  transition                                                                                                                  31  Start   from   the   following   coordinates:    Fe 0 0 0 0.000000 0.000 0.000 O 1 0 0 2.064887 0.000 0.000

O 1 2 0 2.063420 175.071 0.000

O 1 2 3 2.105932 89.618 85.198

O 1 2 4 2.106399 89.557 195.176

O 1 2 5 2.100535 87.401 82.433

O 1 2 6 1.768057 92.691 179.818

H 2 1 7 1.009685 122.106 359.695

H 2 1 7 1.008473 127.717 179.862

H 3 1 7 1.009525 121.926 0.061

H 3 1 7 1.008441 127.868 180.210

H 4 1 2 1.008208 123.239 15.085

H 4 1 2 1.008169 123.475 169.580

H 5 1 2 1.008168 123.401 190.709

H 5 1 2 1.008201 123.457 345.633

H 6 1 2 1.007542 124.625 90.015

H 6 1 2 1.007546 124.702 269.828

H 7 1 2 1.005566 179.700 45.575  

Page 32: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

metal  nitrosyls.  Here  the  NO  ligand  is  believed  to  be  able  to  coordinate  as  NO-­‐(S=1),  

NO-­‐(S=0),  NO•(S=1/2)  and  NO+(S=0)  depending  on  the  metal,  its  formal  oxidation  

state  and  the  remaining  ligand  set.  Since  the  interpretation  is  often  ambiguous,  

Enemark  and  Feltham  have  invented  the  {MNO}x  notation,  where  x  is  the  number  of  

metal-­‐d  plus  NO-­‐π*  electrons.  Thus,  the  reaction  of  Fe(II)  with  NO•  yields  a  {FeNO}7  

complex.  We  will  study  here  the  iron  pentaaquo-­‐NO  complex  –  a  molecule  which  is  

surely  well  known  to  every  chemist32  but  has  a  quite  complicated  electronic  

structure.33  

• Optimize  the  geometry  of  [Fe(H2O)5(NO)]3+/2+/1+  in  the  spin  states  S=5/2,3/2  

and  1  (you  can   initially  assume  an  NO  bond  distance  of  1.76  Angström  and  

Fe-­‐O  bonds  of  2.12;  an  N-­‐O  distance  of  1.2  Angström  and  an  O-­‐N-­‐Fe  angle  of  

170°).   The   standard   input ! UKS BP RI SlowConv NoFinalGrid Opt

TightSCF  should  work.  

• How  does  the  N-­‐O  distance  change  with  oxidation  state?  How  does  the  Fe-­‐N  

distance  change?  What  does  it  tell  you  about  the  oxidation  state  of  the  ligand?  

(Compare  to  free  NO+,  NO•  and  NO-­‐)  Do  you  observe  trans  effects?  

• Analyze   the   bonding   using   orbitals,   spin-­‐   and   electron-­‐populations,   bond  

orders   and   all   the   tricks   that   you   have   learned. 34  What   is   your   best  

description  of  the  electronic  structure  in  terms  of  dN  configurations?  In  fact,  

is  the  dN  designation  sensible  in  this  case?    

• Are  there  alternative  spin  states?  May  they  be  lower  in  energy  than  the  ones  

that  you  have  calculated?    

                                                                                                               32  It  is  the  “brown  ring”  complex  in  analytic  chemistry.  33  Wanat,  T.;  Schneppensieper,  G.;  Stocher,  G.;  van  Eldik,  R.;  Bill,  E.;  Wieghardt,  K.  Inorg.  Chem.  (2002),  41,  4;  a  very  interesting  combined  experimental  and  theoretical  study  on  a  related  system  is  found  in:  Brown,  C.  A.;  Pavlosky,  M.  A.;  Westre,  T.  E.;  Zhang,  Y.;  Hedman,  B.;  Hodgson,  K.  O.;  Solomon,  E.  I.  J.  Am.  Chem.  Soc.  1995,  117,  715.  34  In  fact,  there  is  an  additional  trick  that  you  can  learn  here  –  unrestricted  corresponding  orbitals  (UCOs).  In  order  to  get   them  use  ! UCO   in   the  ORCA   input   file.   This  will   produce   a   file   called  MyJob.uco   and   additional   output.   The  additional  output  will  consist  of  spatial  overlaps  of  spin-­‐up  and  spin-­‐down  orbitals.  Overlaps  close  to  zero  or  at  least  significantly   smaller   than   one   indicate   a   spin   coupled   system.   In   this   case   the   spin-­‐expectation   value   <S2>   will  significantly  deviate  from  S(S+1)  [Look  in  the  output  for  this  number].  What  is  your  interpretation  if  you  find  such  a  spin  coupled  system  here?  Look  at  the  UCO’s  and  observe  which  orbitals  are  spin  coupled  (if  any).  (For  a  discussion  see  Neese,  F.  J.  Phys.  Chem.  Solids,  2004,  65,  781-­‐785)  

Page 33: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

1.2.8 High-­‐Spin/Low-­‐Spin  and  Spin-­‐Crossover  An  important  field  of  investigation  deals  with  transition  metal  complexes  that  have  

two  thermally  accessible  spin  states  –  such  complexes  are  spin-­‐crossover  systems;  

in  the  LFT  language  they  feature  ligand  field  strengths  that  are  just  on  the  boarder  of  

the  strong-­‐field  and  the  weak  field  part  of  the  Tanabe-­‐Sugano  diagram.  This  case  is  

particularly  frequently  met  in  Fe(II)  complexes,  where  the  5T2g  (t2g)4(eg)2  and  1A1g  

(t2g)6(eg)0  states  are  part  of  the  spin-­‐crossover  system.  A  very  accurate  calculation  of  

the  energy  difference  between  these  two  states  is  obviously  very  complicated  as  one  

needs  to  reach  an  accuracy  that  is  on  the  order  of  the  thermal  energy  (~200  cm-­‐1).  

Moreover,  the  spin-­‐crossover  is  frequently  driven  by  cooperative  and  entropic  

effects  which  are  particularly  difficult  to  model  with  quantum  chemical  methods.  

Nevertheless,  it  has  been  found  to  be  an  especially  difficult  problem  for  DFT  

methods  to  come  at  predictions  that  are  at  least  within  a  reasonable  range  and  

predictions  of  different  functionals  were  found  to  differ  by  very  large  margins.35  

Here  we  simply  study  a  very  simple  model  for  one  frequently  studied  spin-­‐crossover  

complex  [Fe(phen)2(NCS)2].  We  model  this  complex  by  [Fe(NH3)4(NCS)2]  in  order  to  

save  computation  time.  The  optimized  structure  for  the  low-­‐spin  and  the  high-­‐spin  

state  are  shown  below.  

JOB:  

• First   look   at   the   two   structures   given   below.   What   is   the   main   difference  

between   the   high-­‐spin   and   low-­‐spin   structure?   How   do   the   geometric  

structures  reflect  the  electronic  structures.  

• Run   single   point   calculations   on   both   structures   using   the   UHF  method   as  

well   as   the  BHLYP  B3LYP,   BLYP,   BP96   and   LSD   functionals.   Compare   your  

results  using  the  standard  small  basis  set  of  this  experiment.  To  which  extent  

do  you  want  to  rely  on  DFT  methods  in  order  to  predict  spin  state  energetics?  

Is  the  Hartree-­‐Fock  method  an  attractive  alternative?    

                                                                                                               35  See  Reiher,  M.;  Salomon,  O.;  Hess,  B.A.  Theor.  Chem.  Acc.  (2001),  107,  48-­‐55;  Paulsen,  H.;  Duelund,  L.;  Winkler,  H.;  Toftlund,   H.;   Trautwein,   A.X.   Inorg.   Chem.,   (2001),   40,   2201-­‐2203;   Zhang,   Y.;   Oldfield   J.   Phys.   Chem.,   (2003),   107,  4147-­‐4150;  Fouqueau,  A.;  Mer,  S.;  Casida,  M.E.;  Daku,  L.M.L.;  Hauser,  A.;  Mieva,  T.;  Neese,  F.  J.  Chem.  Phys.,  (2004)  120,  9473-­‐9486  

Page 34: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

 

 The  Low-­‐Spin  Structure:  

The  High-­‐Spin  Structure:36  

                                                                                                               36  SCF   convergence   may   be   somewhat   challenging   for   this   species   since   the   (“t2g”)

4(eg)2   configuration   is   almost  

orbitally  degenerate  and   the  orbital   energy  difference  between   the   spin-­‐down  HOMO  and   the   spin-­‐down  LUMO   is  only  ~0.2  eV!  The  calculations  with  the  hybrid  functionals  are  more  likely  to  converge  better  and  might  be  taken  as  input  for  the  calculations  that  do  not  involve  HF  exchange.  Try  ! SlowConv  and  perhaps  NRSCF  Consult  the  ORCA  manual  for  more  details.  

Fe 0 0 0 0.000000 0.000 0.000 N 1 0 0 2.049885 0.000 0.000 N 1 2 0 2.052645 171.400 0.000 N 1 2 3 2.075900 92.686 228.011 N 1 2 4 1.891473 87.079 183.611 N 1 2 5 2.076027 92.843 83.846 N 1 2 5 1.890882 87.097 260.185 C 7 1 2 1.216067 163.994 272.872 C 5 1 2 1.215828 165.562 59.575 S 8 7 1 1.670676 179.710 208.661 S 9 5 1 1.669960 179.595 121.297 H 2 1 7 1.039864 102.694 252.981 H 2 1 7 1.039853 101.663 5.151 H 2 1 7 1.036898 122.329 128.624 H 3 1 7 1.039350 102.218 352.185 H 3 1 7 1.039385 101.030 104.131 H 3 1 7 1.036609 122.740 227.615 H 4 1 2 1.043434 99.578 85.432 H 4 1 2 1.036103 114.396 201.937 H 4 1 2 1.036150 114.727 328.610 H 6 1 2 1.036117 114.516 29.302 H 6 1 2 1.036059 114.770 156.207 H 6 1 2 1.043065 99.101 272.880

Page 35: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

1.2.9 Exchange  Coupling  Between  Transition  Metal  Ions  A  final  important  subject  to  be  mentioned  in  this  course  is  the  “magnetic”  coupling  

between  open-­‐shell  transition  metal  ions.  There  is  a  large  class  of  multinuclear  

transition  metal  systems  with  constituent  ions  that  –  based  on  their  electron  count  –  

must  have  unpaired  electrons.  These  unpaired  electron  on  different  sites  typically  

couple  weakly  to  produce  a  ladder  of  spin  states.  This  ladder  may  

phenomenologically  be  described  by  a  so-­‐called  “Heisenberg  Hamiltonian”  that  has  

the  form  -­‐2J*SA*SB  where  SA  and  SB  are  the  spin-­‐operators  for  sites  ‘A’  and  ‘B’  and  J  is  

the  “exchange  coupling  constant”.  If  it  is  negative,  low-­‐spin  states  are  preferred  and  

the  system  is  said  to  be  “antiferromagnetically”  coupled  Positive  J  refers  to  

“ferromagnetic  coupling”.  The  ordering  of  spin-­‐states  has  important  implications  for  

many  branches  of  chemistry  and  physics.  We  will  not  enter  a  detailed  discussion  of  

this  important  subject  here.37,38    

                                                                                                               37  See  for  example  L.  Noodleman  J.  Chem.  Phys.  74  (1981)  5737;  L.  Noodleman,  E.R.  Davidson  Chem.  Phys.  109  (1986)  131;  A.  Bencini,  D.  Gatteschi  J.  Am.  Chem.  Soc.  108  (1980)  5736;  K.  Yamaguchi,  Y.  Takahara,  T.  Fueno,  in:  V.H.  Smith  (Ed.),  Applied  Quantum  Chemistry,  155;  O.  Kahn,  'Molecular  Magnetism',  VCH  Publishers,  New  York,  1993;  Calzado,  C.  J.;  Cabrero,  J.;  Malrieu,  J.  P.;  Caballol,  R.  J.  Chem.  Phys.  (2002),  116,  2728  38  It  is  important  to  understand  that  this  “magnetic  coupling”  has  nothing  to  do  with  genuine  magnetic  interactions.  The  origin  of  the  exchange  coupling  is  purely  electrostatic  in  origin  and  is  intimately  associated  with  the  antisymmetry  

Fe 0 0 0 0.000000 0.000 0.000 N 1 0 0 2.247312 0.000 0.000 N 1 2 0 2.248019 169.744 0.000 N 1 2 3 2.307572 93.715 223.307 N 1 2 4 2.037538 86.562 188.204 N 1 2 5 2.312841 93.535 80.789 N 1 2 5 2.040200 87.326 252.683 C 7 1 2 1.217279 165.572 266.841 C 5 1 2 1.217076 166.485 63.735 S 8 7 1 1.664217 179.384 344.402 S 9 5 1 1.664331 179.291 29.396 H 2 1 5 1.036127 101.817 352.840 H 2 1 5 1.037918 97.395 104.836 H 2 1 5 1.033648 124.170 226.576 H 3 1 5 1.036474 101.009 243.118 H 3 1 5 1.037390 97.718 355.185 H 3 1 5 1.033489 124.285 117.374 H 4 1 2 1.037219 96.213 87.978 H 4 1 2 1.032760 114.165 204.515 H 4 1 2 1.033161 113.380 331.970 H 6 1 2 1.033002 113.409 27.309 H 6 1 2 1.032651 114.389 155.169 H 6 1 2 1.037071 95.577 271.582

Page 36: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

The  accurate  calculation  of  exchange  coupling  constants  is  quite  challenging.  In  

principle,  multideterminantal  methods  are  required  to  give  a  proper  description  of  

the  low-­‐spin  states.  The  values  of  J  are  on  the  order  of  a  few  dozen  to  a  few  hundred  

wavenumbers  –  thus  one  has  to  calculate  small  differences  between  large  numbers  

to  high  accuracy  which  is  quite  difficult.  The  multideterminantal  (but  not  

multiconfigurational)  nature  of  the  low-­‐spin  states  would  seem  to  invalidate  DFT  

methods,  which  are  based  on  a  single  determinant.  However,  the  situation  is  not  so  

bad,  if  some  precaution  is  taken.  Consider  for  example  the  case  of  two  interacting  

Cu(II)  ions.  Since  both  are  d9  systems,  a  singlet  and  a  triplet  arise.  The  triplet  is  

straightforward  to  calculate  with  standard  DFT  methods,  but  the  singlet  is  not.  In  

fact,  the  best  way  to  approach  the  problem  in  DFT  that  is  known  up  to  date  is  to  

generate  a  so-­‐called  “broken-­‐symmetry”  solution  in  which  one-­‐electron  localizes  

with  spin-­‐up  on  one  site  and  another  one  with  spin-­‐down  on  the  second  site.  This  

does  not  represent  a  correct  singlet  wavefunction  but  merely  a  determinant  with  

MS=0.  Nevertheless,  the  physics  of  weakly  interacting  electrons  is  basically  correctly  

described.  Together  with  some  approximate  formalisms,  it  is  possible  to  extract  a  

value  for  J  from  a  high-­‐spin  and  a  broken  symmetry  calculation.  In  ORCA,  it  is  

possible  to  specifically  ask  for  such  solutions.    

We  study  the  famous  case  of  bis-­‐µ-­‐hydroxy-­‐bridged  Cu(II)-­‐dimers  which  represents  

a  classical  case  of  magnetostructural  correlations.  It  has  been  shown  experimentally,  

that  a  switch  between  ferromagnetic  and  antiferromagnetic  coupling  occurs  as  a  

function  of  bridging  angle  α.  The  critical  angle  is  found  somewhere  around  97°.  You  

can  do  a  calculation  and  see  whether  you  can  qualitatively  reproduce  this.    

JOB:    • Run  the  job  shown  below.  Vary  the  angle  Alpha  and  search  in  the  output  for  

the  predicted  exchange  coupling  constant  (choose  J(3)  for  plotting).  Plot  the  exchange   coupling   versus   angle   and   observe   whether   you   can   detect   an  change  from  J<0  to  J>0.  

                                                                                                                                                                                                                                                                                                                                         of   the   N-­‐particle   wavefunction.   There   is   no   “exchange   interaction”   in   nature.   What   we   describe   as   “exchange  interaction”  is  merely  the  combined  action  of  antisymmetry  and  electron-­‐electron  repulsion.  

Page 37: 1 Computer Experiment* 15: Computational Coordination ...scienide2.uwaterloo.ca/~nooijen/Chem-440-computational/ORCA_labs... · oriented$modifications$are$known$collectively$as$“ligand$field$theory”$but$we$will$

       

! UKS BP RI SV DeMon/J SlowConv TightSCF %basis newgto h "STO-3G" end end %scf BrokenSym 1,1 end %paras Alpha = 95.0 ATilt = 60.0 AHNCu = 112.0 ANCuN = 94.7 RCuO = 1.970 ROH = 1.034 RNH = 1.066 RCuN = 2.060 end * int 2 3 Cu 0 0 0 0 0 0 Cu 1 0 0 {2*RCuO*sin(0.5*Alpha)} 0 0 O 1 2 0 {RCuO} {90-0.5*Alpha} 0 O 1 2 3 {RCuO} {90-0.5*Alpha} 180 H 3 4 1 {ROH} {180-ATilt} 90 H 4 3 1 {ROH} {180-ATilt} 90 N 1 2 3 {RCuN} {180-ANCuN/2} 0 N 1 2 3 {RCuN} {180-ANCuN/2} 180 N 2 1 3 {RCuN} {180-ANCuN/2} 0 N 2 1 3 {RCuN} {180-ANCuN/2} 180 H 7 1 2 {RNH} {AHNCu} 0 H 7 1 2 {RNH} {AHNCu} 120 H 7 1 2 {RNH} {AHNCu} 240 H 8 1 2 {RNH} {AHNCu} 0 H 8 1 2 {RNH} {AHNCu} 120 H 8 1 2 {RNH} {AHNCu} 240 H 9 2 1 {RNH} {AHNCu} 0 H 9 2 1 {RNH} {AHNCu} 120 H 9 2 1 {RNH} {AHNCu} 240 H 10 2 1 {RNH} {AHNCu} 0 H 10 2 1 {RNH} {AHNCu} 120 H 10 2 1 {RNH} {AHNCu} 240 *