41
Computational UV/vis, IR and Raman Spectroscopy 1 1 Computer Experiment 8: Computational UV/vis, IR and Raman Spectroscopy 1.1 Introduction to Theoretical Spectroscopy Whenever it comes to the discussion of spectroscopy, it is essential to orient oneself about the molecular phenomena that are being studied (Figure 1). The natural order parameter for spectroscopy is the energy of the photons that are applied to the system. For high energy gammaray photons (>10 4 eV) one studies nuclear processes (for example in Mössbauer spectroscopy transition between different states of a 57 Fe nucleus are probed). Radiation in the hard and soft xray region (~10 4 10 2 eV) induces electronic transitions from corelevels to empty valence levels or into the nonbound continuum. Ultraviolet and visible (UV/vis) photons (14 eV) induce electronic excitations from filled to empty valence (and perhaps Rydberg) levels. It is this energy region where the photon energy is of the same order of magnitude than the energy of chemical bonds. Thus, electronic spectroscopy directly probes chemical bonding! Below the energy of visible photons, it is (usually) no longer possible to induce electronic transitions. Thus, infrared photons (0.010.5 eV ~1004000 cm 1 ) merely induce transitions between different vibrational levels of the molecule within a given electronic configuration. 1 At even lower energy (10 4 105 eV; ~110 cm 1 ) there occur the phenomena that are associated with the electron spin and that are probed by ESR (=EPR for our purposes) spectroscopy. Finally, with radiowave photons (10 6 10 7 ~0.0010.01 cm 1 ) one is only able to induce transitions between different states of magnetic nuclei and these are probed in NMR (and ENDOR) spectroscopy. In this computer experiment we study the UV/vis and infrared region of the spectrum and calculate the energy levels that are associated in electronic and vibrational transitions. In the next experiment, the lowenergy region covered by magnetic resonance spectroscopy is covered. In interpreting the results of the computations, we will use the language of molecular orbitals. It is, however, very important to understand that in actual experiments one never observes molecular orbitals. Every measurement always only reports energies and properties of many electron states. It is crucial to properly distinguish between the oneelectron (MO) and manyelectron (states) level in 1 An electronic configuration is defined by specifying the integer occupation number for each MO in the system such that the sum of occupation numbers equals the total number of electrons and each MO is occupied by either 0, 1 or 2 electrons.

1 Computer Experiment* 8:* Computational* UV/vis,* IR and ...scienide2.uwaterloo.ca/.../Orca_labs_7_UVVIS_IR_Raman.pdf · Computational+UV/vis,+IRandRamanSpectroscopy+ 1+ 1 Computer

  • Upload
    buinhu

  • View
    236

  • Download
    1

Embed Size (px)

Citation preview

Computational  UV/vis,  IR  and  Raman  Spectroscopy   1  

1 Computer   Experiment   8:   Computational   UV/vis,   IR   and   Raman  Spectroscopy  

1.1  Introduction  to  Theoretical  Spectroscopy  Whenever  it  comes  to  the  discussion  of  spectroscopy,  it  is  essential  to  orient  oneself  

about  the  molecular  phenomena  that  are  being  studied  (Figure  1).  The  natural  order  

parameter  for  spectroscopy  is  the  energy  of  the  photons  that  are  applied  to  the  system.  

For  high  energy  gamma-­‐ray  photons  (>104  eV)  one  studies  nuclear  processes  (for  

example  in  Mössbauer  spectroscopy  transition  between  different  states  of  a  57Fe  nucleus  

are  probed).  Radiation  in  the  hard-­‐  and  soft  x-­‐ray  region  (~104-­‐102  eV)  induces  

electronic  transitions  from  core-­‐levels  to  empty  valence  levels  or  into  the  non-­‐bound  

continuum.  Ultraviolet  and  visible  (UV/vis)  photons  (1-­‐4  eV)  induce  electronic  

excitations  from  filled  to  empty  valence  (and  perhaps  Rydberg)  levels.  It  is  this  energy  

region  where  the  photon  energy  is  of  the  same  order  of  magnitude  than  the  energy  of  

chemical  bonds.  Thus,  electronic  spectroscopy  directly  probes  chemical  bonding!  Below  

the  energy  of  visible  photons,  it  is  (usually)  no  longer  possible  to  induce  electronic  

transitions.  Thus,  infrared  photons  (0.01-­‐0.5  eV  ~100-­‐4000  cm-­‐1)  merely  induce  

transitions  between  different  vibrational  levels  of  the  molecule  within  a  given  electronic  

configuration.1  At  even  lower  energy  (10-­‐4-­‐10-­‐5  eV;  ~1-­‐10  cm-­‐1)  there  occur  the  

phenomena  that  are  associated  with  the  electron  spin  and  that  are  probed  by  ESR  (=EPR  

for  our  purposes)  spectroscopy.  Finally,  with  radiowave  photons  (10-­‐6-­‐10-­‐7~0.001-­‐0.01  

cm-­‐1)  one  is  only  able  to  induce  transitions  between  different  states  of  magnetic  nuclei  

and  these  are  probed  in  NMR  (and  ENDOR)  spectroscopy.    

In  this  computer  experiment  we  study  the  UV/vis  and  infrared  region  of  the  spectrum  

and  calculate  the  energy  levels  that  are  associated  in  electronic  and  vibrational  

transitions.  In  the  next  experiment,  the  low-­‐energy  region  covered  by  magnetic  

resonance  spectroscopy  is  covered.  In  interpreting  the  results  of  the  computations,  we  

will  use  the  language  of  molecular  orbitals.  It  is,  however,  very  important  to  understand  

that  in  actual  experiments  one  never  observes  molecular  orbitals.  Every  measurement  

always  only  reports  energies  and  properties  of  many  electron  states.  It  is  crucial  to  

properly  distinguish  between  the  one-­‐electron  (MO)  and  many-­‐electron  (states)  level  in  

                                                                                                               1 An electronic configuration is defined by specifying the integer occupation number for each MO in the system such that the sum of occupation numbers equals the total number of electrons and each MO is occupied by either 0, 1 or 2 electrons.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   2  

any  discussion  of  electronic  structure  in  order  to  avoid  unnecessary  mistakes  and  

confusion!  

 Figure  1:  The  energy  scale  of  molecular  spectroscopy.  

1.2  Background  

1.2.1 UV/vis  spectroscopy  UV/Vis  spectroscopy  is  a  routinely  used  experimental  technique  throughout  all  branches  

of  chemistry.  As  explained  above,  bands  in  UV/vis  spectra  correspond  to  the  excitation  

of  a  molecule  from  the  electronic  ground  state  to  electronically  excited  states  with  

simultaneous  excitations  of  vibrational,  rotational,…  quanta.  Thus,  electronic  transitions  

lead  to  quite  large  changes  in  the  electronic  structure  of  the  investigated  molecules.  

Since  the  nature  of  the  bonds  that  are  involved  in  the  transitions  change  between  the  

ground-­‐  and  the  excited  states,  molecules  in  excited  states  generally  assume  a  geometry  

that  is  different  from  the  ground  state  equilibrium  geometry.  In  fact,  these  large  changes  

in  bonding  are  the  basis  for  the  large  and  important  field  of  photochemistry  where  the  

making  and  breaking  of  bonds  in  electronically  excited  states  is  of  central  interest.  

1.2.1.1 The  Experiment  An  experimental  UV/vis  absorption  spectrum  consists  of  a  plot  of  the  molar  decadic  

extinction  coefficient   !  versus  the  excitation  energy.2  The  extinction  coefficient  ε(λ)  is  

the  characterisitic  molecular  property  that  we  are  going  to  calculate  in  this  computer                                                                                                                  2 More commonly, the wavelength of the radiation is plotted on the x-axis. The relation between wavelength (in nm) and frequency (in cm-1) is given by E(cm-1)=107/λ(nm). The wavenumber scale is to be preferred since it is linear in energy and consequently corresponds better with the molecular energy level scheme than the wavelength scale which is reciprocal.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   3  

experiment.  The  quantitative  relation  between  absorbance,  extinction  coefficient,  

concentration  and  pathlength  is  given  by  the  Bourger-­‐Lambert-­‐Beer  law.  It  is  defined  as  

follows:  

! =

A(")c !d

            (  1)  

where   A =! log10(I I

0)  is  the  measured  absorbance, I0

 is  the  intensity  of  the  incident  

light  at  a  given  wavelength   ! ,  I  is  the  transmitted  intensity;  c  is  the  concentration  (in  

mol⋅l-­‐1)  and  d  is  the  pathlength  (in  cm).  Broadly  speaking,  the  electronic  transition  

energies  correspond  to  absorption  maxima  in  the  UV/vis  spectrum.  The  integral  under  

an  absorption  band  characterizes  the  intensity  of  a  given  electronic  transition.  The  

excitation  energies  may  be  given  on  the  wavelength  ( ! ,  in  nm),  frequency  ( ! ,  in  Hz),  

wavenumber  ( !! ,  in  cm-­‐1),  eV,  and  atomic-­‐unit  scale  (Eh).  The  reason  for  introducing  the  

wavelength,  frequency  and  wavenumber  as  energy  units  in  spectroscopy  lies  in  the  

Planck-­‐Einstein  relation  between  these  parameters  and  the  energy  of  incident  photon:  

E = h! = hcL/" = hc

L!!         (  2)  

Here   cL  is  the  speed  of  light,  and   h is  Planck’s  constant.  The  conversion  factors  between  

different  energy  units  conventionally  used  in  spectroscopy  are  given  in  Table  1  Table  1:  Conversion  factors  between  the  energy  units  used  in  spectroscopy.  

  eV   Hz   cm-­‐1   Eh  eV   1   2.4179696×1014   8.065479×103   3.674901×10-­‐2  Hz   4.1357012×10-­‐

15  1   3.3356412×10-­‐

11  1.519829×10-­‐

16  cm-­‐1   1.239852×10-­‐4   2.9979243×1010   1   4.556333×10-­‐6  Eh   27.21161   6.579686×1015   2.194747×105   1  

 

An  absorption  spectrum  basically  consists  of  a  number  of  absorption  bands.  Each  

absorption  band  corresponds  to  a  transition  of  the  ground-­‐  electronic  state  to  an  excited  

electronic  state.  For  reasons  to  be  discussed  below,  however,  such  transitions  do  not  

take  the  appearance  of  sharp  absorption  lines  (as  in  the  spectra  of  atoms  and  ions)  but  

are  usually  considerably  broadened.  In  many  cases  there  will  be  overlapping  bands  and  

one  is  faced  with  the  problem  of  how  to  deconvolute  the  broad  absorption  envelope  into  

contributions  from  individual  transitions.  In  most  cases,  one  simply  performs  a  “Gauss-­‐

Fit”.3  That  is,  one  assumes  that  the  shape  of  each  individual  band  is  that  of  a  Gaussian  

                                                                                                               3 This is an approximate procedure which is not without problems. First of all, accurate band-shapes do not follow the shapes of Gaussian functions. Secondly, the fits are often not well defined and unless one has

Computational  UV/vis,  IR  and  Raman  Spectroscopy   4  

functions  and  then  applies  as  many  (or  as  few)  Gaussian  functions  as  are  necessary  in  

order  to  accurately  represent  the  absorption  envelope.  A  typical  example  is  shown  in  

Figure  2.  

0

2

4

122

613

3

ε ( m

M -1

cm

-1 )

Δε

( M -1

cm

-1 )

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

Δε ( m

M -1 cm

-1 )

2

9

10

12

10

3

6

9

12

30 25 20 15 10

-0.9

-0.6

-0.3

0.0

0.3

0.6

23

A.

B.

13

5

8

9

Wavelength (nm)

Wavenumber (1000 cm-1)

C.

350 550 750 9501150

 Figure  2:  Deconvolution  of  an  absorption  spectrum  into  contributions  from  individual  electronic  transitions  using  Gauss  fits.  The  upper  panel  represents  the  absorption  spectrum  plotted  on  a  wavenumber  scale.  The  second  panel  represents  the  MCD  spectrum  and  the  lower  panel  the  CD  spectrum  of  the  same  compound.  Note  how  the  MCD  and  CD  spectra  help  the  deconvolution  due  to  the  fact  that  they  are  signed  quantities.  

Having  obtained  a  reasonable  Gauss-­‐fit,  it  is  possible  to  calculate  the  so-­‐called  oscillator  

strength4   f0!I  of  the  transition  from  the  ground  state  to  the  I’th  electronically  excited  

state.  The  oscillator  strength  is  simply  proportional  to  the  area  of  the  absorption  band:  

        f0!I

=4.32 "10#9

n! I( )(!")d!"

Band$         (  3)  

(note  that   !I( )(!")  is  given  in  wavenumbers).  The  refractive  index   n  is  usually  set  equal  

to  one.    

The  phenomenon  on  electronic  circular  dichroism  (CD)  consists  in  that  the  left-­‐handed  

and  right-­‐handed  circularly  polarized  light  are  not  only  propagated  with  different  

velocities,  but  they  are  also  absorbed  to  different  extents  in  the  region  of  an  electronic  

absorption  band.  This  behavior  is  most  simply  described  by  the  difference  of  the  molar  

                                                                                                                                                                                                                                                                                                                                                         additional information – for example from CD or MCD spectra – it is advisable to view the fit results with some caution. Additional Gaussians should not be used to improve the fits unless there is at least a well defined maximum or a well defined shoulder in the absorption envelope.

4 Its name stems from the classical dispersion theory and amounts to the number of virtual oscillators equivalent to a given transition.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   5  

decadic  absorption  coefficients  corresponding  to  the  left-­‐handed  and  right-­‐handed  

circularly  polarized  light   !! = !L" !

R.  Similar  to  the  case  of  absorption  the  CD  

spectrum  can  be  deconvoluted  into  individual  bands  corresponding  to  different  

electronic  transitions.  The  fact  that  each  electronic  band  in  the  CD  signal  can  be  of  

positive  or  negative  sign  allows  to  resolve  ambiguity  in  the  spectral  deconvolution.  For  

example,  two  stongly  overlapping  bands  in  an  absorption  spectrum  can  have  different  

signs  in  the  corresponding  CD  spectrum,  thus  providing  the  evidence  for  two  electronic  

transitions  in  the  given  spectral  range.  Typically,  CD  spectra  are  analyzed  together  with  

corresponding  absorption  spectra,  since  they  arise  from  the  same  set  of  electronic  

transitions.  Each  band  in  CD  spectrum  corresponding  to  the  transition  from  the  ground  

state  to  the  I’th  electronically  excited  state  is  characterized  by  the  the  so-­‐called  rotatory  

strength   R0!I  which  is  calculated  as  following  

R

0!I= 0.229"10#38 $! I( )(!")

d!"!"Band%           (  4)  

in  units  of  esu2·cm2,  where   !!  is  in  units  of  molar  extinction  coefficient.    

1.2.1.2 Elementary  Discussion  of  Electronic  Transitions  On  a  most  elementary  level,  electronic  transitions  correspond  to  the  transitions  from  

one  electronic  configuration  to  another.  An  example  is  shown  in  Figure  3.  What  is  shown  

there  is  the  ground  state  electronic  configuration  of  a  square-­‐planar  transition  metal-­‐

dithiolene  complex.  It  consists  of  a  series  of  molecular  orbitals  that  are  drawn  according  

to  increasing  energy.  According  to  the  Aufbau  principle,  all  MOs  are  filled  by  two  

electrons  and  in  the  present  case  one  unpaired  electron  remains  and  occupied  a  singly-­‐

occupied  MO  (SOMO).  The  MOs  transform  under  the  irreducible  representation  of  the  

molecular  point  group  and  the  symmetry  labels  are  given  for  each  MO.5    

The  many  electron  states  found  from  distributing  the  electrons  among  the  available  

orbitals  also  have  a  symmetry  that  is  designated  by  a  term  symbol.  It  is  given  by  2S+1Γ .  

Here  2S+1  is  the  multiplicity  of  the  electronic  state  (S  is  the  total  spin)  and  Γ  is  its  

symmetry.  In  a  nutshell,  the  symmetry  is  found  by  taking  the  direct  product  of  all  

partially  occupied  MOs.  In  general,  this  leads  to  a  reducible  representation  of  the  

molecular  point  group.  However,  for  non-­‐degenerate  point  groups  (D2h  and  subgroups),  

                                                                                                               5 Note that lowercase symbols are used to represent one-electron functions and uppercase symbols are used to represent many-electron states.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   6  

the  direct  product  directly  yields  the  symmetry  of  the  electronic  state.  In  the  example  

above,  a  single  b2g  MO  is  singly  occupied  which  leads  to  a  2B2g  ground  state.    

Excited  states  are  formed  by  promoting  electrons  from  lower-­‐lying  orbitals  into  partially  

occupied  or  empty  MOs.  For  example,  the  transition  1b1u→2b2g  leads  to  (1b1u)1(2b2g)2  

which  corresponds  to  a  2B1u  excited  state.  Transition  from  1au  to  1b1g  yields  

(1au)1(1b2g)1(2b1g)1  which  corresponds  to  an  electron  states  au⊗b2g⊗b1g=2B3u.  However,  

this  configuration  has  three  unpaired  electrons  and  consequently,  it  also  gives  rise  to    a  4B3u  excited  state.  In  order  to  distinguish  different  states  of  the  same  symmetry  one  

typically  uses  “X”  for  the  ground  state,  the  letters  A,B,C  for  excited  states  of  the  same  

multiplicity  and  a,b,c  for  excited  states  of  different  multiplity.  Thus,  in  the  present  case,  

we  have  identified  a  X-­‐2B2g→A-­‐2B1u  and  X-­‐2B2g→B-­‐2B3u  transitions  as  well  as  a  X-­‐2B2g→a-­‐4B3u  transition.    

 

 Figure  3:   Exemplification  of   electronic   states   and   transitions.   Shown   is   the   ground   state   configuration  with  orbitals   in  order   of   decreasing   energy   (top   to   bottom).   Excited   states   are   essentially   formed  by   promoting   electrons   from   lower  lying  MOs  into  higher  lying  partially  occupied  or  empty  MOs.  

1.2.1.3 Selection  Rules  In  order  to  perform  an  assignment  of  the  electronic  spectrum,  it  is  necessary  to  

determine  the  symmetries  of  the  excited  states  involved  and  identify  the  donor  and  

acceptor  MO  pairs  involved  in  the  transitions.  A  typical  example  is  shown  in  Figure  4.  

In  order  to  find  out  whether  an  electronic  transition  is  allowed  or  not,  the  following  rules  apply:  

• Transitions  between  states  of  different  multiplicity  are  forbidden.  This  is  a  fairly  strong  selection  rule  and  such  transitions  typically  have  ε  values  <  1  M-­‐1  cm-­‐1.6  

                                                                                                               6 This selection rule is lifted by spin-orbit coupling. For heavier elements formally “spin-forbidden” transitions become increasingly allowed.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   7  

• The   direct   product   of   the   irreps   of   the   initial   and   final   states   involved   the  transition  with   the   irreps   spanned  by   the  electric  dipole  operator  must   contain  the   totally   symmetric   representation.   Otherwise   the   transition   is   forbidden   by  symmetry.  

 

The  electric-­‐dipole  operator  transforms  like  the  molecular  translations  which  can  be  

found  with  the  labels  “x”,  “y”  and  “z”  in  usual  group  theoretical  tables.7  As  a  consequence  

of  these  rules:  

• If  the  point  group  contains  a  center  of  inversion,  only  g→u  transitions  are  

allowed  (and  vice  versa).  Other  transitions  are  said  to  be  “parity  forbidden”.    

The  selection  rules  are  a  good  guide  to  identifying  strongly  and  weakly  allowed  

transitions  in  a  spectrum.  However,  don’t  be  surprised  to  find  a  formally  forbidden  

transition  in  your  spectrum.  In  the  real  world  there  are  almost  always  additional  

perturbations  that  break  the  symmetry  of  the  system  and  make  formally  forbidden  

transitions  weakly  allowed.8  

0

1

2

3

4

MCDABS

calc

exp

8

7

6

54

3

287

65

42 -1

0

1exp

Δε

(mM

-1 c

m-1 T

-1)

ε (m

M-1 c

m-1)

25 20 15 100

1

2

3

calc

2a2g

3a1g

2b1g

1b3g

1b2g

1b1g1ag

2a2g

calc

Wavenumber (1000 cm-1)

3

25 20 15 10

-0.2

0.0

0.2calc

3ag2b1g

1b3g

1b2g

1b1g

1ag

2a2g

 Figure  4:  An  example  for  an  assigned  absorption  spectrum.  The  experimental  absorption  and  MCD  spectra  are  given  in  the  upper  panel  and  calculated  spectra  are  shown  in  the  lower  panel.  The  acceptor  MO  is  always  a  b3u  level  in  this  case  and   the   lowercase   symbols   on   the   calculated   transitions   merely   represent   the   symmetry   of   the   donor   MO.   The  theoretical   calculations   yield   “only”   the   black   bars   in   the   lower   panels   which   have   been   empirically   convoluted   by  

                                                                                                               7 The most popular text in chemistry is F.A. Cotton: Chemical Applications of Group Theory, John Wiley and Sons, New York, 1990. It has very useful tables. A more theoretically oriented text is R. McWeeny Symmetry – An introduction to Group Theory and Its Applications. Dover, New York, 2002.

8 A typical example are the parity forbidden d-d transitions studied in chapter Error! Reference source not

found..

Computational  UV/vis,  IR  and  Raman  Spectroscopy   8  

Gaussian  functions  in  order  to  produce  an  envelope  that  can  be  compared  with  the  experimental  measurement.  For  this  example  see9  

Since  the  rotational  strength  of  a  transition  is  proportional  to  the  scalar  product  of  the  

electric  and  magnetic  dipole  transition  moments  (see  the  next  section)  the  selection  

rules  are  different  than  those  for  the  case  of  normal  absorption.  Since  g→g  and  u→u  

transitions  are  formally  electric  forbidden,  whereas  g→u  and  u→g  are  magnetic  dipole  

forbidden,  it  is  immidiatley  apperent  that  centrosymmetric  molecules  are  not  optically  

active.  More  generally,  it  can  be  shown  that  a  molecule  must  lack  any   Sn  axis  (including  

S1! !  and   S2

! i )  to  be  optically  active,  which  translates  to  only  those  molecules  with  

nonsuperimpsable  mirror  images.  Such  molecules  are  called  chiral.  Enantiomers,  which  

are  stereoisomers  that  are  nonsuperimposable  complete  mirror  images  of  each  other,  

show  CD  signals  of  equal  amplitudes  but  different  signs.  Only  molecules  belonging  to  the  

point  groups   Cn, Dn

, O , T ,  or   I  are  optically  active  and  discrete  transitions  must  be  

both  electric  and  magnetic  dipole  allowed  to  exhibit  CD  .  

1.2.1.4 Refined  Discussion  of  Electronic  Transitions  In  this  section,  we  will  purse  a  slightly  more  “physical”  description  of  the  aborption  

processes  that  leads  the  way  towards  a  more  quantitative  description  of  the  spectra.  The  

theoretical  description  of  the  absorption  process  rests  on  two  major  assumptions:  The  

first  assumption  is  to  assume  that  the  strength  of  the  external  electromagnetic  field  

(provided  by  the  electromagnetic  wave)  is  much  smaller  than  the  electric  fields  that  act  

“inside”  the  molecules.  In  such  a  case  the  external  electromagnetic  force  can  be  reliably  

modeled  as  a  small  perturbation.  In  this  case,  the  radiation  may  be  described  as  a  time-­‐

dependent  electromagnetic  field  and  the  vast  majority  of  observable  electronic  

transitions  are  well  interpreted  within  the  first-­‐order  time-­‐dependent  perturbation  

treatment.  Secondly,  it  is  assumed  that  the  wavelength  of  light  is  much  larger  than  the  

dimensions  of  the  investigated  molecules.  Thus,  the  oscillating  electromagnetic  field  is  

assumed  to  be  essentially  uniform  over  the  extension  of  the  molecule.  This  is  the  essence  

of  the  so-­‐called  electric-­‐dipole  approximation.    

Both  assumptions  can  and  need  to  be  refined  in  special  circumstances,  like  very  intense  

electromagnetic  fields  provided  by  lasers  or  very  short  wavelengts  used  in  X-­‐ray  

                                                                                                               9 Neese, F.; Zumft, W.G.; Antholine, W.E.; Kroneck, P.M.H.; (1996) J. Am. Chem. Soc., 118, 8692-8699; Farrar, J.; Neese, F.; Lappalainen, P.; Kroneck, P.M.H.; Saraste, M.; Zumft, W.G.; Thomson, A.J. (1996) J. Am. Chem. Soc. 118:11501-11514.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   9  

absorption  spectroscopy.  Taking  into  account  higher-­‐order  terms  in  the  time-­‐dependent  

perturbation  approach  leads  to  multi-­‐photon  non-­‐linear  processes,  for  which  molecular  

transitions  occur  with  simultaneous  absorption  of  two  and  more  photons.  There  are  also  

corrections  to  the  long-­‐wavelength  approximation  that  describe  magnetic  dipole,  

electric  quadrupole  and  higher-­‐order  multipole  transitions  which  usually  give  rise  to  

low-­‐intensity  spectral  lines  in  ordinary  experiments.    

On  the  thereotical  side,   fik  can  be  related  to  the  electronic  structure  of  the  molecule  

under  investigation.  The  relevant  expression  for  the  oscillator  strength  in  the  electric-­‐

dipole  approximation  takes  the  form  (SGS  units):  

f0!I

=4!m

e

3e2!"

0ID

0!I

2           (  5)  

Alternatively,  in  atomic  units:  

f0!I

=23E

0ID

0!I

2             (  6)  

E0I= h!

0I  is  the  transition  energy,   D0!I

 is  the  transition  dipole  moment  matrix  

element  between  the  ground-­‐  and  excited  state  I  is  defined  as  follows:  

D

0!I= "

0(r,R)d"

I(r,R)drdR#           (  7)  

      d = Z

AR

AA! " r

ii!             (  8)  

Here,   d  is  the  electric  dipole  operator  which  consists  of  a  sum  over  nuclei  (A)  with  

charges  ZA  and  electrons   ri .10  The  functions   !0

(r,R)  and   !I(r,R)  represent  the  many-­‐

electron  wavefunctions  that  describe  the  ground  and  final  states  of  the  system  

respectively.    

The  rotatory  strength  depends  on  the  scalar  product  of  the  transition  dipole  moment  

D0!I  and  transition  magnetic  dipole  moment   M0!I

 matrix  elements.  If   D0!I  and   M0!I

 

are  given  in  atomic  units  then   R0!I  in  units  of  [esu2·cm2]  is  obtained  according  to  the  

formula:  

  R

0!I= 4.71450602"10#38 D

0!I$M

0!I( )         (9)  

                                                                                                               10 The positive and negative signs reflect the positive and negative charges of electrons and nuclei respectively.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   10  

where   M0!I  is  defined  by  the  following  expression:  11  

  M

0!I= "

0(r,R)µ"

I(r,R)drdR#      

  (10)  

  µ = 1

2(l

i+ 2s

i)

i!            

  (11)  

In  the  BO  approximation   !I(r,R)  describes  the  joint  electronic,  vibrational  and  

rotational  state  of  the  system.  Thus,  as  mentioned  above,  an  electronic  absorption  band  

also  contains  vibrational  and  rotational  structure;  Although  for  small  molecules  in  the  

gas  phase  this  rovibrational  structure  may  be  resolved  and  analyzed,  the  UV/vis  

absorption  spectra  of  larger  molecules  are  usually  measured  in  solution.  Since  the  

rotational  quanta  are  small  and  the  rotation  is  hindered  by  the  solvent  molecules,  no  

rotational  structure  is  seen  in  solution,  even  with  high-­‐resolution  spectrometers.  Thus,  

any  remaining  structure  on  the  absorption  band  –  if  any  can  be  resolved  -­‐  may  be  

attributed  to  vibrations.  Upon  the  neglect  of  rotations,   !S(r,R)  may  be  written  as  

vibronic  wavefunctions  of  the  form:  

!In(r,R) = "

I(r,R)!

n

I( )(R)             (  

12)  

Here,   !n

I( )(R)  is  the  nuclear  vibrational  wavefunction  and  I  and  n  are  the  indices  of  the  

electronic  and  vibrational  states,  respectively.  Even  within  the  limitations  of  the  BO  

approach,  there  are  a  number  of  approximations  that  are  usually  made  in  order  to  make  

the  analysis  and  calculations  of  the  absorption  spectra  of  molecules  feasible:  

1. Molecules   are   assumed   to   be   excited   from   the   ground   vibrational   state   of   the  

electronic  ground  state.  (this  is  the  zero-­‐temperature  approximation).  Typically  the  

account   of   thermal   population   of   the   excited   vibrational   levels   leads   to   some  

broadening  of  the  absorption  band.  

                                                                                                               11 Note that M0!I defined by eqs.(10)-(11) should be multipled by the fine-structure constant

! = 1/ 137.039 in order to obtain actual transition magnetic dipole moment in atomic units. For the given

definition of M0!I ! is included in the numerical coefficient in eq. (9).

Computational  UV/vis,  IR  and  Raman  Spectroscopy   11  

2. The  electronic  transition  dipole  moment  defined  as  

d

0I(R) = !

0(r,R)d!

I(r,R)dr"  does  not  depend  upon  nuclear  coordinates,  and  equal  

to  the  value  corresponding  to  the  ground-­‐state  equilibrium  geometry   R0:  

d0I(R)! d

0I(R

0) = const .  This  statement  constitutes  the  essence  of  the  so-­‐called  

Franck-­‐Condon  approximation.  

3. Due   to   the   vibronic   nature   of   the   molecular   states   the   excitation   of   the   0! k  

electronic   transition   implies   a   co-­‐excitation   of   of   vibrational   quanta   from   the   ground  

vibrational  state  of  the  electronic  ground  state  into  the  ground-­‐  and  excited  vibrational  

states  on  the  k’th  electronic  state.  For  a  given  electronic  transition   0! k  the  excitation  

of  the  vibrational  quanta  in  a  given  normal  mode  constitute  a  vibrational  progression.  

In  the  general  case,  the  intensity  distribution  within  such  a  progression  corresponding  

to  transitions  of  the  form   !0(r,R)!

0

0( )(R)" !I(r,R)!

n

I( )(R)  is  determined  by  the  square  

of   the   vibrational-­‐overlap   integrals   !

0

0( ) !n

I( )2

 (Franck-­‐Condon   factors).   Their  

precise   calculation   for   larger   molecules   and   realistic   PESs   is   a   quite   complicated  

problem  and  calls  for  further  simplifications.  The  Franck-­‐Condon  (FC)  principle  states  

that   the   electronic   ground-­‐state   vibrational   wavefunction   !0

0( )(R)  has   large   overlap  

with  the  excited-­‐state  wavefunctions   !n

I( )(R)  only  if  their  energy  levels  are  close  to  the  

vertical  transition.  The  vertical  transition  occurs  when  the  molecule  is  promoted  from  

the   electronic   ground   state   to   the   electronically   excited   state   at   the   ground   state  

equilibrium  geometry  (Figure  5).12  

1.2.1.5 Vibrational  Fine  Structure  Although  the  absorption  spectra  of  various  chemical  species  are  characterized  by  

vibrational  structure,  the  latter  is  frequently  obscured  due  to  intermolecular  collisions,  

solvent  effects  and  spectral  crowding.  In  such  a  case  and  absorption  band  may  be  viewed  

as  the  envelope  of  different  vibronic  transitions  (Figure  5).  Its  intensity  (area  under  the  

absorption  band)  is  given  by  the  oscillator  strength.  The  simultaneous  neglect  of  

vibrational  and  rotation  finestructure  together  with  the  zero-­‐temperature  and  Franck-­‐

Condon  approximation  on  harmonic  potential  energy  surfaces  present  the  most  

                                                                                                               12 The FC principle can be rigorously derived only for diatomic harmonic PES’s and for many-atom molecules and more complicated shapes of the PESs it should be used with caution.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   12  

simplified  level  of  theoretical  description.  It  only  requires  the  knowledge  of  the  energies  

and  electronic  wavefunctions  of  the  ground  and  excited  states  which  are  all  evaluated  at  

the  ground-­‐state  equilibrium  geometry.  Despite  the  many  simplifying  assumptions  this  

already  represents  a  very  useful  level  of  theory.  

The  prediction  of  the  vibrational  structure  of  the  absorption  spectra  is  based  on  the  

calculations  of  vibrational-­‐overlap  integrals   !

0

0( ) !n

I( )2

 which  in  general  case  depend  in  

a  very  complicated  way  on  the  shape  of  PES’s.  However,  under  the  simplifying  

assumption  of  harmonic  potentials,  the  problem  can  be  solved  exactly.  In  this  case  the  

distribution  of  intensity  between  various  vibronic  bands  only  depends  on  the  

equilibrium  shift,13  frequency  alteration14  and  normal  mode  rotation  in  the  excited  

state,15  such  that  the  entire  absorption  bandshape  can  be  written  in  the  closed  form  as  a  

function  of  these  parameters.  Of  these  factors,  only  the  first  one  –  the  equilibrium  shift  –  

is  of  major  importance.  The  neglect  of  the  normal  mode  rotations  leads  to  the  

Independent  Mode  Displaced  Harmonic  Oscillator  (IMDHO)  model.  In  this  simplified  

framework,  the  vibrational  progression  is  determined  solely  by  the  equilibrium  shift  and  

vibrational  frequency  alteration  between  the  ground-­‐  and  excited-­‐state  PES’s.16    

                                                                                                               13 The quantities Δ are defined above

14 The change of vibrational frequencies in the excited state due to the change of bonding.

15 This is a fairly complicated effect known as “Duschinsky rotation”. Is arises from the fact that the force field of the molecule is different for each electronic state. Thus, the shapes of normal modes also change from state to state which leads to additional complexity in the calculation of FC factors.

16 For a detailed recent discussion see Petrenko, T.; Neese, F. (2007) Analysis and Prediction of Absorption Bandshapes, Fluorescence Bandshapes, Resonance Raman Intensities and Excitation Profiles using the Time Dependent Theory of Electronic Spectroscopy. J. Chem. Phys., 127, 164319; Neese, F.; Petrenko, T.; Ganyushin, D.; Olbrich, G. (2007) Advanced Aspects of ab initio Theoretical Spectroscopy of Open-Shell Transition Metal Ions. Coord. Chem. Rev., 205, 288-327.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   13  

 Figure   5:   Illustration   of   the   Franck-­‐Condon   principle.   The   excitation   into   electronic   band   involves   the   coexcitation   of  vibrational  levels  which  results  in  the  vibrational  progression  of  the  absorption.  The  vibronic  peak  of  maximum  intensity  corresponds  to  the  excited-­‐state  vibrational   level  which   is  most  close  to  the   intersection  of   the  vertical   transition   (red  line)  with   the   excited-­‐state   PES.  As   a   result   the  maximum  of   the   broadened   absorption   spectrum  occurs   close   to   the  vertical  transition  energy.  For  the  IMDHO  model  the  number  of   intense  peaks   in  the  absorption  progression  correlates  with  the  value  of  Δ.  

 Let  us  consider  in  more  detail  the  quantitative  aspects  of  the  vibrational  structure  in  

optical  spectra.  In  the  harmonic  approximation  the  nucler  motion  corresponding  to  the  

ground  and  excited  electronic  states  is  described  according  to  the  Hamiltonians   HN

(0)  and  

HN

(I ) ,  respectively:  

          H

N

(0) = !!2

2Mii

3M

" #i

2 +12

Vij

(0)X

iX

j

ij

3M

"  

  (13)  

          H

N

(I ) = !!2

2Mii

3M

" #i

2 +12

Vij

(I )X

i!X

i0(I )( ) X

j!X

j 0(I )( )

ij

3M

"  

  (14)  

where   V

ij(0){ }  and   Vij

(I ){ }  are  the  Hessian  matrices  for  the  ground  and   I -­‐th  excited  state.  

X

i{ }  are  the  Cartesian  coordinates  of  the  nuclei  in  the  frame  of  reference  which  origin  corresponds  to  the  equilibrium  positions  of  the  nuclei  for  the  ground  electronic  state.  

X

i0(I ){ }  are  the  Cartesian  displacements  of  the  nuclei  equlibrium  positions  in  excited  

state   I  relative  to  the  ground  state.  

Computational  UV/vis,  IR  and  Raman  Spectroscopy   14  

Note,  that  if   HN

(0)  and   HN

(I )  are  identical,  i.e.   V

ij(I ) =V

ij(0)  and   Xi0

(I ) = 0  for  all   i , j ,  then  the  

ground-­‐  and  excited-­‐state  vibrational  Hamiltonians  are  characterized  by  the  common  set  

of  vibrational  eigenfunctions,  so  that   !

0

0( ) !0

I( ) = 1 ,  whreas  all  other  integrals  

!

0

0( ) !n

I( ) = 0 .  As  soon  as  the  equilibrium  positions  of  atoms  in  the  excited  state  are  

changed  compared  to  the  ground  state  ( Xi0(I ) ! 0 ),  and/or  there  is  a  change  in  the  force  

constants   V

ij(I ) !V

ij(0) ,  then  

!

0

0( ) !0

I( ) <1 ,  and  there  will  be  certain  vibrational  levels   !n

I( )  

for  which   !

0

0( ) !n

I( ) ! 0 .  This  means  that  one  would  observe  different  transitions  

between  the  ground  vibrational  level  in  the  ground  state  and  various  vibrational  levels  

corresponding  to  the  excited  state.  The  intensity  pattern  of  such  a  vibrational  structure  

depends  in  a  rather  complicated  way  on  the  difference  between   HN

(0)  and   HN

(I ) .    

Upon  appropriate  linear  transformation  from  Cartesian  to  normal  coordinates  it  is  

possible  to  obtain  decoupled  representation  of   HN

(0)  and   HN

(I ) ,  which  eigenfunctions  are  

the  products  of  one-­‐dimensional  Hermitian  functions  of  normal  coordinates.  

Consider  particular  case     V

ij(I ) =V

ij(0) =V

ij  for  all   i , j ,  for  which,  as  can  be  shown,  there  is  

no  change  in  vibrational  frequencies  and  shapes  of  normal  modes  between  the  ground  

and  excited  states.  Let  us  assume  also  that  the  equilibrium  positions  of  atoms  in  the  

excited  state  are  changed  compared  to  the  ground  state  ( Xi0(I ) ! 0 ).  A  given  vibrational  

state !n

I( )  is  completely  determined  by  a  set  

n!{ }  of  occupation  numbers  for  individual  

modes  with  frequencies   !" .  In  such  a  case  

          !

0

0( ) !n

I( )2

=(s")n"

n"!

e!s"

"

3M!6

"  

  (15)  

where   s!

=(!

!)2

2  are  the  so-­‐called  Huang-­‐Rhys  factors,  and   !! is  displacement  of  the  

excited-­‐state  origin  relative  to  the  ground-­‐state  one  along  dimensionless  normal  

coordinate   !  which  can  be  related  to   X

i0(I ){ }  as  follows:  

Computational  UV/vis,  IR  and  Raman  Spectroscopy   15  

        !!

="!

!

"

#$$$$

%

&'''''

1/2

Uk!

Xi0(I ) M

kk=1

3M

(  

  (16)  

U

m!{ }  is  the  orthogonal  matrix  obtained  upon  numerical  diagonalization  of  the  mass-­‐weighted  Hessian  matrix.  

Thus,  one  can  observe  in  a  given  electronic  band  a  manifold  of  vibrational  transitions  at  

energies    

        E

n= E

0+ n

!!"!

!

3M!6

"   (17)  

corresponding  to  different  sets  

n!{ } ,  which  have  relative  intensities  given  by  the  

values  of   !

0

0( ) !n

I( )2

 in  eq.  (15).   E0  is  the  separation  between  the  ground-­‐  and  excited-­‐

state  PES  minima.  The  vertical  transition  energy  and   E0  can  be  related  in  the  following  

way:  

E

V= E

0+ !!

"s"

"

3M!6

"   (18)  

In  general,  the  intensity  pattern  can  be  quite  complicated  if  there  are  many  modes  

characterized  by  significant  displacements.  As  one  can  see,  for  the  single-­‐mode  species,  

like  diatomic  molecules,  the  relative  intensities  are  given  by  the  values  

sn

n !e!s  which  

result  in  a  rather  simple  character  of  vibrational  progression.  Figure  5  illustrates  this  

case  for   != 2.9 .  For  the  one-­‐dimensional  species  one  can  show  that   n  corresponding  

to  the  most  intense  vibrational  transition  satisfies  the  condition   n ! s ,  and  the  number  

of  the  most  significant  vibrational  transitions  is  close  to   2s .  

1.2.2 Quantum  Chemical  Calculation  of  Absorption  Spectra  The  calculation  of  electronically  excited  states  of  molecules  presents  a  number  of  

additional  challenges  to  quantum  chemistry  that  can  not  be  discussed  in  depth  in  the  

framework  of  this  course  (for  some  aspects  of  accurate  treatments  see  chapter  Error!  

Reference  source  not  found.,  in  particular  Error!  Reference  source  not  found.).  Here  

we  pursue  a  rather  simple  picture  of  the  excitation  process  as  involving  the  promotion  

of  an  electron  from  an  occupied  molecular  orbital  in  the  ground-­‐state  into  an  

unoccupied  orbital  (one-­‐electron  picture).  Within  this  approximation,  the  assignment  of  

Computational  UV/vis,  IR  and  Raman  Spectroscopy   16  

an  absorption  spectrum  consists  of  the  determination  which  donor  and  acceptor  MOs  

give  the  dominant  contribution  to  the  observed  absorption  bands  as  exemplified  above.  

In  general,  this  picture  is  oversimplified  since  the  excited  state  is  usually  not  well  

represented  in  terms  of  a  single  Slater  determinant  and  secondly,  the  orbitals  of  the  

ground  state  will  not  be  appropriate  for  the  electronically  excited  state.  Nevertheless,  

the  simplest  reasonable  approximation  is  to  write  the  ground  state  as  a  single  

determinant  (for  example,  the  Hartree-­‐Fock  determinant):  

  !

0= !

1...!

i...!

N                   (  

19)  

And  the  excited  states  may  be  assumed  to  be  represented  as  a  superposition  of  singly-­‐

excited  determinants:  

  !

I= t

ia!

ia

i,a"                   (  20)  

Where  the  „excited  determinants“   !ia  have  an  occupied  MO   !i

 replaced  by  a  virtual  MO  

!a  and  the  quantities   ti

a  are  the  “amplitudes”  of  the  excitation.  The  quantity   tia( )2  is  the  

“weight”  of  the  excited  determinant   !ia  in  the  excited  state   !I

.  The  amplitudes  are  

determined  from  solving  the  so-­‐called  configuration-­‐interaction  with  single  

excitations  (CIS)  eigenvalue  problem.  The  eigenvalues  of  this  matrix  correspond  to  the  

(vertical)  excitation  energies  at  this  level  of  approximation  and  the  eigenvectors  are  the  

excitation  amplitudes.17  Calculations  in  the  CIS  framework  do  not  lead  to  very  accurate  

results  but  are  suitable  to  provide  an  initial  orientation  prior  to  more  accurate  

calculations  and  is  also  feasible  for  larger  molecules.    

In  the  framework  of  density  functional  theory  (DFT),  one  employs  the  so-­‐called  time  

dependent  density  functional  theory  (TD-­‐DFT)  in-­‐order  to  predict  vertical  transition  

energies.  TD-­‐DFT  is  based  on  a  different  philosophy  than  CIS  but  leads  to  equations  that  

are  fairly  similar  and  that  can  be  solved  with  comparable  computational  effort.  The  

accuracy  of  TD-­‐DFT  is  usually  superior  to  that  provided  by  CIS.18  

In  ORCA  CIS  or  TD-­‐DFT  calculations  are  invoked  with  the  block:  

                                                                                                               17 Note that there is a different set of excitation amplitudes for each excited state.

18 However, there are many problems with that method and a number of artifacts are known that one should be aware of before applying the method in actual calculations. Some of the relevant aspects are 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

Computational  UV/vis,  IR  and  Raman  Spectroscopy   17  

 For  HF  reference  wavefunction  (RHF  or  UHF)  the  program  automatically  chooses  CIS  

calculations  and  for  DFT  model  (RKS  or  UKS)  TD-­‐DFT.  Below  is  a  summary  of  variables  

that  can  be  assigned  within  the  block:  

 Table  2:  Variables  for  CIS  or  TD-­‐DFT  calculations  with  the  ORCA  program.  

Variable   Description  NRoots=N Calculate  the  first  N  excited  states  Triplets Parameter  is  only  valid  for  closed  shell  references.  If  chosen  as  true  the  program  

will  also  determine  the  triplet  excitation  energies  in  addition  to  the  singlets  Ewin e0, e1 Slects  the  MOs  within  the  orbital  energy    range  [e0,  e1]  (Eh)  to  be  included  in  the  

correlation  treatment  MaxDim The  maximum  dimension  of  the  expansion  space  in  the  Davidson  procedure.  If  

MaxDim  ≈  5-­‐10  times  NRoots  the  calculations  will  show  favorable  convergence  but  also  increased  disk  space  demands  

Etol Gives  the  required  convergence  of  the  energies  of  the  excited  states  (in  Eh)  RTol Gives  the  required  convergence  on  the  norm  of  the  residual  vectors.  TDA Parameter  is  valid  for  TD-­‐DFT  treatment.  If  chosen  as  true  the  so-­‐called  Tamm-­‐

Dancoff  approximation  (TDA)  is  invoked.  For  hybrid  functionals  the  choice  TDA=true  is  required  and  will  be  enforced  by  the  program.  

Mode=riints   Invokes  the  method  that  allows  to  use  RI-­‐integrals  transformed  to  the  MO  basis  to  generate  the  CI  matrix.  It  shows  speedups  on  the  order  of  10  or  more.  An  auxiliary  basis  set  needs  to  be  assigned  if  this  option  is  used.  

 

Here  is  the  key  part  of  the  ORCA  output  from  a  TD-­‐DFT  or  CIS  calculation:  

 In  this  part  of  the  output  the  results  on  each  state  are  summarized,  including  the  

excitation  energy  (in  atomic  units,  eV  and  cm-­‐1),  and  the  largest  amplitudes  (squared)  in  

the  CI  expansion.  For  example,  the  1st  excited  state  is  mainly  presented  by  one  leading  

configuration  (with  the  weight  of  0.988777)  in  which  the  spin-­‐up  orbital  78  (the  “donor  

MO”)  in  the  reference  determinant  is  replaced  with  the  virtual  orbital  80  (the  “acceptor  

%cis end # or equivalently %tddft end

------------------------------------ TD-DFT/TDA EXCITED STATES (SINGLETS) ------------------------------------ the weight of the individual excitations are printed if larger than 0.01 STATE 1: E= 0.095176 au 2.590 eV 20888.6 cm**-1 78a -> 80a : 0.988777 (c= 0.99437248) STATE 2: E= 0.108238 au 2.945 eV 23755.5 cm**-1 79a -> 80a : 0.945608 (c= -0.97242400) 79a -> 83a : 0.010346 (c= 0.10171342)

Computational  UV/vis,  IR  and  Raman  Spectroscopy   18  

MO”);  the  symbol  “a”  in  the  output  next  to  the  orbital  number  means  that  this  is  a  spin-­‐

up  orbital;  alternatively,  “b”  corresponds  to  spin-­‐down  orbitals.  

Transition  energies  and  oscillator  strengths  for  the  calculated  excited  states  are  given  in  

the  following  part  of  the  output:  

 The  column  fosc  shows  calculated  oscillator  strengths.  TX,  TY,  TZ  are  the  components  

of  electronic  transition  dipole  moment  (in  atomic  unit),  T2  is  the  square  of  transition  

moment.    

The  output  of  rotatory  strengths  for  the  calculated  excited  states  has  similar  structure:    

 The  column  R  shows  calculated  rotatory  strengths.  MX,  MY,  MZ  are  the  components  of  

electronic  transition  dipole  moment  (in  atomic  unit).    

If  you  want  to  obtain  a  plot  of  the  absorption  spectrum  then  call  the  small  utility  

program  orca_mapspc:  

 The  program  will  produce  outputfile.abs.dat file  containing  absorption  

spectrum  that  can  be  plotted  with  standard  graphics  programs.  Options  are  explained  

here:  

 

 

 -x0value Start  of  the  x-­‐axis  for  the  plot  -x1value End  of  the  x-­‐axis  for  the  plot  -wvalue Full-­‐width  at  half-­‐maximum  height  in  cm-­‐1  for  each  transition  

----------------------------------------------------------------------------- ABSORPTION SPECTRUM ------------------------------------------------------------------------------ State Energy Wavelength fosc T2 TX TY TZ (cm-1) (nm) (au**2) (au) (au) (au) ------------------------------------------------------------------------------ 1 20888.6 478.7 0.094387411 1.48758 0.08858 -1.21644 -0.00006 2 23755.5 421.0 0.167847587 2.32609 -0.00002 -0.00013 -1.52515 3 32083.8 311.7 0.295677742 3.03395 0.13992 1.73619 0.00009

----------------------------------------------------------------------------- CD SPECTRUM ------------------------------------------------------------------------------ State Energy Wavelength R MX MY MZ (cm-1) (nm) (1e40*cgs) (au) (au) (au) ------------------------------------------------------------------------------ 1 20888.6 478.7 3.659617854 0.32713 0.01744 -0.00002 2 23755.5 421.0 -166.793655155 0.96929 -0.50766 0.23200 3 32083.8 311.7 329.282848783 0.42056 0.36840 -0.12245

orca_mapspc outputfile abs –x015000 –x135000 –w200 –n2000

Computational  UV/vis,  IR  and  Raman  Spectroscopy   19  

-nvalue Number  of  points  to  be  used    

Thus,  in  the  above  example  the  absorption  spectrum  will  be  produced  for  2000  

equidistant  points  in  the  range  15000-­‐35000  cm-­‐1  using  full-­‐width  at  half-­‐maximum  

height  equal  to  200  cm-­‐1  for  each  transition.  

Likewise,  CD  plot  can  be  obtained  by  calling  orca_mapspc  with  another  option:  

 The  program  will  produce  outputfile.cd.dat file  containing  CD  spectrum.  

1.2.3 IR  and  Raman  spectroscopy  

1.2.3.1 Normal  Modes  and  Vibrational  Frequencies  Infrared  (IR)  and  Raman  spectroscopies  are  mutually  complementary  parts  of  molecular  

vibrational  spectroscopy.  Their  molecular  physics  is  essentially  the  same  since  both  are  

concerned  with  the  observation  of  the  excitation  of  molecular  vibrational  energy  states  

associated  with  the  electronic  ground-­‐state  PES.  In  IR  spectroscopy,  vibrational  

excitations  occur  upon  the  absorption  of  electromagnetic  radiation.  In  Raman  

measurements  the  excitation  is  a  consequence  of  inelastic  light  scattering  by  molecule.  

Upon  the  assumption  that  the  PES  is  quadratic  in  atomic  displacements  (harmonic  

model)  the  vibrational  states  of  an  M-­‐atom  molecule  are  described  as  a  superposition  of  

3M-­‐6  (3M-­‐5  for  linear  molecules)  independent  harmonic  oscillators  which  describe  the  

collective  harmonic  vibrational  motion  of  the  nuclei  about  their  equilibrium  

configuration.  The  shapes  of  different  vibrations  are  characterized  by  different  patterns  

of  the  joint  nuclear  displacements  and  are  called  normal  modes  (coordinates).19  They  

transform  under  the  irreducible  representations  of  the  molecular  symmetry  group  and  

may  therefore  be  classified  by  their  symmetry.  The  textbook  example  of  the  H2O  

molecule  is  shown  in  Figure  6.20  

                                                                                                               19 Their construction will not be pursued in this course.

20 Still the most popular book that describes the calculation and use of normal coordinates from force fields in detail is classic text: Wilson, E.B.; Decius, J.C.; Cross, P.C. Molecular Vibrations – The Theory of Infrared and Raman Spectra, Dover, New York, 1980; the original text is from 1955.

orca_mapspc outputfile cd –x015000 –x135000 –w200 –n2000

Computational  UV/vis,  IR  and  Raman  Spectroscopy   20  

 Figure  6:  The  normal  coordinates  of  the  H2O  molecule  together  with  its  IR  and  Raman  spectra  on  natural  abundance  as  well  as  isotopically  labeled  forms  of  H2O.  

A  quantum  mechanical  treatment  provides  the  exact  vibrational  energies  for  the  

harmonic  oscillator  model:  

E = !!

1(n

1+ 1

2)+ !!

2(n

2+ 1

2)+ !!

3(n

3+ 1

2)+…     (  

21)  

where   ni= 0,1,2,3...are  the  vibrational  quantum  numbers  and   !i

 are  harmonic  

vibrational  frequencies  corresponding  to  the  i-­‐th  normal  mode.  For  a  purely  harmonic  

oscillator  the  solution  of  the  wavefunctions  is  shown  in  Figure  7.  (ω=ν/2π).  

 Figure  7:  Eigenfunctions  and  eigenenergies  of  the  harmonic  oscillator.  

The  vibrational  frequencies  of  the  normal  modes  reflect  the  force-­‐field  or  the  molecule.  

It  is  a  measure  of  the  shape  of  the  PES  of  the  electronic  state  under  investigation  (usually  

the  electronic  ground  state)  which,  in  turn,  is  depends  on  the  strengths  of  the  bonds  

Computational  UV/vis,  IR  and  Raman  Spectroscopy   21  

between  the  different  nuclei.  For  a  diatomic  molecule,  the  force  constant  k  is  simply  the  

second  derivative  of  the  electronic  energy  at  the  equilibrium  distance:  

   

k =!2E I( )

!R2

R=R

                (  

22)  

From  which  the  vibrational  frequency  is  readily  calculated  as:21  

    ! =

12"c

km                   (  

23)  

Where  the  reduced  mass  m  is  defined  in  Figure  8  below.  

 Figure  8:  Vibrations  of  a  diatomic  molecule.  

For  a  general  polyatomic  molecule,  the  definition  of  the  force  constant  must  be  

generalized.  It  is  replaced  by  the  Hessian  matrix  already  used  extensively  in  the  

previous  computer  experiments.  Its  definition  is:  

   

HAB

=!2

EI( )

!XA!X

BX

A=X

B=...X

              (  

24)  

(XA,  XB,…  are  Cartesian  coordinates  of  atoms  A,B,…)  Essentially,  diagonalization  of  the  

mass  weighted  Hessian  matrix  ( !HAB

= HAB

mAm

B( )"1/2

)  yields  the  normal  modes  and  

harmonic  vibrational  frequencies  of  the  system.  

                                                                                                               21 For correct units and conversion factors see chapter Error! Reference source not found.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   22  

1.2.3.2 Selection  Rules  The  important  difference  between  IR  and  Raman  spectroscopies  is  that  for  non-­‐zero  IR  

intensity  there  must  be  a  change  of  dipole  moment  along  a  given  normal  mode.  For  

diatomic  molecules,  this  means  that  they  must  have  a  permanent  dipole  moment  in  

order  to  be  IR  active.  For  nonzero  Raman  intensities  there  must  be  a  change  in  the  

polarizability  during  the  vibration.    

Taking  into  account  only  linear  terms  in  variations  of  dipole  moment  (for  IR)  and  

polarizability  (for  Raman)  during  the  vibration  leads  to  the  vibrational  selection  rule  for  

harmonic  case.  It  states  that  only  excitations  with   !n = ±1  in  only  one  single  mode  are  

allowed  upon  the  vibrational  transition.  For  such  a  case,  IR  and  Raman  spectra  consist  

only  of  fundamental  transitions  at  the  frequencies  which  coincide  with  the  vibrational  

harmonic  frequencies  of  the  molecule.  The  use  of  group  theory  and  symmetry  

arguments  can  be  of  great  assistance  in  determining  of  which  vibrations  are  IR  or  

Raman  active  and  which  are  not.  Character  tables  for  the  various  symmetry  groups  can  

be  used  to  predict  how  many  of  the  3N-­‐6  vibrations  are  IR  active  and  how  many  are  

Raman  active.  In  particular,  if  the  symmetry  point  group  possesses  a  center  of  inversion,  

there  is  a  mutual  exclusion  rule  which  states  that  vibrations  allowed  in  the  Raman  

spectrum  are  forbidden  in  IR,  and  vice  versa.  The  symmetry  of  a  molecule  may  also  

dictate  that  certain  bands  are  forbidden  in  both  the  IR  and  Raman  spectra,  in  which  case  

even  the  combination  of  the  two  spectra  will  not  provide  the  full  information  about  the  

3N-­‐6  normal  frequencies.  

1.2.3.3 Anharmonicities  A  more  detailed  treatment  of  the  vibrational  spectra  of  molecules  can  be  made  if  

anharmonicity  is  included  into  consideration.  In  this  case  the  quadratic  model  for  PES  

function  is  extended  by  cubic  and  higher-­‐power  terms.22  The  occurance  of  

anharmonicities  has  several  important  consequences  for  vibrational  spectroscopy.    

1. First,  the  energy  levels  are  no  longer  equally  spaced.    

2. The   first   transition   n '' = 0! n ' = 1  (so-­‐called   fundamental   transition)  occurs  at   a  

slightly  lower  frequency  than  in  the  harmonic  case.  

3. The  vibrational  selection  rule   !n = ±1  breaks  down  and  leads  to  the  observation  of  

overtone   vibrations   in   the   experimental   spectra.   Because   the   energy   levels   are   no  

                                                                                                               22 Such anharmonicities will be studied in chapter Error! Reference source not found. for diatomic

molecules.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   23  

longer  equidistant,  the  positions  of  the  overtones  are  not  given  by  integer  multiples  

of  the  fundamental  frequency   n '' = 0! n ' = 1 .  

4. In   addition   to   overtones,   combination  bands  become  weakly   allowed   in  which   two  

different  vibrational  modes  are  simultaneously  excited.  

 Figure  9:  Description  of  overtone,  combination  and  “hot”  bands  in  IR  and  Raman  spectra.  

1.2.3.4 Isotope  Shifts  Since  the  vibrational  energies  depend  on  the  masses  of  the  vibrating  atoms  it  is  possible  

to  isotopically  label  the  compounds  of  interest  in  order  to  arrive  at  an  experimentally  

substantiated  assignment  of  the  experimental  spectra  of  in  order  to  assist  the  fitting  of  

the  force  fields  of  molecules.  In  the  case  of  a  diatomic  molecule,  the  isotrope  shift  is  

particularly  transparent  since  the  form  of  the  reduced  mass  is  simply  mAmB/(mA+mB)  

where  mA  and  mB  are  the  masses  of  the  two  vibrating  atoms  (see)  

 Figure  10:  The  reduced  mass  and  isotope  shifts  in  vibrational  spectra  of  diatomic  molecules.  

 

Computational  UV/vis,  IR  and  Raman  Spectroscopy   24  

1.2.4 Quantum  Chemical  Calculation  of  IR  and  Raman  Spectra  Within  the  harmonic  approximation  IR  and  Raman  vibrational  spectra  are  calculated  in  

the  course  of  frequency  calculations.  Since  harmonic  vibrational  frequencies  are  

determined  by  the  value  of  the  Hessian  matrix  at  the  ground-­‐state  equilibrium  

geometry,  it  is  necessary  that  frequencies  are  calculated  for  the  geometrical  structure  

optimized  at  the  same  level  of  theory  as  the  frequency  job.  ORCA  frequency  calculation  is  

requested  by  NumFreq  keyword  which  can  be  combined  with  keyword  Opt  in  order  to  

perform  geometry  optimization  prior  to  the  frequency  run.  The %freq  block  contains  

some  algorithm-­‐specific  information:  

 IR  spectral  intensities  are  calculated  automatically  in  frequency  runs.  Here  is  the  part  of  

the  output  of  a  frequency  calculation  containing  information  on  IR  intensities:  

 The  Mode  indicates  the  number  of  the  vibrational  modes,  then  the  vibrational  frequency  

follows.  The  value  T**2  is  the  square  of  the  change  of  the  dipole  moment  along  a  given  

vibrational  mode  in  KM/mole.  This  number  is  directly  proportional  to  the  intensity  of  a  

given  fundamental  in  an  IR  spectrum.    

Use  orca_mapspc program  to  plot  IR  spectrum:  

 The  produced  outputfile.ir.dat file  will  contain  the  IR  spectrum.  In  order  to  

predict  the  Raman  spectrum  of  a  compound  one  has  to  know  the  derivatives  of  the  

polarizability  with  respect  to  the  normal  modes.  Thus,  the  Raman  spectrum  will  be  

automatically  calculated  if  a  frequency  run  is  combined  with  a  polarizability  calculation  

which  is  requested  with  the  block:  

%freq CentralDiff true # use central-differences # (default is false) Increment 0.02 # increment in bohr for the # differentiation (default 0.005) end

----------- IR SPECTRUM ----------- Mode freq (cm**-1) T**2 TX TY TZ ------------------------------------------------------------------- 6: 1278.77 6.157280 ( -2.481387 -0.000010 -0.000287) 7: 1395.78 29.682711 ( -0.000003 -5.448182 -0.004815) 8: 1765.08 4.180497 ( 0.000537 -0.022003 2.044508) 9: 2100.53 8.550050 ( 0.000080 0.011990 2.924022) 10: 3498.72 1.843422 ( 0.001027 -0.178764 -1.345907) 11: 3646.23 19.778426 ( 0.000035 4.446922 -0.057577)

orca_mapspc outputfile ir –w20

Computational  UV/vis,  IR  and  Raman  Spectroscopy   25  

 The  output  consists  of  the  Raman  activity  (in  A4/AMU)  and  the  Raman  depolarization  ratios:    

 As  with  IR  spectra  you  can  get  a  plot  of  the  Raman  spectrum  using:  

 The  produced  outputfile.raman.dat file  will  contain  the  Raman  spectrum.  

1.3 Description  of  the  Experiment  

1.3.1 Calculation  of  UV/vis  spectra.    PART  1:  Calculate  the  absorption  spectra  of  the  protonated  and  unprotonated  forms  of  

the  dye  methyl  orange  (Figure  11).  Since  the  two  forms  show  distinct  absorption  

spectra,  such  dyes  can  be  used  as  pH  indicators.  The  protonated  form  (pH<3.1)  is  

characterized  by  intense  absorption  in  the  blue-­‐green  spectral  range  (450-­‐550  nm),  

while  the  absorption  of  the  deprotonated  form  (pH>4.4)  mainly  occurs  in  the  violet-­‐blue  

range  (400-­‐470  nm)  (Figure  11).23  Consequently,  upon  a  change  of  the  pH  from  3.1  to  

4.4,  methyl  orange  changes  its  color  from  red  to  yellow.  In  this  experiment  you  should  

try  to  understand  this  change.  

We  are  interested  in  the  calculation  of  the  most  intense  electronic  absorption  band  of  

both,  the  protonated  and  the  unprotonated  form  of  methyl  orange.  Below  we  specify  the  

input  files  which  contain  geometries.24  TD-­‐DFT  calculations  for  the  10  lowest  excited  

states  will  be  carried  out.  The  COSMO  approach  is  applied  in  order  to  model  the  solvent  

                                                                                                               23 Nero, J.D.; Araujo, R.E.; Gomes, A.S.L.; Melo, C.P.; (2005) J. Chem. Phys., 122, 104506.

24 Optimized with BP86/SV(P)

%elprop CalcPolar true end

-------------- RAMAN SPECTRUM -------------- Mode freq (cm**-1) Activity Depolarization -------------------------------------------------- 6: 1278.77 0.007349 0.749649 7: 1395.78 3.020010 0.749997 8: 1765.08 16.366586 0.708084 9: 2100.53 6.696490 0.075444 10: 3498.72 38.650431 0.186962 11: 3646.23 24.528483 0.748312

 

orca_mapspc outputfile raman –w20

Computational  UV/vis,  IR  and  Raman  Spectroscopy   26  

effect  on  the  calculated  transition  energies  of  the  species  in  aqueous  solution  (keyword  

cosmo(water)in  the  input  line).  We  use  the  SV(P)+    basis  set.25  

 

 

 Figure  11:  Cis  isomers  of  the  alkaline  and  monoprotonated  azonium  forms  of  methyl  orange  which  exist  in  alkaline  and  acidic  conditions,  respectively.  Experimental  UV/vis  spectra  of   (a)  aqueous  and  (b)  acetone  solutions  of  methyl  orange  for  different  values  of  the  pH.    

 

                                                                                                               25 This basis set, although being small, includes diffuse functions since they are important for the quality of excited-state calculations.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   27  

 

# # Calculate absorption spectrum of the cis isomer of the # monoprotonated azonium form of methyl orange molecule # ! RKS BP86 RI SV(P)+ TightScf cosmo(water) %basis Aux Auto end %tddft mode riints NRoots 10 MaxDim 200 triplets false end * int 0 1 S 0 0 0 0.000000 0.000 0.000 O 1 0 0 1.503294 0.000 0.000 O 1 2 0 1.505496 113.903 0.000 O 1 2 3 1.504846 113.963 132.114 C 1 2 3 1.821147 105.823 246.180 C 5 1 2 1.412488 118.923 184.915 H 6 5 1 1.101906 119.884 1.391 C 6 5 1 1.396656 120.278 181.201 H 8 6 5 1.099690 121.170 179.947 C 8 6 5 1.415430 119.104 359.943 C 5 1 2 1.405438 120.673 5.790 H 11 5 1 1.101105 119.344 358.753 C 11 5 1 1.401946 119.910 178.849 H 13 11 5 1.102862 120.413 180.017 N 10 8 6 1.407048 121.294 179.582 N 15 10 8 1.292068 122.718 359.407 C 16 15 10 1.355081 121.021 180.663 C 17 16 15 1.435712 115.336 179.613 H 18 17 16 1.101331 117.832 0.440 C 18 17 16 1.379608 121.689 180.593 H 20 18 17 1.098042 118.873 179.922 C 20 18 17 1.441237 120.588 0.036 C 17 16 15 1.438195 126.765 359.889 H 23 17 16 1.102101 121.369 359.401 C 23 17 16 1.379013 120.789 179.375 H 25 23 17 1.098225 118.466 180.009 N 22 20 18 1.355001 121.367 179.530 C 27 22 20 1.466602 120.443 359.964 H 28 27 22 1.112145 110.957 60.750 H 28 27 22 1.105488 109.004 180.193 H 28 27 22 1.112207 110.954 299.654 C 27 22 20 1.467211 120.863 180.183 H 32 27 22 1.111950 111.045 299.134 H 32 27 22 1.105431 108.919 179.800 H 32 27 22 1.111988 111.009 60.471 H 15 10 8 1.037594 115.838 179.885 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   28  

   

# # Calculate absorption spectrum of the cis isomer of the # alkaline form of methyl orange molecule # ! RKS BP86 RI SV(P)+ TightScf cosmo(water) %basis Aux Auto end %tddft mode riints NRoots 10 MaxDim 200 triplets false end * int -1 1 S 0 0 0 0.000000 0.000 0.000 O 1 0 0 1.505131 0.000 0.000 O 1 2 0 1.506757 113.600 0.000 O 1 2 3 1.506150 113.681 131.250 C 1 2 3 1.818920 106.015 245.700 C 5 1 2 1.413708 118.996 185.848 H 6 5 1 1.102898 119.621 1.627 C 6 5 1 1.396697 120.074 181.327 H 8 6 5 1.100787 121.258 179.965 C 8 6 5 1.422751 120.253 359.920 C 5 1 2 1.405535 120.698 6.867 H 11 5 1 1.101648 119.146 358.587 C 11 5 1 1.403465 119.673 178.764 H 13 11 5 1.102720 121.111 179.947 N 10 8 6 1.410630 124.803 179.669 N 15 10 8 1.282691 114.919 359.230 C 16 15 10 1.393811 116.204 180.367 C 17 16 15 1.420530 116.764 179.724 H 18 17 16 1.102756 118.044 0.310 C 18 17 16 1.392566 121.688 180.427 H 20 18 17 1.098954 118.761 179.958 C 20 18 17 1.434004 120.756 0.043 C 17 16 15 1.424999 125.381 0.000 H 23 17 16 1.101255 118.390 359.646 C 23 17 16 1.389114 121.131 179.553 H 25 23 17 1.099517 118.637 179.995 N 22 20 18 1.371544 121.464 179.666 C 27 22 20 1.460234 120.164 359.930 H 28 27 22 1.114577 111.584 60.709 H 28 27 22 1.106833 109.151 180.161 H 28 27 22 1.114654 111.583 299.634 C 27 22 20 1.460612 120.462 180.172 H 32 27 22 1.114413 111.662 299.198 H 32 27 22 1.106774 109.075 179.828 H 32 27 22 1.114474 111.629 60.462 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   29  

 • From   the   output,   determine   the   most   intense   transitions,   their   excitation  

energies  and   their  oscillator   strengths.  Are   the  variations  of   transition  energies  

consistent  with  experimentally  observed  spectral  changes  between  two  forms  of  

methyl  orange?  

• Analyze   the   electronic   structure   of   the   two   forms.  Determine   the   nature   of   the  

most   intense   transitions   by   inspecting   the   shapes   of   donor   and   acceptor   MOs  

with   a   visualization   program.   How   do   these   orbitals   change   upon   going   from  

unprotonated  to  protonated  species?  

• Use  the  orca_mapspc  program  to  plot  absorption  spectra  in  the  visible  spectral  

range  and  compare  them  with  the  experimental  spectra.    

PART  2:  Calculate  the  vibrational  structure  of  the  absorption  band  and  resonance  

Raman  spectra  corresponding  to  the  electronic  transition  11Ag  →11Bu  of  trans-­‐1,3,5-­‐

hexatrien.  Assuming  the  IHMDO  model,  the  calculation  of  vibronic  structure  in  

absorption  spectra26  for  dipole-­‐allowed  transitions  involves  two  stages.  First  one  

should  calculate  the  transition  energy,  transition  dipole  moment  and  origin  

displacement  of  the  excited  state  relative  to  the  ground  state  along  totally  symmetric  

normal  modes.  On  the  second  stage  the  calculated  parameters  are  employed  in  order  

to  simulate  the  absorption  spectrum  in  a  user-­‐specified  spectral  range.  Within  the  

harmonic  approximation,  the  origin  displacements  along  different  normal  modes  

may  be  approximated  by  means  of  excited-­‐state  energy-­‐gradient  calculations.  This  

type  of  job  is  specified  below  in  the  ORCA  input.  We  have  used  the  geometry  of  

hexatriene  optimized  at  the  RHF/SV(P)  level.  The  program  employs  the  Hessian  

matrix  obtained  from  a  corresponding  frequency  calculation  since  it  provides  

frequencies  and  normal  modes  which  are  used  in  the  transformation  of  the  energy  

gradient  from  Cartesian  to  normal  coordinates.  Thus,  the  name  of  the  Hessian  file  is  

specified  in  the  %rr  block  via  the  keyword  HessName.  

   

                                                                                                               26 as well as resonance Raman spectra; however, the resonance Raman technique will not be covered here.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   30  

 At  the  end  of  the  ORCA  run  you  get  the  file  hexatriene-uv.asa.inp  which  provides  

the  input  information  for  spectral  simulation.  The  simulation  part  is  performed  in  the  

framework  of  Heller  time-­‐dependent  theory  implemented  in  stand-­‐alone  program  

orca_asa.27  The  basic  structure  of  the  file  is  shown  below.  In  the  input  you  will  need  to  

modify  the  parameters  concerning  various  details  of  the  simulation  like  spectral  range  

and  number  of  points  for  absorption.  The  input  file  for  spectral  simulation  also  contains  

                                                                                                               27 The theory is explained in Petrenko, T.; Neese, F. (2007) Analysis and Prediction of Absorption Bandshapes, Fluorescence Bandshapes, Resonance Raman Intensities and Excitation Profiles using the Time Dependent Theory of Electronic Spectroscopy. J. Chem. Phys., 127, 164319; Neese, F.; Petrenko, T.; Ganyushin, D.; Olbrich, G. (2007) Advanced Aspects of ab initio Theoretical Spectroscopy of Open-Shell Transition Metal Ions. Coord. Chem. Rev., 205, 288-327

# hexatriene-uv.inp # # CIS Normal Mode Gradient Calculation # ! RHF TightSCF SV(P) Grid4 NoFinalGrid NMGrad %cis NRoots 1 EWin -10000, 10000 MaxDim 200 ETol 1e-7 RTol 1e-7 triplets false end %rr HessName= "hexatriene.hess" states 1 # perform energy-gradient calculation for the 1st #excited state Tdnc 0.005 # threshold for dimensionless displacement to be # included in the input file for spectra simulations # generated at the end of the program run ASAInput true # generate the input file for spectral simulations end * xyz 0 1 C -0.007965 0.666889 -0.000000 H -0.961692 1.187002 -0.000000 C 1.194639 1.503680 0.000000 H 2.146084 0.980277 0.000000 C 1.184814 2.831999 -0.000000 H 0.257404 3.396037 -0.000000 H 2.105024 3.404491 0.000000 C 0.007965 -0.666889 -0.000000 H 0.961692 -1.187002 -0.000000 C -1.194639 -1.503680 0.000000 H -2.146084 -0.980277 0.000000 C -1.184814 -2.831999 -0.000000 H -0.257404 -3.396037 -0.000000 H -2.105024 -3.404491 0.000000 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   31  

the  following  blocks  specifying  the  paramteres  for  the  IMDHO  model  which  were  

calculated  upon  ORCA  run:  

• %el_states   block   specifies   information   about   each   electronic   state   including  

adiabatic   minima   transition   energy,   electronic   transition   dipole   moment  

components,  homogeneous  linewidth  parameter  (Gamma),  standard  deviation  of  

transition  energy  (Sigma,  also  called  inhomogeneous  linewidth  parameter).  

• %vib_freq_gs   block   contains   ground-­‐state   vibrational   frequencies   of  

vibronically  active  modes.  

• %sdnc   block   specifies   dimensionless   normal   coordinate   displacements   for  

vibronically  active  modes.    

Computational  UV/vis,  IR  and  Raman  Spectroscopy   32  

   

# # hexatriene-uv.asa.inp # %sim model IMDHO method Heller # spectral range for simulation of absorption: AbsRange 40000.0, 55000.0 # number of points in the simulated absorption spectrum NAbsPoints 5000 end #--------------------------------------------------------------------------- # Transition Gamma Sigma Transition Dipole Moment (atomic unit) # Energy (cm**-1) (cm**-1) (cm**-1) Mx My Mz #--------------------------------------------------------------------------- $el_states 1 1 43422.32 50.00 0.00 -0.8533 -3.3690 -0.0000 $vib_freq_gs 13 1 373.151457 2 468.331475 3 1008.146916 4 1301.212433 5 1414.106060 6 1432.662207 7 1542.963264 8 1796.175363 9 1899.633103 10 3309.402656 11 3315.060790 12 3324.587826 13 3397.380154 $sdnc 13 1 1 1 -0.689568 2 -0.121420 3 0.332391 4 -0.998839 5 0.452707 6 0.186902 7 0.210313 8 -0.007187 9 -1.782913 10 0.034978 11 0.069401 12 0.031373 13 0.007086

Computational  UV/vis,  IR  and  Raman  Spectroscopy   33  

At  the  end  of  the  program  run  you  will  get  the  file  hexatriene-uv.abs.dat  

containing  the  absoption  spectrum.  

Now:  

• Plot   the   spectrum.   Determine   the   position   of   the   0-­‐0   vibronic   peak   in   the  

absorption.  Locate  the  vibronic  peak  with  the  maximum  intensity  and  compare  it  

with  the  calculated  vertical  transition  energy.  Explain  the  difference.    

• Identify   the   most   important   overtone   and   combination   transitions   in   the  

absorption   band.   How   do   their   intensities   correlate   with   the   values   of  

dimensionless  normal  coordinate  displacements  given  in  the  input?  

PART  3:  Calculate  CD  spectrum  of  the  (M)-­‐heptahelicene,  Figure  12).28  Expeimental  

transition  energies  and  intensities  of  the  lowest  energy  transitions  are  summerized  in  

Table  3.  

 

 Figure  12:  Structure  of  (M)-­‐pentahelicene  and  its  experimental  CD  spectrum.  

 

Table  3:  Experimental  results  for  the  lowest  excitations  of  (M)-­‐pentahelicene  in  ethanol  solution.  

State   !E (nm)   f   R  (10-­‐40  cgs)  1   392   0.002   0.1  2   328   0.115   -­‐150  3   306   0.400   -­‐400  

 

                                                                                                               28 Brown, A.; Kemp, C.M.; Mason, S.F.; (1971) J. Chem. Soc. A, 751.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   34  

Below  we  specify  the  input  file  which  contains  geometry  optimized  at  the  BP86/SV(P).  

TD-­‐DFT  calculations  for  the  10  lowest  excited  states  will  be  carried  out.  The  COSMO  

approach  is  applied  in  order  to  model  the  solvent  effect  on  the  calculated  transition  

energies  and  intensities.  

# # Calculate CD spectrum of (M)-pentahelicene # ! RKS BP86 RI SV(P)+ cosmo(ethanol) %basis aux auto end %tddft mode riints NRoots 10 MaxDim 300 triplets false end * xyz 0 1 C -0.032098 0.727529 0.865728 H 0.368360 -1.199356 4.290991 C 0.032098 -0.727529 0.865728 C 0.292271 -1.387042 2.114305 C -0.292271 1.387042 2.114306 C -0.954485 -1.133225 -1.443248 C -0.191948 0.661637 3.344271 C -0.616376 2.787102 2.138058 C 0.191948 -0.661637 3.344271 H -0.368360 1.199356 4.290991 C -0.242892 -1.583277 -0.290736 C 0.616376 -2.787102 2.138058 C -0.589489 3.546253 0.992284 H -0.890736 3.240068 3.105649 C -0.095388 2.985042 -0.233202 H -0.868017 4.613230 1.014928 C 0.242892 1.583277 -0.290736 C 0.147631 3.819306 -1.363589 C -0.147631 -3.819306 -1.363589 C 0.954486 1.133225 -1.443248 C 0.589489 -3.546253 0.992284 H 0.890735 -3.240068 3.105649 C 0.095388 -2.985042 -0.233202 H 0.868017 -4.613230 1.014928 C 0.784716 3.327293 -2.497196 H -0.149295 4.880184 -1.305305 C 1.221968 1.978849 -2.516823 H 0.982510 3.987823 -3.357927 C -1.221968 -1.978849 -2.516823 H 1.787046 1.595442 -3.382869 H 0.149294 -4.880184 -1.305305 C -0.784716 -3.327293 -2.497196 H -0.982510 -3.987823 -3.357927 H -1.787046 -1.595442 -3.382869 H -1.326263 -0.099422 -1.475319 H 1.326263 0.099422 -1.475319 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   35  

Now:  • From   the   output,   determine   the   most   intense   transitions,   their   excitation  

energies,  oscillator  strengths,  and  rotatory  strengths.  Perform  the  assignement  of  

the  experimentally  observed  optical  transitions.  29  

• Determine  the  nature  of  the  most  intense  transitions  by  inspecting  the  shapes  of  

donor  and  acceptor  MOs  with  a  visualization  program.    

• Use   the   orca_mapspc   program   to   plot   CD   spectrum   in   appropriate   spectral  

range  and  compare  it  with  the  experimental  spectrum.  

 

PART  4:  Calculate  the  absorption  and  CD  spectra  corresponding  to  the  lowest  energy  

transition  in  the  1,7,7-­‐trimethyl-­‐bicyclo[2.2.1]heptan-­‐2-­‐one    (d-­‐(+)-­‐camphor,  Figure  13  

).30    

 

 Figure   13:   Strucure   of   d-­‐(+)-­‐camphor   and   its   experimental   UV   absorption   (dashed   line)   and   CD   (solid   line)   spectra   in  ethanol  solution  corresponding  to  the  lowest  energy  electronic  band.  

 We  are  interested  in  the  calculation  of  the  low-­‐energy  electronic  absorption  and  CD  

band  of  the  compound.  Below  we  specify  the  input  file  which  contains  geometry  

optimized  at  the  BP86/SV(P).  TD-­‐DFT  calculations  for  the  3  lowest  excited  states  will  be  

carried  out.  The  COSMO  approach  is  applied  in  order  to  model  the  solvent  effect  on  the  

calculated  transition  energies  and  intensities.  

 

 

                                                                                                               29 For a discussion see also Furche, F.; Ahlrichs, R; Wachsmann, C.; Weber, E.; Sobanski, A.; Vögtle, F.; Grimme, S.; (2000) J. Am. Chem. Soc., 122, 1717.

30 Gillard, R.D.; Mitchell, P.R.; (1969) Trans. Fraraday. Soc., 65, 2611.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   36  

   

# # Calculate the lowest energy optical transitions # in d-(+)-camphor # ! RKS BP86 RI SV(P)+ cosmo(ethanol) %basis Aux auto end %tddft tda true mode riints NRoots 3 MaxDim 300 triplets false end * xyz 0 1 O 1.877429 6.781782 7.163138 C 1.373114 7.681449 4.918091 C 1.308771 7.598232 6.450938 C 0.405529 8.751374 6.913391 C 0.099863 9.480852 5.585535 H -0.775079 10.160694 5.637362 C 1.406490 10.155015 5.090954 H 1.881650 10.771421 5.884999 H 1.199121 10.824910 4.228371 C 2.287481 8.937404 4.662510 H 3.248107 8.870379 5.216696 H 2.533414 8.970132 3.578223 C -0.026564 8.284845 4.609006 H -0.862265 7.593915 4.861784 H -0.124454 8.594531 3.544670 C 1.813380 6.399538 4.228001 H 1.886092 6.541741 3.126493 H 1.097222 5.570740 4.426288 H 2.810451 6.077492 4.602254 H -0.517516 8.317884 7.364558 H 0.900650 9.367303 7.696537 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   37  

Now:  

• From  the  output,  determine  the  transitions,  which  energy  most  closely  matches  

the   experimental   one.  Determine   the   nature   of   the   transition   by   inspecting   the  

shapes  of  donor  and  acceptor  MOs  with  a  visualization  program.    

• Use  the  orca_mapspc  program  to  plot  absorption  and  CD  spectra  in  the  spectral  

range  around  300  nm  using  appropriate  value  of   the  bandwidth  which   roughly  

matches  the  experimental  one.  Compare  experimental  and  theoretical  plots.  

• Does  the  rotatory  strength  have  the  proper  sign  for  the  given  transitions?  What  

parameter,   that   can  be   rather   sensitive   to   the   level   of   calculation,   can   result   in  

observed   disagreement?   As   a   hint,   consider   the   the   scalar   product   of   the  

transition   electric   dipole   and   magnetic   dipole   moments   in   the   form  

D0!I"M

0!I= D

0!IM

0!Icos! .  

1.3.2 Calculation  of  IR  and  Raman  spectra  PART  1:  Optimize  the  geometries,  calculate  vibrational  frequencies,  IR  and  Raman  

intensities  of  the  following  diatomic  molecules  using  the  BP86  functional  and  the  TZVP  

basis  set:  

• CO   :  ωexp=  2170  cm-­‐1  • HF   :  ωexp=  4138  cm-­‐1  • ClF   :  ωexp=    786  cm-­‐1  • N2   :  ωexp=  2359  cm-­‐1  • HCl   :  ωexp=  2991  cm-­‐1  • Cl2   :  ωexp=    560  cm-­‐1    Now:  • Analyze  the  origin  of  the  trends  in  the  calculated  IR  intensities  using  population  

analysis  and  chemical  intuition    

• Why  do  the  stretching  vibrations  of  the  homonuclear  diatomics  do  not  show  any  

IR  intensity?  Provide  a  qualitative  explanation  on  the  origin  of  IR  intensity  for  the  

heteronuclear  species.  

• Compare   the   calculated   harmonic   frequencies   with   the   experimental   harmonic  

frequencies.  How  reliable  are  the  DFT  results?31    

                                                                                                               31 In general, the calculations always produce harmonic force constants and therefore also harmonic frequencies. The underlying assumption is that the potential energy surfaces behave exactly quadratically. In reality, however, the potentials are anharmonic and this leads to important alternations in the spacing of the vibrational levels. Some of these aspects will be studied for diatomic molecules in section Error! Reference source not found. on page 166. However, experimental harmonic frequencies are only known for very small molecules and anharmonic frequencies are difficult to calculate. In practice, this limits the accuracy of the

Computational  UV/vis,  IR  and  Raman  Spectroscopy   38  

 

PART  2:  Optimize  the  geometry,  calculate  vibrational  frequencies,  IR  and  Raman  

intensities  of  the  benzene  molecule  using  the  BP86  functional  and  the  TZVP  basis  set.  

Experimental  vibrational  frequencies  of  the  benzene  are  known  from  IR  and  Raman  

measurements:  

• IR   :  1485  cm-­‐1  • Raman   :  605.0,  991.6,  1178.0,  1595.0  cm-­‐1    

Here  is  the  input  file  for  ORCA  calculation:  

   

Now:  • Assign  experimental  vibrational   frequencies.  Determine   the   character  of   IR  and  

Raman  active  vibrations  using  the  gOpenMol  program  for  visualization.  

• Plot  the  experimental  versus  the  calculated  frequencies.  

• Calculate   the   parameters   of   a   linear   regression   analysis.  How   reliable   are   your  

predictions   of   vibrational   frequencies?   What   is   your   mean   deviation   from  

experiment,  what  is  your  maximum  deviation?  

• The   benzene   molecule   possesses   a   center   of   inversion.   Show   complementary  

nature  of  Raman  and  IR  spectra  on  the  basis  of  the  calculated  IR  intensities  and  

                                                                                                                                                                                                                                                                                                                                                         comparison between theory and experiment and one usually compares calculated harmonic frequencies with observed fundamentals (which contain anharmonic contributions).

! RKS BP86 RI TZVP TZV/J TightOpt TightScf Grid4 NoFinalGrid NumFreq %freq CentralDiff true Increment 0.02 end %elprop Polar true end # vibrational analysis of benzene molecule * xyz 0 1 C 0.000000 -0.7000000000 1.2124355653 H 0.000000 -1.2500000000 2.1650635095 C 0.000000 -1.4000000000 0.0000000000 H 0.000000 -2.5000000000 0.0000000000 C 0.000000 -0.7000000000 -1.2124355653 H 0.000000 -1.2500000000 -2.1650635095 C 0.000000 0.7000000000 1.2124355653 H 0.000000 1.2500000000 2.1650635095 C 0.000000 1.4000000000 0.0000000000 H 0.000000 2.5000000000 0.0000000000 C 0.000000 0.7000000000 -1.2124355653 H 0.000000 1.2500000000 -2.1650635095 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   39  

Raman  activities.  Which  symmetry  species  are  active  in  IR  and  which  in  Raman?  

Which  vibrations  are  forbidden  in  both  IR  and  Raman  spectra?  

 

PART  3:  Optimize  the  geometries  and  calculate  vibrational  frequencies  of  the  following  

molecules  using  the  BP86  functional  and  the  TZVP  basis  set:  

• C2H2    • C2H4  • C2H6    JOB:  

• Discuss  the  trend  in  the  variation  of  frequency  of  CC  stretching  mode  between  the  

molecules.   How   do   the   frequencies   depend   on   the   bond   order?   Compare   the  

results  with  the  characteristic  frequencies  known  for  C–C,  C=C  and  C≡C  bonds.  

• Use   the   orca_vib   program   to   obtain   estimates   of   the   C-­‐C   stretching   force  

constants.  How  do  they  vary  with  bond  order?    

• Provide  a  qualitative  explanation  of  the  observed  trend  on  the  basis  of  chemical  

intuition.  

PART  4:  Optimize  the  geometry,  calculate  vibrational  frequencies,  IR  and  Raman  

intensities  of  the  glycine  molecule  (H2N-­‐CH2-­‐COOH)  using  the  BP86  functional  and  the  

TZVP  basis  set.  

The  input  file  is  specified  below:  

 The  experimental  IR  and  Raman  powder  spectra  of  glycine32  are  given  in  Figure  14.  

                                                                                                               32 Kumar, S.; Rai, A.K.; Singh, V.B.; Rai, S.B.; (2005) Spectrochim. Acta Part A, 61, 2741-2746.

! RKS BP86 RI TZVP TZV/J TightOpt TightScf Grid4 NoFinalGrid NumFreq %freq CentralDiff true Increment 0.02 end %elprop Polar true end * xyz 0 1 N 0.417502 -1.939309 0.000000 C -0.563503 -0.866622 0.000000 C 0.000000 0.555853 0.000000 H 1.029096 -1.853448 0.815576 H 1.029096 -1.853448 -0.815576 H -1.225791 -0.963395 -0.876292 H -1.225791 -0.963395 0.876292 O 1.179353 0.850856 0.000000 O -1.002263 1.487634 0.000000 H -0.564825 2.365543 0.000000 *

Computational  UV/vis,  IR  and  Raman  Spectroscopy   40  

 Figure  14:  Experimental  infrared  (a)  and  Raman  (b)  powder  spectra  of  glycine.  

The  experimental  vibrational  frequencies  of  were  determined  as  follows:  

• IR   :  504,  584,  607,  698,  893,  1034,  1334,  1410,  1504,  1610,            1703,  2128,  2920,  3084,  3414  cm-­‐1  

• Raman  :  497,  602,  697,  893,  1033,  1323,  1410,  1508,  1667,  2123,              2930,  3050  cm-­‐1  

Now:  • Plot  calculated  IR  and  Raman  spectra  using  the  orca_mapspc  program.  

• Assign  experimental  vibrational   frequencies.  Determine   the   character  of   IR  and  

Raman   active   vibrations   using   the   gOpenMol   program   for   visualization.   Find  

corresponding   vibrations   which   are   clearly   identified   in   both   IR   and   Raman  

spectra.33  Which  normal  modes  have  noticeable  intensities  only  in  IR  or  Raman?  

• Plot  the  experimental  versus  the  calculated  frequencies.  

• Calculate   the   parameters   of   a   linear   regression   analysis.  How   reliable   are   your  

predictions   of   vibrational   frequencies?   What   is   your   mean   deviation   from  

experiment,  what  is  your  maximum  deviation?  

 

                                                                                                               33 Some of the vibrations may show slightly different frequencies in IR and Raman measurements which is due to slight variations in the wavelength calibration and the experimental conditions.

Computational  UV/vis,  IR  and  Raman  Spectroscopy   41