• Keine Ergebnisse gefunden

2.   Molecular  spectroscopy:  theoretical  background  and  datasets

2.1.   Levels  and  transitions

Electromagnetic  radiation  can  be  characterised  either  in  terms  of  the  wavelength   or  frequency.  Conventionally  the  wavelength  λ  [nm/µm]  is  used  in  the  UV-­‐VIS-­‐

NIR,   the   wavenumber  𝜈𝜈  [cm-­‐1]   in   the   NIR-­‐MIR-­‐FIR,   and   the   frequency   ν   [MHz/GHz]  in  the  microwave  regions,  Figure  2.1.  These  units  are  related  as:  

𝜈𝜈 = 𝜈𝜈 𝑐𝑐  = 1

𝜆𝜆  

Figure  2.1.  Electromagnetic  spectrum    

According  to  the  quantum  mechanics,  atoms  and  molecules  can  exist  in  certain   discrete   states   that   are   different   with   respect   to   the   electron   configuration,   angular   momentum,   parity   and   energy.   Transitions   between   different   energy   states   are   associated   with   the   absorption   and   emission   of   the   electromagnetic   radiation  by  the  molecule.  The  actually  observed  resulting  spectrum  may  consist   of   both   individual   lines   corresponding   to   specific   transitions   and   continuous   bands,  depending  on  the  density  of  transitions  on  wavelength  scale  and  typical   Doppler  widths.  

The   use   of   the   word   “line”   to   describe   an   experimentally   observed   transition   goes  back  to  the  early  days  of  observations  of  visible  spectra  with  spectroscopes.  

In   those   instruments,   the   lines   observed   in,   for   example,   the   spectrum   of   a   sodium   flame   are   images,   formed   at   various   wavelengths,   of   the   entrance   slit.  

Although  modern  observations  are  usually  in  the  form  of  a  plot  of  some  measure   of   the   intensity   of   the   transition   vs.   wavelength,   frequency   or   wavenumber,   peaks  in  such  a  spectrum  are  referred  to  as  “lines”.  

Using   somewhat   simplified   approach,   the   full   energy   state   a   molecule   can   be   divided  into  electronic,  rotational,  and  vibrational  constituents:  

E  =  Evib  +  Erot  +  Eelec           (2.1)   The  simplest  way  describe  the  vibrational  energy  state  is  to  consider  an  example   of  a  diatomic  molecule.  The  bond  between  the  two  atoms  in  a  diatomic  molecule   can  be  viewed  as  a  spring,  which,  as  the  internuclear  distance  r  changes  from  the   equilibrium  value  re,  exerts  a  force  that  tends  to  restore  atoms  to  the  equilibrium   position.   Such   a   system,   assuming   the   parabolic   shape   of   the   potential   energy   curve,  is  approximated  by  a  harmonic  oscillator.  For  the  vibrational  motion  of  a   harmonic  oscillator,  the  vibrational  energy  term  is  given  by:  

Evib  =  ħω  (v  +!!)           (2.2)   where  ħ  =  !!!,  h  is  the  Plank  constant  h=6.626×10-­‐34  J s,  ω  is  the  angular  frequency   (   ω   =  2𝜋𝜋∙  ν   ),   v   is   a   vibrational   quantum   number   (v   =   0,1,2,…),   so   that   energy   levels  of  the  ideal  oscillator  are  equally  spaced  (Figure  2.2  a-­‐b).

   

Figure   2.2.  (a)  Vibration  of  heteronuclear  diatomic  molecule  (HCl)  (b)  potential  energy  of  an  ideal   harmonic  oscillator,  and  (c)  an  anharmonic  oscillator  described  by  the  Morse  function  (d)  

This  idealized  approximation  is  valid  for  low  vibrational  energy  levels  only.  For   real  molecules,  the  potential  energy  rises  sharply  at  small  values  of  r,  whereas   for   large  r   the   bond   stretches   until   it   ultimately   breaks   and   the   dissociation   occurs.   The   real   potential   is   not   symmetric   (Figure   2.2c,d).   The   energy   of   vibrational   transition   is   on   the   order   of   0.1   eV,   with   the   corresponding   wavelengths   in   the   IR   spectral   range.   Because   the   vibrational   energy   level   spacing   is   relatively   large   (typically   of   the   order   of   103   cm-­‐1)   compared   to   the   thermal   energy,   most   molecules   at   room   temperature   are   in   their   lowest   vibrational  energy  level  and  light  absorption  normally  occurs  from  v  =  0.    

A  selection  rule  (transition  rule)  formally  constrains  the  possible  transitions  of  a   system   from   one   state   to   another.   Selection   rules   have   been   derived   for   electronic,  vibrational,  and  rotational  transitions.  The  strength  or  energy  of  the   interaction   between   a   charge   distribution   and   an   electric   field   depends   on   the   dipole  moment  of  the  charge  distribution  in  a  molecule.  

To  obtain  the  strength  of  the  interaction  that  causes  transitions  between  states   (each  characterized  by  its  wave  function),  the  transition  dipole  moment  integral   is  used  rather  than  the  dipole  moment.  If  transition  dipole  moment  integral  is   zero,  then  the  interaction  energy  is  zero  and  no  transition  occurs  or  is  possible   between  the  two  states.  Such  a  transition  is  said  to  be  forbidden  (electric-­‐dipole  

forbidden).  If  transition  dipole  moment  integral  is  large,  then  the  probability  of  a   transition  is  large  /Herzberg,  1988/.  

For  a  purely  vibrational  transition,  the  selection  rule  requires  the  change  of  the   dipole  moment  during  the  vibration.  This  oscillating  dipole  moment  produces  an   electric  field  that  can  interact  with  the  oscillating  electric  and  magnetic  fields  of   the   electromagnetic   radiation.   Heteronuclear   diatomic   molecules   such   as   NO,   HCl,   and   CO   absorb   infrared   radiation   and   undergo   vibrational   transitions,   contrary   to   homonuclear   diatomic   molecules   such   as   N2   and   02,   whose   dipole   moments  remain  constant  during  vibration.  

Thus,  the  vibrational  transitions  occur  for  ∆v  =  ±1,  ±2,  ±3,  …  .  For  the  harmonic   oscillator,  only  the  transitions  with  ∆v  =  ±1  are  allowed  (fundamental  modes).  

The   anharmonicity   of   the   real   molecules   leads   to   weaker   overtone   transitions   with   ∆v   =   ±2,   ±3,   ±4,   …   .   For   linear   diatomic   molecules   only   one   mode   of   vibration  is  possible  (stretching  of  the  bond  length).  For  polyatomic  molecules,   several  modes  of  vibration  are  possible,  resulting  in  changes  of  the  bond  lengths   and  angles  between  the  bonds  (Figure  2.3a)    

   

Figure  2.3.  (a)  Internal  vibrations  of  the  bonds  in  the  H2O  molecule,  (b)  rotational  motion  of  H2O,  and   (c)  translation  of  the  H2O  molecule.

 

For  a  given  electronic  configuration  and  a  given  vibrational  level  (v  value),  it  is   also   possible   to   discriminate   the   rotation   of   the   molecule   with   respect   to   particular  axis  (Figure  2.3b).  The  rotational  energy  is  given  as:    

Erot  =  B∙J∙  (J  +  1)           (2.3)   B= ħ2/2Θ  

where  B  is  the  rotational  constant  of  a  molecule  (assumed  to  be  a  rigid  rotator   with  constant  internuclear  distances)  for  a  particular  rotation  mode,  connected  

with  the  moment  of  inertia  Θ  for  the  given  rotational  axis,  and  J  is  the  rotational   quantum   number   (J   =   0,   1,   2,…).   Again,   this   approximation   is   limited   to   small   values   of  J,   since   bond   lengths   of   real   polyatomic   molecules   are   subject   to   stretching   due   to   the   centrifugal   force,   resulting   in   increase   of  Θ   for   higher  J   values.  

The   difference   in   the   angular   momentum   quantum   number   of   initial   and   final   states  is  given  as  ∆J  =±1,  since  the  photon  exchanged  with  the  atom  or  molecule   has  a  spin  of  unity.  The  rotational  transition  can  be  observed  when  ∆J  =  0,  ±1.  

The  transitions  are  denoted  as  P-­‐branch  (∆J  =  -­‐1),  and  R-­‐branch  (∆J  =  +1).  For  Q-­‐

branch  transitions,  observed  simultaneously  with  an  electronic  transition,  ∆J  =  0.    

The  energy  difference  between  two  consecutive  rotational  states  (∆J  =+1)  is:  

 

∆Erot  =  EJ+1  -­‐  EJ  =  B[  (J+1)(J+2)  -­‐  J(J+1)  ]  =  2B(J+1)    

The   spacing   between   rotational   energy   levels   increases   with  J.   Energy   level   spacings  are  small  compared  to  those  of  vibrational  transitions,  typically  of  the   order  of  10  cm-­‐1  (10-­‐3  eV)  in  the  lower  levels;  this  corresponds  to  absorption  in   the  microwave  region.  The  resulting  observed  sequence  of  rotational  transitions   (called  “band”)  consists  of  a  series  of  equally  spaced  lines.  

At   room   temperature   conditions,   the   thermal   energy   available   is   sufficient   to   populate  the  energy  levels  above  J  =  0.  

The  Eelec  component  characterizes  the  energy  of  excited  electronic  configurations   of  the  molecule.  Energy  difference  ∆E  associated  with  electronic  transitions  is  of   the  order  of  1  eV,  corresponding  to  wavelengths  in  the  visible  or  UV  parts  of  the   spectrum.  The   change   in   the   electronic   energy   may   occur   simultaneously   with   the   changes   in   vibrational   and   rotational   energy   (molecular   bond   length   and   interaction   potential   changes   with   the   electronic   configuration).   As   a   result,   observed   electronic   transitions   may   have   a   vibrational   structure   with   a   rotational  “fine  structure”.    

Electronic  states  notation  arises  from  the  symmetry  group  theory,  and  involves   consideration   of   the   angular   momentum   of   the   molecules   (electronic   and   rotational)   and   their   electronic   spin.   The   detailed   explanation   of   concept   is   available  in  classical  monograph  by  /Herzberg,  1988/.    

The  energy  levels  are  thus  described  by  a  sum  of  the  corresponding  terms:    

E  =  Eelec  +  ħω  (  v  +  !!  )  +  B⋅J⋅(  J+1)       (2.4)   Real   molecular   spectra   have   a   complex   structure   and   involve   electronic,   vibrational-­‐rotational  (also  called  ro-­‐vibrational)  and  pure  rotational  transitions   at   typical   ambient   temperatures.   The   simple   example   for   transitions   between   two   vibrational   levels   v´´=0   and   v´=1   belonging   to   the   same   electronic   energy   level  and  separated  in  the  rotational  levels  J´´  and    is  shown  in  Figure  2.4.    

The  photon  energy  of  allowed  transitions  is  given  by  the  difference  of  energy  of   two  consecutive  states:  

∆𝐸𝐸= 𝐸𝐸(v′,𝐽𝐽′)  −𝐸𝐸(v′′,𝐽𝐽′′)=  

=ℏ𝜔𝜔 v!+!! +𝐵𝐵!∙𝐽𝐽!∙ 𝐽𝐽! +1 −ℏ𝜔𝜔 v!!+!! +𝐵𝐵!!∙𝐽𝐽!!∙ 𝐽𝐽′!+1     (2.6)   Assuming  B’  ≈ B’’  =  B,  Δ  v = +1  

∆𝐸𝐸= ℏ!! +𝐵𝐵∙ 𝐽𝐽!∙ 𝐽𝐽!+1 −𝐽𝐽!!∙ 𝐽𝐽′!+1     (2.6a)   and  for  P,  R  and  Q  branches  

∆J  =  +1,     J’=J’’+1,    ∆𝐸𝐸≈ ℎ𝜈𝜈!+2𝐵𝐵∙ 𝐽𝐽′!+1  

J  =  -­‐1,     J’=J’’-­‐1,      ∆𝐸𝐸≈ ℎ𝜈𝜈!−2𝐵𝐵∙𝐽𝐽!!         (2.6b)  

∆J  =  0,       J’=J’,      ∆𝐸𝐸= ℎ𝜈𝜈!+(𝐵𝐵! −𝐵𝐵′′)𝐽𝐽!∙ 𝐽𝐽! +1      

Schematically,   P,   Q   and   R   branches   are   shown   in   Figure   2.4.   Since   difference   between   rotational   constants   is   not   very   big,   components   of   the   Q-­‐branch   spectral  lines  are  closely  spaced.   The  Frank-­‐Condon  principle  defines  intensity   distribution  in  the  branches  /Herzberg,  1988/.    

 

Figure  2.4.  An  energy  level  diagram  showing  some  of  the  transitions  involved  in  the  IR  ro-­‐vibrational   spectrum  of  a  linear  molecule  

Resolved  ro-­‐vibrational  molecular  spectra  are  of  particular  interest  for  remote   sensing  applications  because  they  allow  to  distinguish  constituents  due  to  their   characteristic  spectral  features.    

Homonuclear  molecules  (like  N2  and  O2)  do  not  have  strong  absorption  features   in  the  thermal  infrared,  making  the  atmosphere  virtually  transparent  for  sunlight  

in   this   region.   In   contrast,   heteronuclear   diatomic   molecules   and   most   polyatomic  molecules  have  strong  resolved  rotational  spectral  features  in  the  IR,   which  makes  this  range  useful  for  spectroscopic  observations.  

The  principles  considered  above  for  diatomic  molecules  are  generally  valid  for   polyatomic  molecules,  but  their  spectra  are  far  more  complicated.    

Nonlinear   polyatomic   molecules   consisting   of   n   atoms   have   three   rotational   constants   with   respect   to   the   principal   axes   and   3n-­‐6   vibrational   degrees   of   freedom   (3n-­‐5   for   linear   polyatomic   molecules).   The   resulting   spectrum   may   contain   multiple   absorption   bands,   as   well   as   overtone   bands   (∆v   >   1)   and   combination  bands  (absorptions  corresponding  to  the  sum  of  two  or  more  of  the   fundamental  vibrations).