orca labs v thermochemistry ts - university of...

19
Computational Thermochemistry 1 1 Computer Experiment 5: Computational Thermochemistry 1.1 Background: Within the frame of this experiment you will employ methods of statistical thermodynamics in order to derive thermochemical reaction quantities and energies of chemical bonds, which both are well known subjects to the experimental chemist. It will become evident that frequency analysis is an essential tool in theoretical thermochemistry. 1.1.1 Basics of Statistical Thermodynamics The link between (microscopic) molecular quantities and the (macroscopic) observables of a chemical system is provided by the system’s partition function Q, which is based on the Boltzmann distribution over the available microstates and in the general formulation reads: Q = g j ! e " E j k B ! T j =1 N # ( 1) Here N denotes the total number of accessible states, Ej the energy of the jth state, kB the Boltzmann constant and T the Kelvin temperature (the term kBT can be interpreted as the available thermal energy). The degeneracy factor gj is directly given by the degeneracy of the jth state and equals one if the corresponding state is not degenerated. Consider a homogenous chemical system consisting of molecules, e.g. a monomolecular gas. From the macroscopic perspective, the system’s components are the molecules and the system’s energy states Ej depend on the arrangement of the components and their energies. The corresponding partition function is referred to as the canonical or system partition function Q, whereas the microcanonical or molecular partition function q treats the single molecule as the superordinated system, thus being a sum over the molecular energy states. In first approximation, the total energy of a molecule can be split up into contributions from each of the molecule’s degree of freedom (which is a direct result from the BornOppenheimer approximation): ! = ! i trans + ! j rot + ! k vib + ! l el . ( 2) Thus the molecular partition can be factorised according to: q = q trans ! q rot ! q vib ! q el . ( 3)

Upload: others

Post on 10-Mar-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     1  

1 Computer  Experiment  5:  Computational  Thermochemistry  

1.1 Background:  Within  the  frame  of  this  experiment  you  will  employ  methods  of  statistical  

thermodynamics  in  order  to  derive  thermochemical  reaction  quantities  and  energies  of  

chemical  bonds,  which  both  are  well  known  subjects  to  the  experimental  chemist.  It  will  

become  evident  that  frequency  analysis  is  an  essential  tool  in  theoretical  

thermochemistry.  

1.1.1 Basics  of  Statistical  Thermodynamics  The  link  between  (microscopic)  molecular  quantities  and  the  (macroscopic)  observables  

of  a  chemical  system  is  provided  by  the  system’s  partition  function  Q,  which  is  based  on  

the  Boltzmann  distribution  over  the  available  microstates  and  in  the  general  formulation  

reads:  

  Q = g

j!e"

Ej

kB !T

j=1

N

#                   (  1)  

Here  N  denotes  the  total  number  of  accessible  states,  Ej  the  energy  of  the  jth  state,  kB  the  

Boltzmann  constant  and  T  the  Kelvin  temperature  (the  term  kBT  can  be  interpreted  as  

the  available  thermal  energy).  The  degeneracy  factor  gj  is  directly  given  by  the  

degeneracy  of  the  jth  state  and  equals  one  if  the  corresponding  state  is  not  degenerated.    

Consider  a  homogenous  chemical  system  consisting  of  molecules,  e.g.  a  monomolecular  

gas.  From  the  macroscopic  perspective,  the  system’s  components  are  the  molecules  and  

the  system’s  energy  states  Ej  depend  on  the  arrangement  of  the  components  and  their  

energies.  The  corresponding  partition  function  is  referred  to  as  the  canonical  or  system  

partition  function  Q,  whereas  the  microcanonical  or  molecular  partition  function  q  treats  

the  single  molecule  as  the  superordinated  system,  thus  being  a  sum  over  the  molecular  

energy  states.  In  first  approximation,  the  total  energy  of  a  molecule  can  be  split  up  into  

contributions  from  each  of  the  molecule’s  degree  of  freedom  (which  is  a  direct  result  

from  the  Born-­‐Oppenheimer  approximation):  

  ! = !

itrans + !

jrot + !

kvib + !

lel    .                 (  2)  

Thus  the  molecular  partition  can  be  factorised  according  to:  

  q = qtrans!q

rot!q

vib!q

el    .                 (  3)  

Page 2: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     2  

With  this  result  the  actual  calculation  of  the  molecular  partition  function  is  possible  by  

applying  the  quantum  chemical  solutions  of  appropriate  problems,  as  for  example  the  

particle  in  a  box  or  the  harmonic  oscillator,  to  the  degree-­‐of-­‐freedom  dependent  

expressions  of  the  molecular  partition  function.  For  example,  the  application  of  the  

harmonic  oscillator  energy  eigenvalues  to  the  vibrational  partition  function  of  a  

diatomic  molecule  yields  the  expression:  

 

qvib

=e!

h!2kBT

1!e!

h!kBT

     .                   (  4)  

Thus  the  only  quantity  which  has  to  be  known  is  the  frequency  of  the  system’s  normal  

mode.  Similar  expressions  for  the  other  degree-­‐of-­‐freedom  dependent  partition  

functions  can  be  derived,  depending  on  molecular  quantities  like  the  moment  of  inertia  

or  the  volume,  which  in  first  approximation  can  be  determined  according  to  the  ideal  gas  

law.  As  an  exception  to  this,  the  electronic  partition  function  includes  a  summation  over  

the  system’s  quantum  states,  which  are  given  by  solving  the  electronic  Schrödinger  

equation:  

  q

el= g

l!e"!lel

kB !T

l=1

N

#                   (  5)  

  In  most  cases,  the  excited  electronic  states  cannot  be  populated  by  the  available  

thermal  energy  and  thus  are  left  out  of  the  summation.  A  further  simplification  is  

realized  by  setting  the  zero  point  of  energy  to  the  electronic  energy  of  the  reactant’s  

ground  state,  in  which  case  the  partition  function  is  reduced  to  the  degeneracy  factor  g1.  

As  laid  out  above,  the  partition  function  contains  all  necessary  information  for  the  

determination  of  the  system’s  thermochemical  quantities.  For  example,  the  general  

expression  of  the  mean  energy  is  proportional  to  the  derivation  of  the  partition  

function’s  natural  logarithm  with  respect  to  the  temperature:  

  E =!

"lnQ"ß

#

$%%%%

&

'((((

= kBT 2 )

"lnQ"T

#

$%%%%

&

'((((    .               (  6)  

In  the  thermodynamic  equilibrium,  this  formulation  equals  the  inner  energy  U  of  the  

system  if  the  volume  remains  constant  and  the  inner  energy  at  0  K  is  zero.  Since  the  

inner  energy  is  the  basic  quantity  in  equilibrium  thermodynamics,  expressions  for  all  

other  thermochemical  state  functions  can  be  derived  on  the  basis  of  the  system’s  

Page 3: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     3  

partition  function  (some  authors  even  say  that  the  partition  function  is  the  “wave  

function  of  modern  thermodynamics”).  

 

 

1.1.2 Thermochemical  Quantities  After  the  system’s  partition  function  is  calculated,  it  may  be  used  for  the  determination  

of  thermochemical  quantities  as  shown  above  for  the  example  of  the  inner  energy.  Since  

partition  function  factorisation  is  the  basis  of  these  calculations,  it  is  common  to  

distinguish  between  the  contributions  from  the  different  molecular  degrees  of  freedom  

to  the  quantity  under  investigation,  e.g.  in  case  of  the  enthalpy:  

qtot= q

trans!q

rot!q

vib!q

elec    à        

Htot= H

trans+ H

rot+ H

vib+ H

elec    .       (  7)  

For  example,  in  classical  thermodynamics  the  entropy  (of  a  reversible  process)  is  

defined  according  to  Clausius:       dS =

CV

TdT  .  The  heat  capacity  CV  is  given  as  the  first  

derivative  of  the  inner  energy  U  with  respect  to  the  temperature  at  constant  volume,  

thus  yielding:  

S !S

0=

1T

0

T

"##T

kBT 2 #lnQ#T

$

%&&&&

'

())))V

dT  .             (  8)  

Evaluation  of  the  integral  and  reintroduction  of  the  inner  energy  leads  to:  

S !S

0= k

BT"lnQ"T

#

$%%%%

&

'((((V

+ kB

lnQ! kB

lnQ( )T=0

 .           (  9)  

The  temperature-­‐independent  term  S0  can  thus  be  identified  as   S

0= k

BlnQ( )

T=0,  which  

is  referred  to  as  the  zero  point  entropy,  the  entropy  at  0  K.  The  remaining  terms  equal  

the  temperature-­‐dependent  entropy  according  to:  

S = k

BT!lnQ!T

"

#$$$$

%

&''''V

+ kB

lnQ    .               (  

10)  

The  remaining  thermochemical  quantities  can  be  derived  in  an  analogue  way  by  starting  

at  the  classical  expression  and  substituting  a  known  function  by  its  statistical  

counterpart  (most  often  the  inner  energy  U).  

 The  following  table  summarizes  the  most  common  thermochemical  functions  and  their  

expression  in  terms  of  the  partition  function.    

Page 4: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     4  

 

 

 Table  1:  Thermodynamic  State  Functions.  

Function   Statistical  expression  Inner energy U

U = k

BT 2 !lnQ

!T

"

#$$$$

%

&''''V

Entropy S

S = k

BT!lnQ!T

"

#$$$$

%

&''''V

+ kB

lnQ

Enthalpy H

H = k

BT

2 !lnQ

!T

"

#$$$$

%

&''''V

+ kBTV

!lnQ

!V

"

#$$$$

%

&''''T

Gibbs free enthalpy G

G = k

BTV

!lnQ!V

"

#$$$$

%

&''''T

(kBT lnQ

 Most  common  quantum  chemical  program  packages  use  the  ideal  gas  law  and  the  

quantum  models  of  the  rigid  rotator  and  the  harmonic  oscillator  as  the  basis  for  the  

calculation  of  the  partition  function.  Carrying  out  the  differentiations  according  to  the  

statistical  expressions  (see  table)  yield  constant  enthalpy  contributions  of  Htrans  =  Hrot  =  

3/2  RT  for  the  rotational  and  translational  degrees  of  freedom  (as  it  is  predicted  by  

classical  thermodynamics),  whereas  the  vibrational  part  is  given  by  

  H

vib= R

h!i

2kB

+h!

i

kB

!1

eh!

i/k

BT "1

#

$%%%%

&

'(((((

i=1

3N"6(7)

)    .             (  

11)  

The  summation  is  carried  out  over  all  vibrational  degrees  of  freedom,  which  in  the  case  

of  a  non-­‐linear  molecule  equal  3N-­‐6.  If  the  structure  under  investigation  is  a  transition  

state,  i.e.  a  maximum  on  the  potential  energy  surface  along  one  direction,  the  

corresponding  vibrational  degree  of  freedom  is  imaginary  and  left  out  in  the  summation,  

thus  reducing  the  number  of  real  vibrational  degrees  of  freedom  to  3N-­‐7.  The  first  part  

of  the  vibrational  enthalpy  is  a  sum  of  temperature-­‐independent  terms  (hυi/2kB),  which  

is  referred  to  as  the  zero  point  energy  (ZPE),  whereas  the  second  part  depends  on  the  

temperature  and  considers  those  molecules  which  are  not  in  the  vibrational  ground  

state.  If  the  electronic  ground  state  is  not  degenerated,  Helec  and  Selec  reduce  to  zero  and  

reaction  enthalpies  are  directly  given  by  the  difference  of  the  reactant’s  electronic  

Page 5: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     5  

energy.  The  analogue  expressions  for  the  entropic  contributions  are  derived  in  the  same  

way:  

 

Strans

=52R + R ! ln

VN

A

!2!Mk

BT

h2

"

#$$$$

%

&'''''

32

"

#

$$$$$$$$

%

&

'''''''''

Srot

=12R 3 + ln

!"! I

1I

2I

3!

8!2kBT

h2

"

#

$$$$

%

&

'''''

32

"

#

$$$$$$$$

%

&

'''''''''

(

)

*****

+

,

-----

Svib

= Rh#

i

kBT!

1

eh#i/kBT .1. ln 1.e.h#i/kBT( )

"

#$$$$

%

&'''''i=1

3.6(7)

/

          (  

12)  

Here  V  and  M  denote  the  system’s  volume  according  to  the  ideal  gas  law  and  the  

molecular  mass,  respectively.  The  symmetry  number  σ    is  given  by  the  order  of  the  

rotational  subgroup  in  the  molecule’s  point  group  and  can  be  understood  as  the  number  

of  sub-­‐turns  which  transfer  the  molecule  into  it’s  own  starting  structure  during  a  full  

360°  turn.  The  principle  moments  of  inertia  I1,  I2,  I3  are  the  eigenvalues  of  the  

diagonalized  inertia  matrix,  which  are  included  in  the  results  of  the  frequency  

calculation  as  well  as  the  vibrational  frequencies  υi.  Knowledge  of  these  basic  

thermochemical  quantities  allows  the  calculation  of  the  free  Gibbs  enthalpy  according  to  

G = H !T "S .  As  already  mentioned  above,  the  output  of  frequency  calculations  

contains  information  about  the  partition  functions  as  well  as  most  thermochemical  

quantities.  

1.1.3 Bond  Dissociation  Energy  and  Atomization  Energy  The  strength  of  intramolecular  bonds  in  chemical  terms  is  defined  by  the  binding  energy  

Ebind,  which  in  case  of  diatomic  molecules  is  identical  to  the  dissociation  energy  Ediss.  For  

larger  molecules  ABn,  the  binding  energy  is  the  arithmetic  mean  of  the  sum  over  all  n  

possible  A-­‐B  bond  dissociation  energies:  

  E

ABbind =

1n

Eidiss

i=1

n

!  .   (  13)  

Even   for  medium-­‐sized  molecules,   there   is   often   a   significant   difference   between   the   dissociation  

energy  and  the  binding  energy.    

In  contrast  to  the  binding  energy,  the  atomization  energy  Eatom   is  defined  as  the  reaction  energy  of  

the   atomization   reaction,   which   conveys   the   gaseous   species   ABn   into   one   atom   of   type   A   and   n  

Page 6: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     6  

atoms   of   type   B   in   the   gas-­‐phase.   A   possible   approach   to   the   determination   of   the   atomization  

energy  of  a  molecular  compound  ABn   lies   in   the  addition  of   the  enthalpy  of   formation   for   the  gas-­‐

phase  formation  reaction  (A  +  n  B  à  ABn)  and  the  atomization  reaction  for  the  transfer  of  the  most  

stable   elemental   modifications   Ax,   By   to   the   gas-­‐phase   according   to   1/x   Ax  à   A   and   1/y   By  à   B,  

respectively.  

1.2 Description  of  the  Experiment  To  understand  the  concepts,  you  will  investigate  simple  chemical  systems  and  

determine  the  change  in  Gibbs  free  enthalpy  ΔG0  as  well  as  bond  dissociation  energies  

and  atomization  energies.  The  reactions  include  the  Haber-­‐Bosch  reaction,  the  

“Knallgas”  reaction  of  hydrogen  and  oxygen  to  water,  and  the  gas-­‐phase  reaction  

between  water  and  carbon  monoxide:  

o H2  +  O2  →  2H2O  

o N2  +  3H2  →  2NH3  

o H2O  +  CO  →  H2  +  CO2    

1.   Optimize  the  geometry  of  all  participating  reactants  as  a  first  step.  Perform  a  DFT  structure  

optimization  starting  with  the  B3LYP  functional  and  the  SVP  basis  set.  

2. Run   a   frequency   analysis   for   the   optimized   geometries   and   look   for   the   relevant  

thermochemical  data  in  the  output  file  

3. Calculate   the   change   in   Gibb   free   enthalpy   for   all   investigated   reactions   and   compare   the  

results  to  experimental  data.  

Next,   the   strength   of   different   C-­‐H   bonds   will   be   examined   on   simple   organic   molecules   and   the  

atomization  energy  of   these  compounds  will  be  determined.  As   sample   systems  you  may  compare  

the  following  molecules:  methane,  acetylene,  benzene,  and  acetic  aldehyde.  Alternatively,  you  could  

choose  your  own  set  of  organic  molecules  and  compare  the  results  of  your  calculation.  

4. Optimize  the  structures  of  all  molecules.  Again  start  with  the  B3LYP/SVP  combination.  

5. Determine  the  bond  dissociation  energies  of  all  C-­‐H  bonds  and  the  C-­‐H  binding  energy  in  

methane  as  well  as  the  atomization  energies  for  all  molecules  and  compare  your  results  

to  experimental  data.  

6. Repeat  your  calculations  for  the  enlarged  basis  sets  TZVP  and  TZVPP.  Do  you  observe  any  

basis  set  effects  ?  

Table  2:  Bond  dissociation  energies  of  some  C-­‐H  bonds  

Bond   D0298  [kJ/mol]  

Page 7: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     7  

H-CH3 438.9

H-CH2 462.0

H-CH 424.0

H-C6H5 473.1

H-CCH 556.1

H-CH2CHO 394.6

H-COCH3 373.8

 

Table  3:  Gibbs  free  reaction  enthalpy  of  some  reactions  

Reaction   ΔG0  

[kJ/mol]  H2  +  O2  →  2H2O(l) -572.0

N2  +  3H2  →  2NH3 -92.3

H2O(g)  +  CO  →  H2  +  CO2 -41.2

 

1.3 Literature  • McQuarry,   D.   A.;   Simon,   J.   D.   Physical   Chemistry   –   A   molecular   approach;  

University  Science  Books:  Sausalito,  1997  

• Wedler,  G.  Lehrbuch  der  physikalischen  Chemie;  Wiley-­‐VCH:  Weinheim,  1997  

• Jensen,   F.   Introduction  to  computational  chemistry;  Wiley-­‐VCH:   Chichester,  1999  

• Holleman,  A.  F.;  Wiberg,  N.  Lehrbuch  der  anorganischen  Chemie;  Walter  de  Gruyter:  Berlin,  1995  

• Lide,  D.  Handbook  of  chemistry  and  physics;  CRC  Press:  Boca  Raton,  2000  

Page 8: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     8  

2 Computer  Experiment  6:  Computational  Chemical  Kinetics  

2.1 Background  

2.1.1 Chemical  Kinetics  Reacting  chemical  systems  are  mathematically  described  by  sets  of  coupled  first  

order  differential  equations.  The  determination  of  the  rate  law  and  the  solution  to  

the  associated  differential  equation  system  is  the  subject  of  descriptive  kinetics.  The  

result  of  the  analysis  is  the  concentration  of  each  species  involved  in  the  reaction  as  

a  function  of  time.  The  rate  at  which  different  species  are  produced  or  consumed  is  

proportional  to  the  rate  constants  k  with  is  a  characteristic  quantity  of  the  reacting  

partners  and  is  a  function  of  temperature  (vide  infra).  The  evolution  of  a  chemical  

reaction  system  is  fully  determined  from  the  initial  concentrations  of  each  species  

and  the  set  of  k’s.1    

Consider  for  example  the  comparatively  simple  reaction:  

      Ak1! "! B k2! "! C          

  (  14)  

In  which  a  substance  A  is  transformed  to  B  which  then  decays  to  the  final  product  C.  

The  time  course  of  the  concentrations  of  A,  B  and  C  is  shown  in  Figure  1  

                                                                                                               1  This  is  only  “half  true”  –  there  are  systems  which  display  “deterministic  chaos”,  e.g.  the  evolution  of  the  system  is  still  fully  determined  by  the  rate  laws  but  the  trajectories  become  “infinitely  sensitive”  to  the  initial  conditions.  In  this  case,  it  is  not  possible  to  predict  the  outcome  of  the  chemical  evolution  since  the  initial  conditions  are  never  precisely  enough   known.   While   in   most   chemical   reaction   systems,   the   time   course   of   the   various   concentrations   can   be  reasonably  well   fitted   to   sets   of   exponential   curves,   it   is   not   necessarily   true   that   the   solutions   of   the   differential  equations  are  always  exponentials.  For  example,  there  are  oscillating  chemical  reactions  and  other  oddities  which  are  of  no  concern  in  the  framework  of  this  course.    

Page 9: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     9  

0 2 4 6 8 100.0

0.2

0.4

0.6

0.8

1.0

A(t) B(t) C(t)

Con

cent

ratio

n (m

M)

Time (Sec)  Figure  1:  The  time  course  of  the  reaction  A  →  B  →  C.  The   intermediate  B   is   formed  from  the  decay  of  A  and   its  concentration  peaks  at  around  1.6  sec  in  this  example.  The  extent  to  which  B  accumulates  depends  on  the  ratio  of  the  rate  constants  k1  and  k2.  The  intermediate  B  slowly  decays  towards  the  final  product  C.  

In  chemistry  and  biochemistry  such  reactions  are  usually  followed  by  recording  

some  type  of  spectra  during  the  course  of  the  reaction.  For  example,  consider  the  

absorption  spectra  in  Figure  2.  Let  species  A  have  an  absorption  spectrum  with  two  

dominant  peaks  at  ~350  and  430  nm,  species  B  with  a  prominent  absorption  at  450  

nm  and  600  nm  and  species  C  a  spectrum  with  peaks  around  310  nm  and  700  nm.  

The  absorbance  as  a  function  of  time  that  results  from  this  reaction  system  is  then  

shown  on  the  right  hand  side  of  Figure  2.  It  has  a  fairly  complicated  appearance  

since  it  is  determined  from  the  overlapping  contributions  of  several  species  the  

relative  contributions  of  which  evolve  in  time.  Nevertheless,  the  successful  

deconvolution  of  such  data2  through  analysis  of  the  experimental  data  yields  the  

absorption  spectra  of  the  individual  species  as  well  as  the  rate  constants  k1  and  k2  

which  may  then  be  subjected  to  theoretical  analysis.  

                                                                                                               2  In   order   to   obtain   the   spectra   of   the   individual   species   from   the   convoluted   absorption   envelope,   one  may   use  techniques  like  singular  value  decomposition  (SVD).  The  details  are  outside  the  scope  of  this  course.  

Page 10: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     10  

300 400 500 600 7000

500

1000

1500

2000

2500

3000

A B C

ε (M

-1 c

m-1)

Wavelength (nm)  

300400

500600

700

0

500

1000

1500

2000

2500

0

24

68

10

Abso

rban

ce (O

D)

time (sec)wavelength (nm)  Figure  2:  Hypothetical  absorption  spectra  of  species  A,  B  and  C  (left).  Three  dimensional  plot  of  absorbance  versus  time   and   wavelength.   Note   the   appearance   and   disappearance   of   the   peak   around   600   nm   caused   by   the  intermediate  B.    

We  give  two  examples  of  rate  laws  where  the  associated  differential  equation  

systems  have  closed  form  solutions.  The  simplest  chemical  reaction  is  a  

unimolecular  decay  of  the  form:  

      Ak! "! B              

  (  15)  

With  the  rate  law:  

     

d A!"#$%&

dt='

d B!"#$%&

dt='k A!"#

$%&          

  (  16)  

We  are  interested  in  the  time  courses   A!"#$%& t( ), B!"#

$%& t( ) .  Since  matter  is  conserved  there  

is  a  conservation  law,  which  states  that:  

      A!"#$%& t( )+ B!"#

$%& t( ) = const          

  (  17)  

And  thus:  

      B!"#$%& t( ) = const' A!"#

$%& t( )          

  (  18)  

which  solves  half  of  the  problem.  Such  conservation  laws  always  exist  in  chemical  

reactions  and  serve  to  reduce  the  dimensionality  of  the  associated  differential  

equation  system.  The  remaining  equation  is  not  difficult  to  solve  and  the  solution  is:  

Page 11: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     11  

      A!"#$%& t( ) = A

0e'kt            

  (  19)  

where  A0  is  the  initial  concentration  of  [A].  Thus   B!"#$%& t( ) = A

01'e'kt( ) .  For  an  

irreversible  bimolecular  reaction  of  the  form:    

      A + B k! "! C + D            

  (  20)  

The  solution  is:  

kt =1

A0!B

0

lnB

0

A0

A"#$%&' t( )

B"#$%&' t( )

(

)

*****

+

,

------            

 (  21)  

unless   A0= B

0,  then:  

   

A!"#$%& t( ) =

1kt + 1

A0

             

  (  22)  

In  the  general  case,  the  differential  equation  systems  will  be  far  too  complex  to  

allow  a  closed  form  mathematical  solution.  In  this  case,  one  has  to  resort  to  

numerical  techniques.  While  this  is  a  rather  specialized  area  of  numerical  

mathematics,  we  briefly  illustrate  the  principle  that  underlies  such  simulations:  

assume  that  you  have  N  species  with  concentrations  C1(t),…,CN(t).  The  rate  laws  are  

of  the  form  dC1/dt=f1(C1,…,CN),  dC2/dt=f2(C1,…,CN)    and  so  on.3  Now,  we  can  try  to  

replace  the  differential  on  the  left  hand  side  by  a  finite  difference  to  obtain  an  

iterative  equation  set:  

 

C1

ti+1( )!C

1ti( )

ti+1! t

i

= f1

C1

ti( ),...,CN

ti( )( )          

  (  23)  

  …  

                                                                                                               3  Each  function  f  contains  the  concentrations  of  potentially  all  species  as  well  as  the  collection  of  rate  constants  k  that  describe  the  chemical  network.  

Page 12: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     12  

 

CN

ti+1( )!C

Nti( )

ti+1! t

i

= fN

C1

ti( ),...,CN

ti( )( )          

  (  24)  

Which  can  be  solved  for  the  unknown   C

Xti+1( )  (X=1…N):  

  C

Xti+1( ) =C

Xti( )+ t

i+1! t

i( ) fX

C1

ti( ),...,CN

ti( )( )         (  25)  

Thus,  given  the  concentrations  at  time  t0=0,  it  is  straightforward  to  obtain  the  

concentrations  at  times  t1,  t2,…  and  therefore  to  simulate  the  entire  evolution  of  the  

chemical  system.  In  order  for  the  simple  finite  difference  approximation  to  work,  

the  time-­‐step  must  be  chosen  small  enough.4  

2.1.2 Transition  State  Theory  From  a  microscopic  point  of  view,  one  is  interested  to  understand  the  value  of  k  

from  first  physical  principles.  From  phenomenological  considerations,  one  already  

knows  that  k  is  temperature  dependent  and  obeys  an  Arrhenius  type  of  equation:  

    k = Aexp !

Ea

RT

"

#$$$$

%

&'''''              

  (  26)  

Where  A  is  the  “pre-­‐exponential  factor”,  R  is  the  gas  constant,  T  the  temperature  and  

the  key  quantity  is  the  “activation  energy”   Ea.  The  larger   Ea

,  the  slower  the  

reaction.  More  detailed  insight  into  the  reaction  scenario  is  obtained  from  Eyring’s  

transition  state  theory.  In  this  case,  one  assumes  that  the  two  reacting  partners  A  

and  B  have  to  pass  through  a  special  geometrical  arrangement   AB{ }*

 (the  

transition  state,  TS)  before  decaying  to  the  products  C  and  D.  The  kinetic  scheme  

that  one  writes  is:  

                                                                                                               4  “Small   enough“   is   roughly   such   that   it   is   sufficiently   smaller   than   the   fastest   event   in   the   chemical   reaction.   The  simple   equations   given   above   are   still   numerically   rather   unstable.  Much  more   sophisticated  methods   such   as   the  Runge-­‐Kutta   procedure   or   ultimately,   the   “stiff-­‐stable”   Gear   algorithm   are   used   in   practice   for   the   integration   of  differential  equation  systems.  

Page 13: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     13  

    A + B !

k1

k"1

# $##% ### AB{ }* k*

# $# C + D        

  (  27)  

If  one  assumes  the  “pseudo-­‐steady  state”  for   AB{ }*

,  one  comes  to  the  conclusion  

that5  one  can  write  the  rate  constant  for  the  bimolecular  decay  as:  

      k = K *k*                

  (  28)  

where   K* = k

1/k!1  (the  equilibrium  constant  for  the  TS)  and  quantum  mechanics  

leads  to  the  conclusion  that:  

    k* = !

kBT

h                

  (  29)  

Where   kB= 1.38x10!38J /K  is  Boltzmann’s  constant,  h  is  Planck’s  constant  and  κ  is  

a  “transmission  coefficient”  which  is  usually  close  to  unity.6  The  “equilibrium  

constant”   K *  is  related  to  the  free  energy  of  the  transition  state  over  the  initial  state  

by:  

    K * = exp !

"G*

RT

#

$%%%%

&

'(((((              

  (  30)  

And  hence:  

    k =

kBT

hexp !

"G*

RT

#

$%%%%

&

'(((((

=k

BT

hexp !

"H*

RT

#

$%%%%

&

'(((((exp

"S*

R

#

$%%%%

&

'(((((    

  (  31)  

Where   !H*  and   !S *  are  the  enthalpy  and  the  entropy  of  the  transition  state.    

Thus,  one  identifies  the  parameters  of  the  Arrhenius  equation  with:  

                                                                                                               5  This   is   explained   in   detail   in   any   textbook   on   physical   chemistry,   e.g.   P.W.   Atkins,   Physikalische   Chemie,   VCH,  Weinheim,  in  the  latest  edition.  6  It  also  accounts  for  tunneling  through  the  barrier.  

Page 14: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     14  

    A =

kBT

hexp

!S *

R

"

#$$$$

%

&'''''              

  (  32)  

    Ea=!H

*                

  (  33)  

The  quantity   !H*  is  the  one  that  one  wants  to  calculate  while   !S *  is  slightly  more  

difficult  to  predict.  

2.1.3 Quantum  Chemical  Calculation  of  Transition  States  The  procedure  to  find  transition  states  from  quantum  chemical  calculations  is  

analogous  to  that  for  finding  minimum  structures.  The  first  step  is  to  search  for  

stationary  points  on  the  potential  energy  surface;  therefore  the  gradients  of  the  

energy  with  respect  to  the  nuclear  coordinates  have  to  vanish.  Next,  the  Hessian  

matrix  has  to  be  determined  in  order  to  characterize  the  stationary  point  as  a  

minimum,  maximum  or  saddle  point.  Only  saddle  points  with  a  single  negative  

frequency  correspond  to  transition  states.  An  example  is  shown  below  which  shows  

the  two-­‐dimensional  potential  energy  surface  of  the  system  H-­‐H-­‐H  in  a  linear  

arrangement:  

 Figure  3:  Two-­‐dimensional  potential  energy  surface  for  the  system  H-­‐H-­‐H  in  a  linear  arrangement.  A  transition  state  is  observed  around  the  H1-­‐H2  and  H2-­‐H3  distances  being  both  1.0  Angström.  

0.60.8

1.01.2

1.41.6

1.82.0

2.22.4

0

20

40

60

80

100

120

0.6

0.9

1.2

1.5

1.8

2.1

2.4

Energy (kcal/mol)

H 2-H

3 D

istan

ce (A

)

H1 -H

2 Distance (A)

78 51

423324

24

3342

5160

69

15

15

12

12

78

9.0

9.0

6.0

6.0

3.0

3.0

87

0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.40.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

H2-H

3 Dist

ance

(Ang

ström

)

H1-H

2 Distance (Angström)

Page 15: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     15  

Once  such  a  point  was  found,  the  electronic  energy  difference   ETS !Eeduct  is  known  

and  is  the  most  important  contributor  to  the  activation  enthalpy;  The  vibrational  

contribution  to  the  activation  enthalpy  is  given  by:  

 

!Hvib

* = Rh!

n

TS

2kB

1+2

exp "h!n

TS /kBT( )"1

#

$

%%%%%%

&

'

(((((((n=0

3N"7

)

"Rh!

n

educt

2kB

1+2

exp "h!n

educt /kBT( )"1

#

$

%%%%%%

&

'

(((((((n=0

3N"6

)      

  (  34)  

Where  the  sum  is  performed  over  all  vibrational  degrees  of  freedom  of  the  TS  and  

the  reactants  respectively.  For  unimolecular  reactions  the  TS  has  one  vibrational  

degree  of  freedom  less  than  the  products  (one  mode  corresponds  to  the  transition  

state  and  does  not  contribute  to  the  ZPE  of  the  TS)     !Hvib

*  is  usually  a  few  kcal/mol  

negative  and  therefore  reduces  the  barrier.    

For  bimolecular  reactions,  there  is  a  constant  contribution  from  the  rotations  and  

translations  to   !H*  of  -­‐4RT.  In  this  case,  the  TS  has  more  vibrational  degrees  of  

freedom  than  the  separated  reactants  which  makes     !Hvib

*  a  few  kcal/mol  positive.  

Likewise,  there  is  a  substantial  contribution  to  the  activation  entropy  for  

bimolecular  reactions  which  raises  the  free-­‐energy  barrier  by  10-­‐15  kcal/mol.    

Nevertheless,  everything  that  is  required  in  order  to  calculate  the  free  energy  of  

activation  can  be  approximately  deduced  from  the  outcome  of  a  frequency  

calculation  at  the  geometries  of  the  reactants  and  the  TS  respectively.  

The  tunnelling  correction   ! T( )  may  be  estimated  from  Wigner's  expression7  

! T( ) = 1+

124

h"TS

kBT

!

"####

$

%&&&&&

2

           

 (  35)  

                                                                                                               7  Wigner,  EP  Z.  Phys.  Chem.  Abt  B  1932,  19,  203  

Page 16: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     16  

where   !TS is  the  absolute  value  of  the  imaginary  frequency  corresponding  to  the  

transition  state.  

A  word  of  caution:  to  find  a  transition  state  with  quantum  chemical  programs  is  not  

easy!  Since  the  programs  have  no  way  to  “guess”  a  transition  state  from  a  stable  

structure  alone,  it  is  important  to  “guide”  the  programs  to  the  desired  transition  

states  by  providing  structures  that  are  close  to  the  final  TS.  However,  this  means  

necessarily,  that  the  outcome  of  the  calculations  depend  on  the  skill  of  the  computer  

chemist  to  guess  a  reasonable  TS  –  there  is  no  guarantee  that  the  TS  the  program  

may  find  based  on  the  guessed  structure  is  also  the  one  of  lowest  energy.  Do  not  be  

disappointed,  if  the  program  does  not  find  the  desired  TS  –  just  provide  a  better  

starting  structure.  In  calculating  chemical  reactions,  the  chemical  intuition  and  

knowledge  of  the  theoretician  is  crucial  for  success!  Please  refer  to  section  Error!  

Reference  source  not  found.  for  appropriate  input  into  the  ORCA  program.  

2.1.4 Kinetic  Isotope  Effects  There  are  two  origins  of  the  kinetic  isotope  effect.  The  first,  quantum  mechanical  

tunnelling  through  the  reaction  potential  energy  barrier  is  usually  only  important  at  

very  low  temperatures  and  for  reactions  involving  very  light  atoms.  It  may  be  

estimated  from  the  change  in  the  transmission  coefficient  caused  by  the  change  in  

the  imaginary  frequency  that  leads  to  the  transition  state  (see  Wigner’s  expression  

above).    

More  importantly  however,  kinetic  isotope  effects  are  caused  by  differences  in  the  

activation  energy  for  reactions  involving  different  isotopes  since  the  reactants  and  

the  TS  have  different  zero  point  vibrational  energies.  The  effect  on  the  activation  

barrier  for  a  reaction  involving  an  H/D  atom  is  shown  below.  

Page 17: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     17  

 Figure   4:   Definition   of   energetic   quantities   in   the   calculation   of   chemical   reactions   on   the   basis   of   qualitative  potential  energy  surfaces.  On  the  x-­‐axis,  the  “reaction  path”  and  on  the  y-­‐axis  the  total  energy  is  plotted.  

The  net  effect  is  that  the  activation  energy  is  higher  for  the  heavier  isotope,  and  

therefore  the  reaction  will  be  slower  (a  'normal'  isotope  effect).  The  maximum  

isotope  effect  is  obtained  when  the  bond  involving  the  isotope  is  completely  broken  

in  the  transition  state,  in  which  case  the  difference  in  activation  energies  is  simply  

the  difference  in  zero  point  energies  of  the  stretching  frequencies  for  the  bond  being  

broken.8    

From  transition  state  theory,  one  readily  deduces:  

kH

kD

= exp !"H

*H( )!"H

*D( )

kBT

#

$

%%%%%

&

'

((((((      

 (  36)  

Where  “H”  and  “D”  have  been  written  for  the  two  isotopes  since  hydrogen  to  

deuterium  is  the  most  frequently  studied  isotope  effect.    The  difference  

!H

*H( )"!H

*D( )  is  entirely  determined  by  the  different  ZPE  contributions  to  

!Hvib

* .  

                                                                                                               8  In   some   reactions   it   is   the   zero   point   energy   difference   between   the   transition   states  which   governs   the   kinetic  isotope   effect.   In   this   case   the   activation   energy   is   larger   for   the   lighter   isotope   and   an   inverse   isotope   effect   is  observed,  in  which  the  heavier  isotope  containing  reactant  undergoes  faster  reaction.  

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

Page 18: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     18  

2.2 Description  of  the  Experiment  

2.2.1 Transition  State  of  Glyoxal  • Take  the  optimized  geometry  of  the  glyoxal  isomer  which  has  been  calculated  in  

the   previous   section   and   rotate   one   carbonyl   group   Error!   Bookmark   not  

defined.   by   90°,   so   that   it   is   arranged   perpendicular   to   the   opposite   carbonyl  

group.  

• Calculate  the  transition  state  using  B3LYP/SVP.    

• Find  out  how  quickly  glyoxal   interconverts  at  room  temperature  by  calculating  

the  activation  energy  and  the  rate  constant.    

2.2.2        H/D  Kinetic  Isotope  Effect  • Calculate  the  H→D  effect  on  the  reaction  rate  of  the  reaction  CH4  +  OH•  →  H2O  +  

CH3.  Build  all  molecules  using  MOLDEN  and  export  the  Z-­‐matrix  to  the  Gaussian  

format.  Then  run  a  geometry  optimization  for  all  compounds  using  B3LYP/SVP.  

(HINT:   in   the   transition   state   one   of   the   C-­‐H   bonds   must   be   significantly  

stretched  and  the  H-­‐O  bond  should  already  be  partially  formed.  You  should  try  to  

start  with  such  a  structure.  A  rule  of  thumb  is  to  stretch  the  bond  to  ~1.5  times  

its  equilibrium  value).  

• Think  of  a  possible  transition  state  structure  and  try  to  guide  the  optimization  to  

this  transition  state.    

• Calculate   the   activation   enthalpy   and   activation   entropy   using   the   zero-­‐point  

corrected   energies.   Calculate   the   reaction   rates   and   predict   and   the   resulting  

isotope  effect.  

2.2.3            Rotational  Barrier  of  Ethane  • Study   the   rotational   barrier   of   ethane.   Determine   the   energies   of   the   eclipsed  

and   staggered   conformations.   Perform   a   relaxed   and   an   unrelaxed   potential  

energy  surface  scan  and  plot  the  results.    

Refer  to  section  Error!  Reference  source  not  found.  for  a  description  of  how  to  

generate  relaxed  surface  scans  with  the  ORCA  program.  

Page 19: ORCA labs V thermochemistry TS - University of Waterlooscienide2.uwaterloo.ca/~nooijen/Chem-440... · Computational+Thermochemistry+ + 3+ partition&function&(some&authors&even&say&that&the&partition&function&is&the&“wave&

Computational  Thermochemistry     19  

• Build  the  ethane  molecule  using  MOLDEN  and  export  the  Z-­‐matrix.    

• Perform  an  unrelaxed  surface  scan  using  the  Scan  keyword  in  the  command  line  

and  enter  the  parameters  to  change  (See  sample  input  below).  

• Perform  a  relaxed  surface  scan  using  the  Opt  keyword  and  enter  the  parameters  

to  change.  (See  sample  input  below).  

• What  happens  generally  if  one  does  not  relax  the  rest  of  the  geometry  during  the  

surface   scan?   Will   you   over-­‐   or   understimate   the   barriers   calculated   by  

unrelaxed  surface  scans  ?  

Compare  to  the  available  experimental  data.  The  rotational  barrier  of  ethane  is  about  1024  cm-­‐1  ≈12.25  kJ/mol.9  

                                                                                                               9Weiss,  S  and  Leroi,  GE,  J.  Chem.  Phys.,  1968,  48,  962