Sei sulla pagina 1di 10

Catalytic
Reduction
of
CO2


Into
Hydrogen
Using

Microreactors


Malavarayan.S

Ravishankar.M


IInd
year
Chemical
Engineering,


Sri
Venkateswara
College
of
Engineering,

Sriperumbudur


Postal
Address:


No:
50,
NGO
Colony
2nd
street,

Vadapalani,

Chennai‐26.


E­mail:
malavarayan@gmail.com


Phone:
+91
9884314549



Abstract:

The
stoichiometric
reduction
of
carbon
dioxide
by
soluble
K[Ru[III]‐(EDTA‐
H)Cl].2H2O
complex
at
CO2
pressure
in
the
range
of
5‐70
atm
and

temperature
40‐80
°C
in
a
microreactor
gave
formic
acid
and

formaldehyde
as
the
reaction
products.
The
rates
of
formation
of

formation
of
formic
acid
and
formaldehyde
are
of
first
order

dependence
with
respect
to
the
catalyst
and
the
dissolved
carbon

dioxide
concentrations.
The
formic
acid
this
produced
is
converted
into

hydrogen
by
water‐gas
shift
method
using
10%
Cu‐Al2O3
as
a
catalyst
in

a
separate
microreactor
at
a
temperature
of
about
130
°C
to
give

hydrogen
gas
.
The
Energy
is
only
given
for
maintaining
the
overall

experimental
conditions.
The
hydrogen
thus
obtained
can
be
used
for

various
industrial
purposes
the
most
significant
being
the
conversion
of

hydrogen
into
Electrical
Energy
using
Fuel
Cells.



Introduction:


Carbon
dioxide
and
water
are
the
end
products
of
many
chemical
and

biological
processes.
The
Utilization
of
CO2
to
synthesize
value‐added

chemicals
is,
therefore,
of
much
importance
in
industry.
CO2
is
used

industrially
in
large
quantities
for
the
preparation
of
Urea
[1]
and

inorganic
compounds
such
as
carbonates
and
bicarbonates.
Activation
of

CO2
by
transition
metal
complexes
for
form
adducts
is
an
area
of

investigation
[2‐8].
The
synthesis
of
methyl
formate
by
the
reaction
of


CO2
,
H2

and
CH3OH
catalyzed
by
Ru(PPh3)Cl2
is
a
good
example
of
the

reductive
fixation
of
CO2
[9]
.
CO2
+
H2
have
been
used
in
the
place
of
CO

+
H2
in
the
synthesis
of
homologous
alkenes,
alcohols
and
amines
in
the

presence
of
NH3

[10‐12].

The
rhodium
complex
Rh(diphos)(BPh4)

catalyzes
the
co‐oligomerization
of
CO2
and
methyl
acetylene
to
give

cyclic
compounds
.


































This
paper
reports
the
reduction
of
CO2
to
formic
acid

and
formaldehyde
by
the
water
soluble
K[RuIII(EDTA)Cl].2H2O
system

under
mild
experimental
conditions
(40
‐80
°C

,
5‐70
atm
).
The
system

suggests
the
possibility
for
a
catalytic
cycle
for
the
fixation
of
CO2
to

value
added
products.
The
reaction
is
carried
out
in
a
microreactor
that

can
provide
the
necessary
experimental
conditions
viz.
the
temperature

and
the
pressure.
A
stack
containing

‘n’
number
of
micro
reactors
can

be
employed
where
‘n’
would
depend
on
the
application
of
use.



Experimental:


The
carbon
dioxide
fixation
was
performed
in
a
300
ml
pressure
reactor

by
[3].
Carbon
dioxide
gas
with
a
purity
greater
than
99.5%
and
complex

K[RuIII(EDTA‐H)Cl).2H2O
were
used
in
the
300
ml
reactor
along
with

double
distilled
water.
The
complex
was
prepared
using
RuCl3.3H2O
and

disodium
salt
of
EDTA
[15].
In
a
sample
run,
a
known
amount
of
complex

1
was
dissolved
in
100
ml
of
distilled
water
and
placed
in
the
pressure

reactor.
The
bomb
was
pressurized
by
CO2
to
the
required
value
when

the
desired
temperature
was
attained.
The
progress
of
the
reaction
was

monitored
by
analyzing
liquid
samples
withdrawn
at
different
intervals

of
time
for
HCOOH
and
HCHO
contents.
The
concentration
of
HCHO
and

HCOOH
present
in
the
liquid
samples
were
estimated

spectrophotometrically
by
monitoring
the
peak
at
412
nm
using
Nash’s

reagent
[16,17].
From
the
graphs
of
time
vs.
amount
of
HCHO
or
HCOOH

formed,
the
rates
of
the
formation
of
both
the
products
evaluated.


Gaseous
samples
were
also
withdrawn
at
suitable
intervals
of
reaction

time
and
analyzed
for
CO2
and
CO
content
using
a
GLC.
The
column
used

for
the
analysis
of
gases
were
2.5
m
long,
100
mesh
with
a
TC
detector

(150
°C),
column
temperature
40
°C,
injector
temperature
50
°C
and
H2

carrier
gas
with
a
flow
rate
of
30
ml
min‐1.
The
required
solubility
data
of

CO2
in
water
at
different
temperatures
and
pressure
were
taken
from

the
reported
values
[18],
and
the
total
dissolved
CO2
concentration

needed
to
interpret
the
kinetic
data
was
calculated.




Results
and
Discussion:


In
order
to
fix
the
conditions
for
kinetic
study,
the
experiment
was


Conducted
with
1
mmol
of
complex
1,
a
CO2
partial
pressure
of
68.0
atm


And
at
80
°C.
The
yields
of
HCOOH
and
HCHO
formed
as
a
function
of


time
is
shown
in
Fig.
1.
It
is
clear
from
the
graph
that
the
concentrations


Of
both
HCHO
and
HCOOH
increase
with
time,
reaching
a
maximum
in


3
‐
4
h;
later
both
the
products
tend
to
dissociate,
resulting
in
a
decrease

in
 their
 concentrations.
 The
 reduction
 in
 the
 concentration
 of
 HCHO
 is

more
 pronounced
 than
 that
 of
 HCOOH.
 Therefore,
 the
 rates
 of

formation
of
HCHO
and
HCOOH
were
evaluated
in
the
interval
between

1‐3
h,
which
ensured
a
linear
dependence.
The
decay
rate
of
HCHO
was

evaluated
in
the
declining
portion
of
the
curve
after
3.5h.







I 1
0 1 2 TIM:.(h) L 5 6

Fig. 1. The yields of HCOOH (0) and HCHO (A) in CO2 reduction by K[ Rum(ED
Fig
1.
The
yields
of
HCOOH
and
HCHO
in
CO2
reduction
by
K[RuIII(EDTA‐H)‐
C1]*2Hz0.
Cl].2H 2O



HCHO and HCOOH were evaluated in the interval between 1 - 3
Kinetics

ensured a linear dependence. The decay rate of HCHO’was evaluat










The
kinetic
experiments
on
CO2
reduction
were
conducted
by

declining portion of the curve after 3.5 h.
varying
the
concentrations
of
CO2
and
rates
of
formation
of
HCHO

and
HCOOH
were
evaluated.

Kinetics
Effect
of
complex
concentration

The kinetic experiments on CO* reduction were conducted by
the concentrations of COz, and rates of formation‐3
M




The
concentration
of
the
complex
1
was
varied
from
0.5
–
2
×10 of HCHO and
were evaluated.
with
a
constant
dissolved
CO 2
concentration
of
3.9
×
10 
m
and

‐2

temperature
of
80
°C
for
all
the
runs.
The
dependence
of
rates
of

formation
of
HCHO
and
HCOOH
on
concentration
of
conplex1
are

Effect of complex concentration
shown
Fig.
2
and
3,
respectively.
Both
HCHO
and
HCOOH
rates
show

The concentration of complex 1 was varied from 0.5 - 2 X lop3
a
first
order
dependence
with
respect
to
complex
concentration.

a constant dissolved CO2 concentration of 3.9 X lo-* M and tempe
80 “CL! for all the 2
runs.
Effect
of
dissolved
CO The dependence of rates of formation of HC
concentration


HCOOH on the concentration of complex 1 are shown in Figs.
Figures
4
&
5
show
the
plot
of
rates
of
formation
of
HCHO
and

respectively. Both HCHO and HCOOH rates show a first order dep
HCOOH
vs.
dissolved
carbon
dioxide
concentration
the
experiments

with respect to complex 1 concentration.
were
performed
keeping
a
constant
complex
concentration
of
1
x

10‐
3
M
,
80
°C
and
varying
the
dissolved

CO 
concentration
from
1.98
X

2
Effect of dissolved CO2 concentration
Figures 4 and 5 show the plot of rates of formation of HC
HCOOH vs. dissolved CO2 concentration. The experiments were pe
keeping a constant complex concentration of 1 X 10e3 M, 80 “C and
the dissolved CO2 concentration from 1.98 X lo-* M to 7.73 X
10‐2
M
to
7.73
X
10‐2
M.
The
rates
of
formation
of
both
HCHO
and

HCOOH
show
a
linear
dependence
of
CO2
concentration.

Temperature

The
temperature
of
the
experiments
for
reduction
of
CO2
by
complex


1
was
varied
between
40
‐
80
°C,
with
other
conditions
constant
such
as


Catalyst
concentration
of
1
X
10
‐3M
and
dissolved
CO2
concentration
of

3.9
X
10‐2
M.
Figure
6
shows
the
graph
of
‐‐In
rate
vs.
l/T.
From
the

slopes
of
both
the
straight
lines,
the
activation
energies
evaluated
for

HCOOH
and
HCHO
are
4.8
kcal
mol‐1
and
3.5
kcal
mol‐1
,
respectively


0 0.5 14 1.5
COMPLEX j_CONCENTRATiON,x103 M COMPLEX 1 CONCENTRATION, ~10~ M
Fig. 2. Effect of complex 1 con~entratj~n on the rate of HCHO formation.

Fig, 3. Effect of complex 1 concentration on the rate of HCOOH formation.


0 0.5 14 1.5 


 COMPLEX j_CONCENTRATiON,x103 M COMPLEX 1 CONCENTRATION, ~10~ M
Fig. 2. Effect of complex 1 con~entratj~n on the rate of HCHO formation.

Fig, 3. Effect of complex 1 concentration on the rate of HCOOH formation.

Fig. 4. Effect of dissolved CO2 concentration on the rate of HCHO formation.

3.9 X low2 M. Figure 6 shows the graph of --In rate us. l/T. From the slopes
of both the straight lines, the activation energies evaluated for HCOOH and
Fig. 4. Effect
HCHO are 4.8of kcal
dissolved CO2and
mol-* concentration on the, rate
3.5 kcd mol-' of HCHO formation.
respectively. 

Mechanism
3.9 X low2 M. Figure 6 shows the graph of --In rate us. l/T. From the slopes
The the
of both reduction
straightoflines,
CO2 the
by activation
complex 1 energies
gave twoevaluated
products, for
namely HCHO
HCOOH and
and
HCHOHCOOH.
are 4.8The
kcalrates
mol-*of and
formation
3.5 kcd ofmol-'
both , the products have a first-order
respectively.
dependence with respect to catalyst and dissolved CO2 concentrations.
Mechanism
The reduction of CO2 by complex 1 gave two products, namely HCHO
Mechanism:


The
reduction
of
CO2
by
complex
1
gave
two
products,
namely
HCHO


and
HCOOH.
The
rates
of
formation
of
both
the
products
have
a
first‐
order
dependence
with
respect
to
catalyst
and
dissolved
CO2

concentrations.
Based
on
these
observations,
the
mechanism
proposed

for
the
reactions
is
shown
in
Scheme
1.
The
complex
1
undergoes
a
rapid

aquation
in
presence
of
water
to
give
Ru(III)
aquo
species
2
[19],
which

reacts
with
CO2
to
form
a
CO,
adduct
3.
The
formation
of
such
CO2

adducts
were
reported
in
the
literature
[3
–
5].
Species
3
activates
water

oxidatively
in
step
(2)
to
form
the
Ru(V)
species
4.
The
insertion
of
CO2

into
the
Ru‐H
bond
of
4
takes
place
in
step
(3)
to
form
an
η1‐formate

species
5.
The
metalloformate
complex
5
undergoes
a
rapid

intermolecular
proto‐
nation
resulting
in
the
formation
of
formic
acid

and
RuV‐oxo
species
6.
Reductive
elimination
of
metalloformate

complexes
to
formate
derivatives
such
as
methyl
formate
from

methanol,
CO2
and
H2
had
been
reported
[9].
The
electrons
needed
for

the
reduction
of
CO2
to
HCOO‐
in
our
case
are
obtained
by
the
hydride

insertion
step
(3).
The
formation
of
ruthenium
(V)
oxo
species
6
is

characterized
by
the
appearance
of
a
band
at
390
nm.
Species
6
also

gives
v(Ru=O)
at
810
cm‐‘.
In
step
(5),
the
formic
acid
formed
in
step
(3)

is
catalytically
decomposed
to
CO
and
Hz0
by
complex
2.



 310
K
LRu**~(H~O) + co, s LRu”‘( COz) (1)
2 3

H
fast
LRu’“‘(C02) + Hz0 - L+v(Co,) (2)
OH

H
k 0
L&F&*) - LRuy_-+C” (3)
k--H ‘OR
!3H
5
fast
1
0
II
LRuV + HCOOH

6


 /o kl H,

c -
c=O + Hz0 + LRuV (4)
LYv-cN
OH- H
OH H’
::
:
+I4 6

kz
HCOOH e CO + Hz0 (5)
LRur”(H20)
fast
1
0
II
LRuV + HCOOH

/o kl H,

c -
c=O + Hz0 + LRuV (4)
LYv-cN
OH- H
OH H’
::
:
+I4 6

kz
HCOOH e CO + Hz0 (5)
LRur”(H20)

k3
CO + H,O - COz + Hz (6)

LRuV + H\C=O kq LRu”’ + HCOOH (7)


H’
8

Scheme 1. Mechanism for the stoichiometric fixation of CO2 by LFlu”‘(H20).




 309
The rate of decomposition of formic acid catalyzed by 1 was estimated in a
separate experiment. It was confirmed by the analysis of both liquid as well
as gas samples withdrawn at different time intervals that HCOOH decom-

DISSOLVED CO 2 CONCENTRATION, x10 ' M

Fig. 5. Effect of dissolved CO* concentration on the rate of HCOOH formation.

111 I
2.8 2.9 3.0 3.1 3.2
' / T , 1O3 K

Fig. 6. Effect of temperature on the rates of formation of HCOOH and HCHO.





 Based on these observations, the mechanism proposed for the reactions is

 shown in Scheme 1.
The complex 1 undergoes a rapid aquation in presence of water to give
Rate
laws:

Ru(II1) aquo species 2 [19], which reacts with CO? to form a CO, adduct 3.

The
rate
constants
for
the
water
–shift
reaction
(k
The formation of such COz adducts were reported in the3)
,
formic
acid

literature [3 - 51.
Species 3 activates water oxidatively in step (2) to form the Ru(V) species 4.
decomposition
(k 2
)
and
the
carbon
dioxide
reduction
to
HCHO
(k
The insertion of CO2 into the Ru-H bond of 4 takes place in step (3) to1
)

are

form an rjl-formate species 5.
The metalloformate complex 5 undergoes a rapid intermolecular proto-
nation resulting in the formation of formic acid and RuV-0x0 species 6.
Reductive elimination of metalloformate complexes to for-mate derivatives
such as methyl formate from methanol, COz and Hz had been reported [9].
given
as
in
the
following
precedence






K3
>>
K2
>>
K1
>>
K
>>
K4


 311
From
the
intercepts
and
slopes
of
the
rate
equations
given
below
the

poses into CO and HzO. The CO obtained in step (5) is rapidly converted
rate
constants
are
evaluated
as


to COz by the water-gas shift reaction [20]. 311

 The $-carboxylate species 5 reacts with Hz in step (4) to give HCHO
and Ru(V)-0x0
into ‐1CO species
HzO. 6.TheThus, the Ru(V)-0x0
in step species is an end product
K=
7.46
M
poses
of reduction

 ofandCOz. In step
CO obtained
(7), HCHO is oxidized
(5) is rapidly
to HCOOH
converted
by LRuV=O
to COz by the‐1water-gas shift reaction [20].
K1=0.33
min

species 

Theregenerating
$-carboxylate species 2. 5 reacts with Hz in step (4) to give HCHO
species
‐1
.
K3and
=
90.1
min

Ru(V)-0x0 
 species 6. Thus, the Ru(V)-0x0 species is an end product
ofRate laws of COz. In step (7), HCHO is oxidized to HCOOH by LRuV=O
reduction

 speciesFrom the kinetic
regenerating observations
species 2. of the reduction of CO2 to HCOOH and
By
taking
a
300
ml
reactor
and
maintaining
at
the
optimal
experimental

HCHO by K[ Ru”‘(EDTA-H)Cl] *2Hz0, the rate equation can be written as:
Rate laws
kKIRuxl’(EDTA-H)(HaO)]r[COz] 3
conditions
the
output
of
hydrogen
obtained
was
found
to
be
0.42
m
Fl A From the kinetic observations of the reduction of CO2 to HCOOH and
/s
.




 HCHO by K[ Ru”‘(EDTA-H)Cl] *2Hz0, the rate equation can be written as:
1+ [CO21
kkll([Ruul(EDTA-H)(H20)]*[C02]
kKIRuxl’(EDTA-H)(HaO)]r[COz]
Fl F2 A = (2)
1+ 1+ [CO21
[CO21

where rl = rate of formation of HCOOH, r2 = rate of formation of HCHO,


kkll([Ruul(EDTA-H)(H20)]*[C02]
F2[ =Ru”‘(EDTA-H)(H20)lT = total complex 1 concentration, [CO,] =(2)dis-
solved CO2 concentration,
1+ [CO21 K = equilibrium
 constant, k = rate constant for
the formation
rl = rate of
of HCOOH,
formationand = rate constant
r2 = rateofofHCHO formation.

 where
Equations (1) and (2)
of k,HCOOH,
can becomplex
rearranged
formation of HCHO,
[ Ru”‘(EDTA-H)(H20)lT = total 1 in the following form
concentration, [CO,]to=evalu-
dis-

 ate theCO2
solved kinetic constants. K = equilibrium constant, k = rate constant for
concentration,
the formation of HCOOH, and
[Ru”‘(EDTA-H)(H20)lT = k, = rate constant of HCHO formation.
Equations (1) and (2) can be rearranged in the following form to evalu- (3)
ate the kineticFl constants.
[Ru”‘(EDTA-H)(H20)IT
[Ru”‘(EDTA-H)(H20)lT == 1 1 1
(4)
(3)
Fl
‘-2 i kk,K x[cozl 1 +kk,

From the intercepts and slopes of the above equations, the values of K, k

 [Ru”‘(EDTA-H)(H20)IT = 1 1 1
and k1 evaluated at 80 “C are: (4)
Results
and
Conclusions:

‘-2 i kk,K x[cozl 1 +kk,
K = 7.46 M-i

 From the intercepts and slopes of the above equations, the values of K, k
k =k10.33
and min-’ at 80 “C are:
evaluated 3

Thus
the
above
paper
reports
on
production
of
0.42
m /
s
of
hydrogen

= 3.03
Kkl= 7.46 min-’
‐2
M-i
with
4
X
10 
M
of
HCOOH
production
from
the
dissolved
Carbon
dioxide

k = 0.33 The water-gas shift reaction occurring between CO and H20 under
min-’
in
a
300
ml
reactor
under
the
temperature
of
around
40
–
70
°C
and

the reaction conditions has a rate constant (k3 = 90.1 min-‘) reported in our
pressure
of
about
70.9
bars.
The
proposal
is
the
usage
of
Microreactors

kl = 3.03 min-’
earlier studies [20]. The values of k and k,, ‘the rate constants for formic
acidTheandwater-gas
formaldehyde shift formation, respectively,
that
can
maintain
the
optimum
experimental
conditions
at
the
same

reaction occurring are much
between CO and smaller
H20 than
underk3
andreaction
the hence during the CO2
conditions has areduction
rate constantthe CO (k3 generated
= 90.1 min-‘)by the decomposition
reported in our
time
give
the
standard
CO
of HCOOH
earlier studiesparticipates
[20]. The values 2
mentioned
above
.
The
advantages
of
using

rapidly ofin ktheandwater-gas shiftconstants
k,, ‘the rate reaction, for
giving
formicCO2
the
Microreactors
are

and and
acid Hz. formaldehyde
The hydrogen formation,
produced by the water-gas
respectively, are shift
muchreaction
smaller isthan
utilized
k3
in the
and hencereduction
during theof HCOOH and HCHO.
CO2 reduction the COThe value ofbyk2the(rate
generated constant for
decomposition

 ofHCOOH
HCOOHdecomposition), determined
participates rapidly independently
in the water-gas shift in the absence
reaction, givingofCO2CO2
and
and• Hz.Microreactors
typically
have
heat
exchange
coefficients
of
at
least

in the
The presence
hydrogenofproduced
complex by 1 atthe80water-gas
“C, is 1.52shift
min-‘. The formic
reaction acid
is utilized
indecomposition rate constant
HCOOH and (k,) is smaller
the 1
megawatt
per
cubic
meter
per
kelvin,
up
to
500
MW
m
reduction of HCHO. The value thanof k2 (water-gas
(k3)(rate constant forshift

³
K‐¹
vs.
a

HCOOH decomposition), determined independently in the absence of CO2‐ ‐
and infew
kilowatts
in
conventional
glassware
(1l
flask
~10
kW
m
the presence of complex 1 at 80 “C, is 1.52 min-‘. The formic acid ³
K ¹)).

Thus,
microreactors
can
remove
heat
much
more
efficiently
than

decomposition rate constant (k,) is smaller than (k3) (water-gas shift
vessels
and
even
critical
reactions
such
as
nitrations
can
be

performed
safely
at
high
temperatures.
Hot
spot
temperatures
as

well
as
the
duration
of
high
temperature
exposition
due
to

exothermicity
decreases
remarkably.
Thus,
microreactors
may

allow
better
kinetic
investigations,
because
local
temperature

gradients
affecting
reaction
rates
are
much
smaller
than
in
any

batch
vessel.
Heating
and
cooling
a
microreactor
is
also
much

quicker
and
operating
temperatures
can
be
as
low
as
‐100
°C.
As
a

result
of
the
superior
heat
transfer,
reaction
temperatures
may
be

much
higher
than
in
conventional
batch‐reactors.
Many
low

temperature
reactions
as
organo‐metal
chemistry
can
be

performed
in
microreactors
at
temperatures
of
‐10°C
rather
than
‐
50°C
to
‐78°C
as
in
laboratory
glassware
equipment.

• Microreactors
are
normally
operated
continuously.
This
allows
the

subsequent
processing
of
unstable
intermediates
and
avoids

typical
batch
workup
delays.
Especially
low
temperature

chemistry
with
reaction
times
in
the
millisecond
to
second
range

are
no
longer
stored
for
hours
until
dosing
of
reagents
is
finished

and
the
next
reaction
step
may
be
performed.
This
rapid
work
up

avoids
decay
of
precious
intermediates
and
often
allows
better

selectivities
.

• Continuous
operation
and
mixing
causes
a
very
different

concentration
profile
when
compared
with
a
batch
process.
In
a

batch,
reagent
A
is
filled
in
and
reagent
B
is
slowly
added.
Thus,
B

encounters
initially
a
high
excess
of
A.
In
a
microreactor,
A
and
B

are
mixed
nearly
instantly
and
B
won't
be
exposed
to
a
large

excess
of
A.
This
may
be
an
advantage
or
disadvantage
depending

on
the
reaction
mechanism
‐
it
is
important
to
be
aware
of
such

different
concentration
profiles.

• Although
a
bench‐top
microreactor
can
synthesize
chemicals
only

in
small
quantities,
scale‐up
to
industrial
volumes
is
simply
a

process
of
multiplying
the
number
of
microchannels.
In
contrast,

batch
processes
too
often
perform
well
on
R&D
bench‐top
level

but
fail
at
batch
pilot
plant
level


• Pressurisation
of
materials
within
microreactors
(and
associated

components)
is
generally
easier
than
with
traditional
batch

reactors.
This
allows
reactions
to
be
increased
in
rate
by
raising

the
temperature
beyond
the
boiling
point
of
the
solvent.
This,

although
typical
Arrhenius
behaviour,
is
more
easily
facilitated
in

microreactors
and
should
be
considered
a
key
advantage.

are +12.3 e.u. for HCOOH and +8.3 e.u. for HCHO. These entropies are
considerably more negative than the entropies observed [21] for the conver-
sion of CO to HCOOH and HCHO. This reflects the higher order of rear-
rangements in the coordination sphere of the metal ion in CO2 insertion as
compared to CO insertion. In general, the enthalpies and entropies are more
Pressurisation
may
also
allow
dissolution
of
reactant
gasses
within

negative in HCHO formation than HCOOH formation. Since a hydrogen
the
flow
stream.

molecule is needed for the reduction of coordinated q’-format0 or a formyl
Thus
using
stacks
of
microreactors
one
can
achieve
very
high
power

group, the transition state requires greater reorganisation in HCHO reaction
yield.
The
hydrogen
thus
produced
can
be
used
in
the
production
of

as compared to HCOOH, hence a more negative entropy. The negative
Electricity
in
fuel
cells
as
well
in
production
of
ammonia
and
other

enthalpy reflects the higher bond strength in HCHO and HCOOH.
organic
compounds.



References
References

1 K. Othmer, Encyclopedia of Chemical Technology, Vol. 23, Wiley-Interscience,


New York, 1983, p. 548.
2 R. P. A. Sneeden, in G. Wilkinson, F. G. A. Stone and E. W. Abel (eds.), Comprehen-
sive Organometallic Chemistry, Pergamon, Oxford, 1982, p. 225 - 283 and references
therein.
3 M. E. Volpin and I. S. Kolommikov, in E. I. Becker and M. Tsutsui (eds.), Organo-
metallic Reactions, Vol. 5, Interscience, New York, 1975, p. 313.
4 I. S. Kolommikov and M. Kh. Grigoryan, Russ. Chem. Rev. (Engl. Transl.), 47 (1978)
334.
5 R. Eisenberg and D. E. Hendriksen, Adu. Catal., 28 (1979) 79.
6 T. Herskovitz, J. Am. Chem. Sot., 99 (1977) 2391.
7 D. J. Darensbourg and C. Ovalles, J. Am. Chem. Sot., 106 (1984) 3750.
8 M. Aresta and C. F. Nobile, J. Chem. Sot. Dalton Trans., (1977) 708.
9 I. S. Kolommikov, T. S. Lobeeva and M. E. Volpin, Zzu. Akad. Nauk SSSR, Ser.
Khim., (1972) 2329.
10 F. S. Karn, J. F. Schultz and R. B. Anderson, Znd. Eng. Chem. Prod. Res. Dev., 4
(1965) 265.
11 R. N. R. Mulford and W. W. Russel, J. Am. Chem. Sot., 74 (1952) 1969.
12 H. Kuster, Brennst-Chem., 17 (1963) 203.
13 P. Albano and M. Aresta, J. Organometall. Chem., 190 (1980) 243.
14 M. M. Taqui Khan, S. B. Halligudi and S. H. R. Abdi, J. Mol. Catal., 45 (1988) 215.
15 A. A. Diamantis and J. V. Dubrawaski, Znorg. Chem., 20 (1981) 1142.
16 T. Nash, Biochem. J., 55 (1953) 416.
17 H. N. Wood and H. Gest, in S. P. Colowick and N. 0. Kaplan (eds.), Methods of Enzy-
mology, Vol. 3, Academic Press, New York, 1957, pp. 287 - 289.
18 A. Seidel and W. F. Linke, Solubilities of Inorganic Compounds, Van Nostand, New
York, 1952, p. 90. 



 19 M. M. Taqui Khan, Amjad Hussain, G. Ramachandraiah and M. A. Moiz, Znorg. Chem.,


 25 (1986) 3023.
20 M. M. Taqui Khan, S. B. Hailigudi and S. Shukla, Angew. Chem. Znt. Ed. Engl.,
27 (1988) 1735.
21 M. M. Taqui Khan, S. B. Halligudi, N. Nageswara Rao and S. Shukla, J. Mol. Catal., 51
(1989) 161.

Potrebbero piacerti anche