Sei sulla pagina 1di 8

View Article Online / Journal Homepage / Table of Contents for this issue

Green Chemistry

Dynamic Article Links

Cite this: Green Chem., 2011, 13, 1267

PAPER

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

www.rsc.org/greenchem

From greenhouse gas to feedstock: formation of ammonium carbamate from


CO2 and NH3 in organic solvents and its catalytic conversion into urea
under mild conditions
Francesco Barzagli,a Fabrizio Mani*a and Maurizio Peruzzinib
Received 12th October 2010, Accepted 8th February 2011
DOI: 10.1039/c0gc00674b
The capture of carbon dioxide by ammonia in both aqueous and non-aqueous solutions was
investigated at atmospheric pressure and 273 K under different operating conditions. The CO2
capture is fast and efcient ranging between 78 and 99%, depending on both the NH3
concentration and the solvent nature. The precipitation of solid mixtures of ammonium
bicarbonate, ammonium carbonate and ammonium carbamate occurred in ethanolwater
solution. Selective precipitation of ammonium carbamate was achieved by reacting gaseous CO2
and NH3 in anhydrous ethanol, 1-propanol or N,N-dimethylformamide (DMF) in a ow reactor
that operates in continuous. In the second step of the process, the pure ammonium carbamate is
used to produce urea with good yield (up to 54% on carbamate basis) at 393413 K in the presence
of inexpensive Cu(II) and Zn(II) catalysts. The yield of urea depends on several factors including
the catalyst, the reaction temperature and the reaction time. Identication and quantication of
urea in the reaction mixtures was obtained by analysis of its 13 C NMR spectrum. A preliminary
mechanistic interpretation of the catalytic reaction is also briey presented and commented.

Introduction
The reduction of anthropogenic CO2 emissions is considered
one of the most urgent challenges1 and imperative strategies must be adopted to limit CO2 emissions by improving
energetic efciency, bolstering the use of alternative energy
sources (biomass, wind farms and photovoltaic cells), favouring
the change in fuels and adopting efcient CO2 capture and
sequestration technologies. Among the last technology, the
ammonia scrubbing process provides the advantage of high CO2
loading capacity and absorption efciency with no absorbent
degradation.2 However, this process suffers from serious energy
penalties due to NH3 loss avoidance and regeneration and for the
cost related to its separation from concentrated CO2 that must
be compressed and sequestered. The ultimate goal for setting up
a sustainable CO2 capture process must be therefore aimed at
lowering the costs of the process and, even more important,
at maximizing the net balance CO2(captured) - CO2(emitted) ,3 that
represents the true CO2 avoided.
In our laboratory we are developing a new concept of CO2
capture technology which combines the necessary CO2 abate-

a
University of Florence, Department of Chemistry, via della Lastruccia 3,
50019, Sesto Fiorentino, Firenze, Italy. E-mail: fabrizio.mani@uni.it
b
ICCOM CNR, via Madonna del Piano 10, 50019, Sesto Fiorentino,
Firenze, Italy

This journal is The Royal Society of Chemistry 2011

ment with the production of commercially valuable products.4


Turning carbon dioxide into a feedstock for producing useful
commodity chemicals in mild conditions would indeed circumvent most of the drawbacks of the energy consuming steps
of CO2 desorption, absorbent regeneration as well as of CO2
transportation and disposal in geological cavities, oceans or
elsewhere.
In a previous experimental study,5 we reported that the
absorption of CO2 by 0.8510.0 M NH3 (1.4618.1 wt%)
aqueous solutions occurs with high efciency and load capacity
producing solutions of bicarbonate, carbonate and carbamate
ammonium salts in a ratio which depends on the amount of
absorbed CO2 with respect to the concentration of free NH3
in solution, as inferred by 13 C NMR spectroscopy. The high
water solubility of the ammonium salts of each species, HCO3 - ,
NH2 CO2 - and CO3 2- , prevented the crystallization of any solid
compound at the end of the absorption experiments.
The reaction of gaseous CO2 and NH3 in dry conditions
at ambient temperature and under atmospheric pressure is
exoenthalpic and produces ammonium carbamate [see below,
reaction (4)]. However, the process is scarcely suited for practical
CCS applications because the low reaction rate of CO2 and NH3
in the gas phase causes a severe loss of NH3 and/or scarce
CO2 removal efciency, accompanied by the difcult removal of
solid ammonium carbamate from the absorbent reactor.6 On the
contrary, doing the reaction in liquid phase facilitates the process
Green Chem., 2011, 13, 12671274 | 1267

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

View Article Online

Fig. 1 Simplied ow diagram of the absorber-ltration cyclic conguration (A) and simplied sketch of the gaseous CO2 NH3 absorber (B).

and makes easy the separation of the solid from the solution,
thus allowing an efcient recycle of the unreacted scrubbing
solution. In summary, the CO2 capture by a wet method should
be highly preferable.
In this paper we report our results on CO2 capture by NH3
in ethanolwater solution as well as in anhydrous ethanol, 1propanol and N,N-dimethylformamide (DMF). The processes
result in the formation of solid mixtures of ammonium bicarbonate and ammonium carbamate or of pure ammonium carbamate
that contain all of the captured CO2 . The capture of CO2 has
been carried out at room pressure with chilled NH3 (273 K)
in order to minimize NH3 loss. Remarkably, the unreacted
ammonia solution, once separated by ltration from the solid
compounds, is entirely reclaimed into the absorbent reactor.
Ammonium bicarbonate is a market product with a variety
of uses. In particular, it has been used as a nitrogen fertilizer in
China for over 30 years.7 Recently, the use of both ammonium
bicarbonate and carbamate has been patented for the recovery of
freshwater from seawater by forward osmosis.8 Ammonium carbamate has been also tested as a NH3 generator for NOx abatement in diesel exhaust gases, to recover manganese from steelmaking plant slag and to soil remediation.9 More important,
ammonium carbamate is the intermediate for the production of
urea, the most used nitrogen fertilizer worldwide (more than 108
metric tons per year).10 The commercial production of urea is
based on substantially similar patented processes where carbon
dioxide is reacted with excess of ammonia (NH3 /CO2 molar
ratio up to 4) at high temperature (450500 K) and pressure
(150250 bar) to produce ammonium carbamate, which is then
dehydrated to urea.
Here we report our studies about the conversion of the solid
ammonium carbamate, obtained in the anhydrous process of
CO2 capture by NH3 , into urea. The process takes place in
relatively mild conditions, i.e. 393413 K and 14 bar (the
pressure generated by the thermal decomposition of ammonium
carbamate). The reaction was accomplished in a sealed vessel
contained ammonium carbamate in the presence of a transition
metal catalyst (11.5 wt%). Depending on both the reaction
parameters (temperature and time) and the catalyst nature,
1268 | Green Chem., 2011, 13, 12671274

conversion of carbamate into urea ranged between 17 and 54%.


Apart for the crucial use of the catalyst, the urea formation from
heating solid ammonium carbamate has been recognized for a
long time.11

Results and discussion


Formation of solid ammonium carbamate and ammonium
bicarbonate
Due to the low solubility of both ammonium carbamate and
bicarbonate in ethanol, some CO2 capture experiments were
carried out in ethanolwater ammonia solutions (275335 mL)
using three different NH3 concentrations (1.10, 2.06 and 2.72 M;
see Experimental). The absorption experiments were carried
out at 273 K using an home-built glass absorber4b immersed
into a thermostatted bath. The CO2 /N2 gas mixture (12% v/v)
simulating ue gas was owing at the bottom of the absorbent
apparatus through a sintered glass diffuser. During the absorption experiments the slurry was circulated with a peristaltic
pump in a closed loop between the absorber and the ltration
unit (Fig. 1A). In the latter, the solid was continuously separated
from the solution that was reclaimed back to the absorber.
In a different type of absorption experiments, gaseous CO2
and, separately, NH3 were continuously introduced through two
sintered glass diffusers in the absorber unit (Fig. 1B) containing
the anhydrous solvent (300 mL, ethanol, 1-propanol or DMF).
The gas exiting from the absorbent unit was dried and puried
from NH3 before being analysed with a gas chromatograph.
The equilibria which describe the reaction of carbon dioxide
with aqueous ammonia are:12
NH3 + CO2 + H2 O  NH4 + + HCO3 - , K eq(273) = 3.22 103 (1)
NH3 + HCO3 -  NH2 CO2 - + H2 O, K eq(273) = 7.30

(2)

NH3 + HCO3 -  CO3 2- + NH4 + , K eq(273) = 2.76 10-1

(3)

If an excess of NH3 is maintained in the system, the overall


reactions can be rewritten as
This journal is The Royal Society of Chemistry 2011

View Article Online

2NH3 + CO2  NH2 CO2 - + NH4 + , K eq(273) = 2.35 104

(4)

2NH3 + CO2 + H2 O  CO3 2- + 2NH4 + , K eq(273) = 8.89 102 (5)

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

The concentration of both NH2 CO2 - and CO3 2- decreases by


increasing CO2 absorption; meanwhile, a simultaneous increase
of bicarbonate occurs according to the equilibria (6) and (7)
CO3 2- + CO2 + H2 O  2HCO3 -

(6)

NH2 CO2 - + CO2 + 2H2 O  2HCO3 - + NH4 +

(7)

In general, a greater ammonia concentration increases the


efciency of CO2 removal, but lowers the loading capacity
and increases the loss of NH3 from the scrubbing solution.
In the NH3 /H2 OC2 H5 OH absorption experiments, no fresh
NH3 was added to the absorber to replace that consumed so
that the absorption efciency decreased with increasing of the
absorbed CO2 . The absorption experiments were stopped after
ca. 200 min. The average CO2 removal efciency in the entire
experiments is comprised between 78.5% (NH3 1.10 M) and
98.9% (NH3 2.72 M). The solids recovered at the end of the absorption experiments were different mixtures of NH4 HCO3 and
NH2 CO2 NH4 together with a smaller amount of (NH4 )2 CO3 .
Pure NH4 HCO3 or NH2 CO2 NH4 were never obtained at the end
of each experiment. In order to estimate the composition of the
solid obtained during the absorption process, 13 C NMR spectra
were recorded in D2 O and compared with those of standard
solutions of NH4 HCO3 and NH2 CO2 NH4 in the same solvent.
The results of three experiments at different NH3 concentrations
are reported in Table 1.
The maximum amount of solid (6.58 g after 184 min) and
the maximum concentration of carbamate in the solid mixture
(97.5% after 114 min) were obtained with the 2.06 M NH3
solution that represents the best compromise amongst several
opposite effects, as, for example, the amount of the species
in solution that increases with NH3 concentration and the
solubility of the different species that increases with increasing
waterethanol ratio that, in turn, increases with NH3 concentration. Furthermore, the amount of carbamate should increase
with NH3 concentration, but should decrease with increasing
of waterethanol ratio, according to the reverse of reaction
(2). Analogous experiments were carried out using other liquid mixtures, such as pyridine/water, dimethylsulfoxide/water,

DMF/water, dioxane/water, 2-propanol/ethanol/water, 1butanol/ethanol/water. In all of these experiments different


amounts and different ratios of the three ammonium salts were
obtained, without any selectivity.
In the experiments, aimed at obtaining pure NH2 CO2 NH4 ,
anhydrous operating conditions were sought in order to avoid
the competitive formation of HCO3 - and CO3 2- [reactions (1)
and (5)]. Gaseous NH3 and, separately, CO2 in molar ratio
comprised between 0.96 and 2.20 were continuously fed at
the bottom of the absorbent reactor (Fig. 1B) containing
300 mL of different organic solvents (ethanol, 1-propanol,
DMF) thermostatted at 273 K. The ow rate of the NH3 /CO2
mixture was the best compromise between the rate of solid
carbamate formation and the CO2 capture efciency. Increasing
the ow rate increases the amount of carbamate at the expense
of a reduced CO2 capture efciency and of an appreciable loss of
ammonia. Decreasing the ow rate has the opposite effect. The
duration of each experiment was xed at 8 h. Each experiment
was carried out at least in duplicate for any solvent and any
concentration used. The efciency of CO2 capture was always
greater than 93% (with a maximum value of 98% in DMF) by
running the reaction with a slight excess of NH3 (NH3 /CO2 ratio
2.102.20) with respect to the stoichiometry of the reaction (8)
2NH3(gas) + CO2(gas) NH2 CO2 NH4(solid)

(8)

The results of typical experiments carried out under different


operating conditions are reported in Table 2 and summarised in
Fig. 2. The slight differences obtained with the three solvents
are mainly due to the slight differences of NH3 /CO2 ratios
and, presumably, to the different solubility of either gas and
carbamate.
The conversion of NH3 into solid NH2 CO2 NH4 ranged
between 71.8 and 79.4% [Table 2, entries 3, 6, 9; on molar
scale, according to the stoichiometry of reaction (8)]. Absorption
experiments with ca. 1/1 NH3 /CO2 ratio (0.961.03) gave 93.6
96.9% (Table 2, entries 1,4,7) conversion of NH3 in crystalline
solids, but the CO2 capture efciency was reduced to 55
66%. Again, the search for the best compromise between CO2
capture efciency and NH3 conversion to pure NH2 CO2 NH4
asked for a NH3 /CO2 molar ratio of ca. 1.5 that resulted
in high CO2 capture (8590%) and NH3 conversion (9498%;
Table 2, entries 2, 5, 8) to crystalline NH2 CO2 NH4 . In all

Table 1 CO2 absorption efciency and composition of the solid mixture of ammonium carbamate, bicarbonate and carbonate obtained from
different aqueous ammoniaethanol solutions
Composition (%)f (n/n)
c/mol dm-3

H2 OC2 H5 OHa
(v/v)

1.10

0.0741

2.06

0.148

2.72

0.222

tb /min

CO2 absc (%)

CO2 /NH3 d
(n/n)

me /g

NH2 CO2 -

HCO3 -

CO3 2-

M g

solid/CO2 h
(n/n)

128
192
114
184
124
187

86.8
78.5
90.5
89.8
100
98.9

0.385
0.537
0.181
0.289
0.145
0.215

3.30
5.61
2.88
6.58
1.33
3.80

75.5
48.7
97.5
86.8
87.3
83.3

15.8
42.6
2.2
8.6
8.3
10.3

8.8
8.6
0.3
4.6
4.5
6.6

79.8
80.6
78.1
79.6
79.0
79.6

0.354
0.431
0.340
0.481
0.127
0.245

Volume ratio between aqueous ammonia and ethanol. b Absorption time. c Average absorption efciency during the absorption time. d Average
loading during the absorption time. e Mass of the solid mixture. f Average composition, on molar scale, of the solid mixtures recovered at the end of
any experiment. g Average molar mass computed from the mixture composition. h Molar ratio between the solid mixture and the absorbed CO2 .

This journal is The Royal Society of Chemistry 2011

Green Chem., 2011, 13, 12671274 | 1269

View Article Online

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

Table 2 CO2 absorption efciency and yield of NH3 conversion into solid ammonium carbamate at different operating conditions
Entry

solvent

NH3 /CO2 n/n

CO2 (abs) %

NH2 CO2 NH4 /NH3 % (n/n)

NH2 CO2 NH4 /CO2 % (n/n)

NH2 CO2 NH4 , m/g h-1

1
2
3
4
5
6
7
8
9

ethanol
ethanol
ethanol
1-propanol
1-propanol
1-propanol
DMF
DMF
DMF

1.03
1.59
2.17
0.96
1.51
2.20
0.96
1.58
2.10

66.4
89.7
93.1
54.6
88.1
94.5
63.4
85.0
98.0

93.6a
94.0a
77.0
96.6a
98.2a
71.8
96.9a
93.7a
79.4

74.7
83.1
89.7b
66.9
84.0
83.7b
82.1
85.9
85.2b

2.33
3.77
4.08
2.34
3.93
3.81
2.61
3.75
4.21

NH3 is the species in defect with respect to the stoichiometric ratio. b CO2 is the species in defect with respect to the stoichiometric ratio.

the reaction (9) is an oversimplication of the multi-phase and


multi-component equilibria occurring at high temperature and
pressure. The set of the main reactions are:
NH2 CO2 NH4 (s)  2NH3 (g) + CO2 (g)
2NH3 (g) + CO2 (g)  (NH2 )2 CO(l) + H2 O(l)
NH2 CO2 NH4 (s) + H2 O(l)  NH4 HCO3 (s) + NH3 (g)
NH2 CO2 NH4 (s) + 2H2 O(l) + CO2 (g)  2NH4 HCO3 (s)

Fig. 2 CO2 absorption efciency () and ammonia conversion into


carbamate (%) (---) as a function of the NH3 /CO2 molar ratio.

of these experiments, ammonium carbamate was obtained in


quite pure form, as inferred from the comparison of 13 C NMR
spectra in D2 O of the different experimental samples with that
of reagent grade commercial ammonium carbamate. In each
experiment, the unreacted solution was entirely recycled to the
absorbent apparatus. Remarkably, the loss of NH3 from the more
concentrated absorbent solution never exceeded 1.5% (on molar
basis) for the entire duration of each absorption experiment.
Conversion of ammonium carbamate into urea
The urea manufacture at industrial scale, typically an urea plant
produces 1500 tons per day, is carried out by feeding excess
ammonia and carbon dioxide to the synthesis reactor at 450
500 K, 150250 bar. The process involves the intermediate
formation of ammonium carbamate [reaction (8)] that is successively dehydrated to form urea
NH2 CO2 NH4 NH2 CONH2 + H2 O

(9)

While reaction (8) is fast and exothermic (DH =


-151 kJ mol-1 ), reaction (9) is slow and endothermic (DH =
32 kJ mol-1 )13 and does not go to the completion under
industrial processing conditions. As a matter of fact, the rather
small enthalpy value suggests that the reaction (9) should be
substantially right hand shifted at high temperature and the
limiting value of urea yield is dictated rather by the reaction
rate than the equilibrium value. Moreover, we can point out that
1270 | Green Chem., 2011, 13, 12671274

Therefore, the nal yield of urea is the result of the competition


between the conversion reactions of carbamate into urea and
its decomposition reactions to, mainly, bicarbonate. The above
equilibria proceed differently to each other at the same working
conditions. In general, the yield of the reaction is in the order 30
55% on NH3 basis (6070% on CO2 basis) and strongly depends
on reaction temperature, pressure, time and NH3 /CO2 ratio.
In order to circumvent the energy penalties affecting the
conventional industrial production of urea requesting high
working temperature and pressure, we looked for the possibility
to promote the dehydration of NH2 CONH4 by carrying out
the reaction in the presence of a transition metal catalyst.
Surprisingly, very few reports may be found in both scientic
and patent literature describing the use of a catalyst to bring
about the synthesis of urea and none of them involves the two
step procedure which is industrially used.14
In keeping with our expectations, heating ammonium carbamate contained in a sealed vessel (home-built stainless steel
airtight container) for 23 days at 393413 K (in one experiment
up to 433 K) in the presence of different catalysts (catalyst
loading: 1.01.5%, on molar basis, Table 3) resulted in an
increased production of urea conrming the possibility to
develop a catalytic synthesis of urea at relatively low pressure
because the maximum pressure generated by the thermal
decomposition of carbamate in the autoclave was 14 bar at
413 K. At the end of the reaction, 13 C NMR spectroscopy in
D2 O (see Experimental) showed that the solid mixture contained
variable amounts of unreacted carbamate (d = 165.84), urea
(d = 162.95) and carbonate/bicarbonate mixture (d = 163.3
164.3) depending on the operating conditions. No other product
was detected in the 13 C NMR spectra, in particular no trace of
biuret (d = 158.08) was found. Presumably, dimerization of urea
to biuret occurs only at temperature and pressure higher than
those occurring in our experimental conditions. Urea could be
This journal is The Royal Society of Chemistry 2011

View Article Online

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

Table 3 Catalysts and experimental conditions employed in the carbamate conversion into urea and yield of the reactionsa
Entry

Catalyst

% (mol)b

T/K

Time/day

Ureac (% mol)

1
2
3
4
5
6
7
8
9
10
11
12

copper
CuCl2 2H2 O
CuCl2 2H2 O
CuO
CuCO3 Cu(OH)2
ZnCl2
ZnCl2
NiCO3 2Ni(OH)2 H2 O
NiCO3 2Ni(OH)2 H2 O
RuCl3
MnCl2
FeCl3

15.3
1.00
1.00
1.00
0.97
2.50
1.25
1.12
1.00
1.26
1.54
1.00

393
393
413
393
403
403
413
413
433
413
393
393

3
3
3
3
2.5
2
3
3
3
2.5
3
3

50.3
48.1
53.6
53.7
51.0
32.0
45.0
25.9
31.0
33.0
21.0
17.3

a
c

Blank experiments up to 403 K and 3 days gave less than 3% yield. b The catalyst percentage is referred to the starting amount of carbamate.
Percentages are referred to the starting amount of carbamate.

recovered from the solid residue by heating the reaction mixture


(333 K) till to constant weight, which removed the unconverted
NH2 CO2 NH4 , the ammonium carbonate/bicarbonate mixture
and water as gaseous NH3 , CO2 and H2 O, leaving pure urea and
the catalyst. The purity of the product was checked by 13 C NMR
analysis and the yield of the reaction (9) could be determined
after subtracting the weight of the catalyst from the constant
weighted residue. Both gaseous NH3 and CO2 obtained from the
decomposition of the mixture of the unreacted carbamate and
of the by-product ammonium bicarbonate could be recovered
and recycled, whereas the catalyst could be separated from the
melt urea and reused.
A variety of reaction promoters in different ratio with respect
to carbamate, were tested and the most efcient ones are
reported in Table 3. The most efcient catalysts were copper(II)
and zinc(II) compounds and elemental copper. Under the same
operating conditions, manganese, iron, chromium and cobalt
compounds gave a conversion lower than 25%. RuCl3 lies
between the two groups of catalysts with a 33% of conversion.
The other tested metal compounds gave generally less than 5%
of urea. Blank experiments run without any catalyst gave less
than 3% of urea again conrming the catalytic activity of the
added metal salt. In general, increasing the catalyst loading did
not improve the yield and in most experiments a catalyst loading
of 11.5% (on molar scale) with respect to carbamate gave the
best results. In contrast, 15% loading of powdered elemental
copper was necessary to obtain the same results of the Cu(II)
compounds indicating that Cu(0) is a poor catalyst with respect
to copper(II) salts. Increasing both decomposition temperature
(Table 3, entries 2, 3 and 8, 9) and time (entries 6, 7) increased
the conversion.
The role of the metal catalyst in the reaction mechanism
responsible for the carbamate conversion to urea (eqn (9)) is,
at the present, uncertain. However, although unravelling the
mechanism of the catalytic reaction is out of our current interest,
some hypotheses seem plausible.
If one consider that (i) either hydrated or anhydrous metal
compounds have a similar effect and (ii) the amount of the
catalyst is 11.5% of the water formed by the reaction (9), it is
conceivable that the metal salts cannot simply shift the reaction
to the right acting as dehydrating agents. Furthermore, the
formation of metal(II)urea complexes that could also favour
This journal is The Royal Society of Chemistry 2011

the conversion reaction or the capacity of the metal(II) ions


to provide a platform to bring two molecules of NH3 in close
proximity to each other thus facilitating the reaction with CO2 ,
cannot explain the better performances of Cu(II) and Zn(II)
compared to Ni(II) and Co(II), for example, as all of the four
metal ions give ammonia15 and urea complexes.16 On these
assumptions, we may tentatively propose that the carbamate
conversion to urea can occur through the following steps [M(II)
stands for either Cu(II) or Zn(II)]
II)
2
NH 2CO 2 NH 4 heat
2NH 3 + CO 2 M(

[M(NH 3 )6 ]2+

CO

3
NH 2CONH 2 + H 2O + [M(NH 3 ) 4 ]2+
[M(NH3 )6 ]2+
(10)
2NH

The ammonia formed by the thermal decomposition of


ammonium carbamate may easily coordinate Cu2+ or Zn2+
ions forming the corresponding hexakis-amino complexes
[M(NH3 )6 ]2+ (I). Then, the reaction of CO2 with two adjacent
molecules of ammonia may well account for the metal-assisted
formation of one urea molecule as a dihapto-coordinated ligand
in the transient [M{k2 -N,N (NH2 )2 CO}(NH3 )4 ]2+ complex (II).
Water is also released along this reaction step. The urea complex,
once formed, may easily exchange urea with the excess of
ammonia restoring the more stable hexakis-amino complex (I),
possibly via the tetracoordinated [M(NH3 )4 ]2+ complex (III).
Fig. 3 illustrates this mechanistic hypothesis. If we taken for
granted this putative hypothesis, the greater efciency of Cu(II)
and Zn(II) with respect to the other tested metal ions could
be readily explained on the basis of the peculiar coordinating
properties of Cu(II) and Zn(II) ions that, in the presence of an
excess of several ligand molecules, including ammonia, form
six-coordinated complexes (I) containing four strongly and
two labile coordinated ligands.15 Thus, two cis disposed NH3
molecules of either [Cu(NH3 )6 ]2+ or [Zn(NH3 )6 ]2+ complexes
may easily react with CO2 yielding the urea complex II which
eventually undergoes substitution reactions where the coordinated urea molecule is easily replaced by two ammonia ligands.
The proposed reaction pathway could also explain the reduced
conversion yield as the metal(II)/carbamate ratio increases and
the poor performances exhibited by other transition metal salts
with respect to Zn2+ and Cu2+ ions. Indeed, the formation of sixcoordinated Cu(II) and Zn(II) complexes at high temperature
Green Chem., 2011, 13, 12671274 | 1271

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

View Article Online

Fig. 3 Proposed pathway of the M2+ catalysed reaction between


CO2 and two coordinated ammonia molecules (M = Cu, Zn). Urea
displacement from the coordination sphere by the stronger ligand NH3
may regenerate easily the active catalytic species.

requires a large excess of ammonia with respect to the metal(II)


ion while Cr(III), Fe(III), Mn(II), Co(II) and Ni(II), which
gave worse catalytic performances, form very stable and inert
octahedral six-coordinated ammonia complexes and, because
of this kinetic inertness, the ligand displacement step is hard to
be accomplished under the used working conditions.
Studies are in progress to substantiate this putative hypothesis
by investigating the specic role of amino complexes in the
catalytic decomposition of ammonium carbamate to urea.
Preliminary results indicate that catalytic tests carried out with
isolated amino complexes [M(NH3 )6 ]2+ may increase both the
decomposition rate of ammonium carbamate and the urea yield.

Experimental
All reagents were reagent grade. Ammonium carbamate,
ammonium bicarbonate, ethanol, 1-propanol and N,Ndimethylformamide (Sigma-Aldrich) were used as received.
Standard NH3 solution 15.2 M (Sigma-Aldrich) was used to
prepare the ethanol/water/ammonia solutions. Pure CO2 and
N2 used to simulate ue gas and pure NH3 were obtained by
Rivoira. Flow rates of N2 , CO2 and NH3 were measured with
gas mass ow meters (Aalborg) equipped with gas controllers
(Cole Parmer). The inlet and outlet CO2 concentrations in the
ue gas mixture were measured with a Varian CP-4900 gas
chromatograph calibrated with a 10% v/v CO2 /N2 reference
gas (Rivoira).
The cyclic absorptionltration device consisted of the absorber and the ltration units that are connected to each
other by means of a peristaltic pump (Masterex L/S) that
1272 | Green Chem., 2011, 13, 12671274

allows the absorbent slurry and the ltered solution to circulate


continuously in a closed loop between the absorber and the
ltration unit (Fig. 1A). The temperature of the absorbent
solution was kept constant at 273 K by a thermostatted water
bath (Julabo model F33-MC refrigerated bath).
To mimic ue gas, a gas mixture containing 12% (v/v) CO2 in
N2 , was continuously fed into the absorber through a sintered
glass diffuser (1640 mm pores) at the bottom of the absorbent
solution with a ow rate of 14 dm3 h-1 . The vent gas came
out from the top of the absorber. The outlet gas was dried
by owing in turn through a condenser cooled at 268 K, a
concentrated H2 SO4 solution and a gas purication tower lled
with P2 O5 , before being analysed with a gas chromatograph.
The gas chromatograph measured the percentage of the CO2
absorbed by the ammonia solutions at intervals of 10 min.
In the NH3 /H2 O/C2 H5 OH experiment, the absorber device
was a home-built glass cylinder with a diameter of 56 mm and
height 300 mm equipped with a thermometer and a combined
pH electrode and tted with three polyethylene disks threaded
on a 2 mm glass rod.4b This arrangement increases the liquid
turbulence therefore providing the reaction mixture with a
sufcient residence time. The absorbent solution was obtained
by mixing 270 mL of ethanol and, separately, 20.0, 40.0, 60.0 mL
of NH3 15.2 M, according to the different experiments. The
volumes of waterethanol solutions were 275, 295 and 335 mL
and the respective NH3 concentrations were 1.10 M, 2.06 M and
2.72 M. During all of the CO2 absorption experiments no new
NH3 was added to the absorber unit. Each experiment lasted
about 3 h.
In the experiments aimed at obtaining pure carbamate, both
CO2 (12% v/v in N2 ) and NH3 were simultaneously introduced
through two separate gas diffusers into 300 mL of the appropriate solvent (ethanol, 1-propanol or DMF) contained in the
thermostatted (273 K) absorber (Fig. 1B). The NH3 /CO2 ow
ratio was in the range 0.962.20 and each experiment lasted 8 h.
At the end of each experiment, the solid collected by the ltration
unit was washed with CO2 saturated ethanol and diethyl ether in
turn before being dried at room temperature with a ow of pure
CO2 to avoid the decomposition of either ammonium carbamate
or bicarbonate. The conversion of ammonium carbamate into
urea was carried out in a home built reactor that comprises an
internal TeonR container (the volume of the cylinder is 50 mL)
sealed by means of an airtight TeonR cap. The TeonR reactor is
contained in stainless steel vessel with a screw cap and equipped
with a pressure gauge. The reactor is heated at the appropriate
temperature (393413 K) by means of a silicone oil heating
bath (IKA HB4). In each experiment, the TeonR reactor is
charged with 8.00 g of ammonium carbamate intimately mixed
with the catalyst selected from a variety of oxides or salts
of chromium(III), manganese(II), iron(III), cobalt(II), nickel(II),
copper(II), zinc(II), ruthenium(III) and from pure Cr, Fe, Cu, and
Zn metals.
As an example, the decomposition of ammonium carbamate
in the presence of CuO is described below. The other experiments
were run similarly by replacing CuO with the appropriate
catalyst.
In a typical experiment, 8.0 g of ammonium carbamate
(0.103 mol) were carefully mixed with 0.080 g of CuO (1.01
10-3 mol) and heated to 393 K for three days. A small amount
This journal is The Royal Society of Chemistry 2011

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

View Article Online

(weighted) of the solid mixture recovered from the reactor at


the end of the conversion experiment was dissolved in D2 O
and analysed by 13 C NMR spectroscopy. The integration of the
signals (see below) due to HCO3 - /CO3 2- (d = 163.15), NH2 CO2 (d = 165.92) and (NH2 )2 CO (d = 162.92) allowed us to quantify
the molar ratio of the species. Additionally, the solid mixture
was heated in the air at 333 K to constant weight in order to
decompose the unreacted ammonium carbamate and the byproduct ammonium bicarbonate and to evaporate water. The
residue, subtracting the mass of the starting CuO and summed
to the mass of the sample used for NMR spectrum, weighted
3.32 g (0.0553 mol) and was identied as pure urea (53.7% of the
starting ammonium carbamate, on molar scale) by 13 C NMR
spectroscopy in D2 O solution (d = 162.92) upon comparison
with that of reagent grade urea. The 13 C NMR spectra were
obtained with a Bruker AvanceIII 400 spectrometer. Chemical
shifts are to high frequency relative to tetramethylsilane as
external standard at 0.00 ppm. CH3 CN was used as internal
reference (CH3 CN, d = 1.47). The standard pulse sequence with
proton decoupling and NOE suppression was used to acquire
the 13 C{1 H} with the following acquisition parameters: pulse
angle = 90.0 , at = 1.36 s, dl = 0 s, data points = 65 K, ns = 5002
103 . Increasing the acquisition time and/or the relaxation delay
(up to 60 s) does not produce substantial changes in the relative
peak areas of the CO carbon atoms. Normally, the integration
of 13 C NMR resonances does not allow reliable quantication
of species having carbon atoms in different environments, due to
the different spinlattice T1 relaxation time.17 In the species we
are dealing with, i.e. carbamate, bicarbonate, carbonate and urea
the CO carbons have a similar chemical environment, so that
they likely exhibit similar T1 . As a matter of fact, to substantiate
this assumption, we have carried out several 13 C NMR spectra
on standard solutions containing accurately weighted amounts
of ammonium carbamate, ammonium bicarbonate and urea
in different molar ratios and we have found a quantitative
relationship (maximum error 5%) between the relative peak
areas of the CO 13 C resonances and the known concentrations of
each species. The quantication method is therefore empirically
quite reliable likely reecting similarities of the relaxation rate
for similar carbons in carbamate, bicarbonate and urea.

Conclusions
The results here reported conrm the advantage of CO2 removal by NH3 in both aqueous and nonaqueous solutions in
terms of absorption efciency and loading capacity. In order
to circumvent the drawbacks of the energy consuming steps
associated with NH3 regeneration, its separation from CO2 and
CO2 sequestration deep underground or under oceans, we have
devised a procedure that transforms all of the captured CO2 into
commodity chemicals. In particular, the reaction of gaseous NH3
and CO2 in organic solvents (ethanol, 1-propanol and DMF)
produces solid ammonium carbamate in a quite pure form and
with a good yield. The best compromise between efciency of
CO2 capture (8593%) and yield of solid carbamate with respect
to NH3 (9398%) is obtained with about 1.5 NH3 /CO2 molar
ratio. The CO2 capture by the wet method here reported is to be
preferred to the reaction of CO2 and NH3 in gaseous phase due
This journal is The Royal Society of Chemistry 2011

to their low rate of reaction, to the severe loss of NH3 and/or to


the scarce CO2 removal efciency.
Even more interesting, the solid ammonium carbamate recovered from the CO2 capture is converted into urea under
mild conditions with respect to those used in the conventional
industrial processes for urea manufacturing. We have found that,
in the presence of 11.5% (on molar scale) of either copper(II)
or zinc(II) compounds, ammonium carbamate heated at 393
413 K in a sealed vessel without applying any external pressure
produces urea with variable yields (4554% with respect to
carbamate) depending on the operating conditions. The entire
procedure of ammonium carbamate and urea production here
described offers signicant advantages in terms of energy gain
with respect to the higher temperatures and pressures required
by the conventional industrial processes still representing an
efcient capture of anthropogenic CO2 ue emission. Even if
the present rate of catalytic formation of urea is too low for
industrial application, we have proved that some catalysts can
accelerate the carbamate dehydration to urea in rather mild
conditions. Studies are in progress to nd more efcient catalysts
and to optimise the process.

Acknowledgements
Financial support from MIUR (Rome, Italy) and Ente CRF
(Florence, Italy), through FLORENCE HYDROLAB Project, is
gratefully acknowledged.

References
1 Intergovernmental Panel on Climate Change Special Report
on Carbon Dioxide Capture and Storage, ed. B. Metz, O.
Davidson, H. C. de Coninck, M. Loos and L. A. Meyer, Cambridge University Press, Cambridge, United Kingdom and New
York, NY, USA, 442 pp, www.ipcc.ch; Chemistry & Energy;
The role of the chemical sciences in the European energy policy,
http://www.euchems.org/binaries/030882EnergyReport_tcm23118847.pdf.
2 J. E. Pelke, P. J. Concannon, D. B. Manley and B. E. Poling, Ind. Eng.
Chem. Res., 1992, 31, 22092215; A.-C. Yeh and H. Bai, Sci. Total
Environ., 1999, 228, 121133; K. Thomsen and P. Rasmussen, Chem.
Eng. Sci., 1999, 54, 17871802; A. Perez-Salado Kamps, R. Sing, B.
Rumpf and G. Maurer, J. Chem. Eng. Data, 2000, 45, 796809; H. P.
Huang, Y. Shi, W. Li and S. G. Chang, Energy Fuels, 2001, 15, 263
268; H. Huang, S.-G. Chang and T. Dorchak, Energy Fuels, 2002, 16,
904910; J.-W. Lee and R.-F. Li, Energy Convers. Manage., 2003, 44,
15351546; A. Corti and L. Lombardi, Int. J. Thermodynam., 2004,
7, 173181; Y.-F. Diao, X.-Y. Zheng, B.-S. He, C.-H. Chen and X.-C.
Xu, Energy Convers. Manage., 2004, 45, 22832296; K. P. Resnik, J.
T. Yeh and H. W. Pennline, Int. J. Environ. Techol. Manage., 2004, 4,
89104; L. Meng, S. Burris, H. Bui and W.-P. Pan, Anal. Chem., 2005,
77, 59475952; J. T. Yeh, K. P. Resnik, K. Rygle and H. W. Pennline,
Fuel Process. Technol., 2005, 86, 15331546; V. Darde, K. Thomsen,
W. J. M. van Well and E. H. Stenby, Int. J. Greenhouse Gas Control,
2010, 4, 131136.
3 CO2(emitted) represents the CO2 emitted in the atmosphere by burning
fossil fuels necessary for the production of all the forms of energy
(electrical, thermal, mechanical) required for the entire cycle of CO2
removal.
4 (a) F. Mani, M. Peruzzini and P. Stoppioni, IT FI2007A000199, 2007;
(b) F. Mani, M. Peruzzini and P. Stoppioni, Energy Fuels, 2008, 22,
17141719; (c) F. Mani, M. Peruzzini and F. Barzagli, ChemSusChem,
2008, 1, 228235.
5 F. Mani, M. Peruzzini and P. Stoppioni, Green Chem., 2006, 8, 995
1000.
6 X. Li, E. Hagaman, C. Tsouris and J. W. Lee, Energy Fuels, 2003, 17,
6974.

Green Chem., 2011, 13, 12671274 | 1273

Published on 22 March 2011. Downloaded by Universiti Teknologi Petronas on 13/08/2015 08:22:29.

View Article Online

7 C.-K. Li and R.-Y. Chen, Fert. Res., 1980, 1, 125136.


8 R. L. McGinnis, US 2001-265745 20010201; J. R. McCutcheon, R.
L. McGinnis and M. Elimelech, Desalination, 2005, 174, 111.
9 I. H. Warren and E. A. Devuyst, Trans. Soc. Mining Eng., 1972,
252, 388391; N. Breuer, C. Schiller and C. Oediger, FR 2003-5411
20030502.
10 Urea:
2010
World
Market
Outlook
and
Forecast
(http://www.mcgroup.co.uk/).
11 A. I. Bazarov, J. Prakt. Chem., 1870, 2, 283312.
12 B. Rumpf, F. Weyrich and G. Maurer, Ind. Eng. Chem. Res., 1998,
37, 29832995; O. Brettschneider, R. Thiele, R. Faber, H. Thielert
and G. Wozny, Sep. Purif. Technol., 2004, 39, 139159.
13 Petrochemical Processes: Synthesis-gas Derivatives and Major Hydrocarbons, A. Chauvel and G. Lefebvre, Editions Technip, Paris,
1989.

1274 | Green Chem., 2011, 13, 12671274

14 (a) J.-J. Herman and A. Lecloux, Catalytic synthesis of urea from


carbon monoxide and amine compound, US Patent 4801744, Jan.
31, 1989; (b) C. Bruneau and P. H. Dixneuf, J. Mol. Catal., 1992, 74,
97107.
15 F. A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry,
5th edn Wiley, New York, 1988; B. J. Hathaway, in Comprehensive
Coordination Chemistry, ed. G. Wilkinson, Pergamon Press, Oxford,
1987, vol. 5, ch. 53, p. 533; R. H. Prince, in Comprehensive
Coordination Chemistry, ed. G. Wilkinson, Pergamon Press, Oxford,
1987, vol. 5, ch 56.1, p. 925.
16 T. Theophanides and P. D. Harvey, Coord. Chem. Rev., 1987, 76,
237264.
17 E. Breitmaier and W. Voelter, Carbon-13 NMR Spectroscopy, 3rd
edn, VCH, Weinheim, Germany, 1990; R. J. Hook, Ind. Eng. Chem.
Res., 1997, 36, 17791790.

This journal is The Royal Society of Chemistry 2011

Potrebbero piacerti anche