Vous êtes sur la page 1sur 7

Energy

Procedia
Energy Procedia 4 (2011) 817–823
Energy Procedia 00 (2010) 000–000
www.elsevier.com/locate/procedia
www.elsevier.com/locate/XXX

GHGT-10

Designing small-molecule catalysts for CO2 capture


Sergio E. Wong, Edmond Y. Lau, Heather J. Kulik, Joseph H. Satcher, Carlos Valdez,
Marcus Worsely, Felice C. Lightstone and Roger Aines
Lawrence Livermore National Laboratory, 7000 East Ave, Livermore, CA 94550, USA
Elsevier use only: Received date here; revised date here; accepted date here

Abstract

One method for CO2 capture is to dissolve CO2 in water to form carbonic acid. This reaction
(CO2 +H20 → H2CO3(aq)) is remarkably slow but is catalyzed in biological systems by an enzyme
called carbonic anhydrase (CA). The catalyzed reaction is diffusion limited and occurs at near
neutral pH. The enzyme catalytic center is composed of a Zn(II) ion that is coordinated by 3
histidine residues and an axial water/hydroxyl group. A nucleophilic attack by the hydroxyl
group on the CO2 molecule is the first step in the reaction mechanism. Cu(II), Hg(II), Cd(II),
Ni(II), Co(II) and Mn(II) can bind the CA binding site and substitute the zinc ion; however, only
Co(II) yields rates comparable to Zn(II). Unfortunately, an enzyme, such as carbonic anhydrase,
is not amenable for industrial applications where a wide range of physico-chemical conditions
exist. Enzymes are vulnerable to large pressures, high temperature, and high ionic strength.
Efforts to isolate the key structural features responsible for catalysis led to the development of
small-molecule mimetics of the CA catalytic site. These mimetics can, in turn, be used as
catalyst for CO2 sequestration. Two of the fastest catalyst are the cyclic molecules: 1,4,7,10-
tetraazacyclododedacane and 1,5,9-triazacyclododedacane (both complexed with a Zn(II) ion).
Nitrogen atoms in these cyclic molecules mimic the imidazole nitrogens of the CA active site.
It is possible to examine the energetics of these compounds using transition state theory for the
purposes of designing more efficient catalysts. Transition state theory predicts that the reaction
rate constant is proportional to exp(-Ea/kT), where Ea denotes the activation energy, k the
Boltzmann constant, and T the temperature of the reaction. The activation energy is the
energetic cost of forming the reaction transition state from the reactants. Ab initio calculations
can yield the activation energy, which can, in principle, be used as a design metric for more
efficient catalysts.
Using this approach, the difference in kinetic rate constant between the tetra- and tri-aza
dodecane catalysts can be determined. Furthermore, the rates of the corresponding Co(II)
catalysts were explored. Our data suggests this is a viable method for the design of inorganic
small-molecule catalysts.

doi:10.1016/j.egypro.2011.01.124
818 S.E. Wong et al. / Energy Procedia 4 (2011) 817–823
2 Author name / Energy Procedia 00 (2010) 000–000


©
c 2010
2011 Elsevier
PublishedLtd.
by All rightsLtd.
Elsevier reserved
Open access under CC BY-NC-ND license.

Keywords:CO2 cycle, CO2 absorption, catalyst

Introduction

This report focuses on the use of catalysts that aid CO2 absorption in industrial steams1. CO2 gas
reacts with water to form carbonic acid, which is far more soluble than CO2 on its own2 in water.
The reaction with water, however, is very slow, and this in turn presents a problem for industrial
applications of gas separation3-5.
An alternative is to use a catalyst to accelerate CO2 hydration to carbonic acid. Such catalysts
already exist and come from the field of biology. They were designed to mimic an enzyme,
Carbonic Anhydrase II (CAII) 6, that catalyzes CO2 hydration at diffusion limited rates7. These
small-molecule CAII mimetics make a significant improvement over the uncatalyzed CO2
hydration kinetics. Examples include 1,4,7,10-tetraazacyclododecane8 and 1,5,9-
triazacyclododecane9.

Figure 1. Structure of 1,4,7,10-tetraazacyclodedecane (N4) and 1,5,9-


triaazacyclododecane (N3). The double positively charged metal may be zinc
or cobalt in the calculations.

They involve a zinc metal that is chelated in a


tetrahedral geometry. One of the chelating ligands
is a water/hydroxyl species that is responsible for
the nucleophilic attack on CO2 (Reaction 1 in Figure
2). Of all other metals tested for CAII activity, only
Co2+ is able to substitute Zn2+ and retain a
comparable rate (kcat/Km ~8.7 E 7 M-1s-1 for Zn2+ vs
~8.8 E 7 M-1s-1 for Co2+)10.

Figure 2. Proposed mechanism for CO2 hydration

This ab initio study used B3LYP density functional calculations examine CO2 hydration as
catalyzed by 1,4,7,10-tetraazacyclododedacane and 1,5,9-triazacyclododedacane (N4 and N3 in
Fig. 1). We examine how switching from Zn2+ to Co2+ in these catalysts affect their activation
energies. The data suggests that N3-Co2+ and N3-Zn2+ should have nearly equivalent kinetics.
Interestingly, the data also suggests N4-Co2+ has a higher activation energy than N3-Zn2+.
S.E. Wong et al. / Energy Procedia 4 (2011) 817–823 819
Author name / Energy Procedia 00 (2010) 000–000 3

Methods

Density functional theory (DFT) calculations were performed to study the reaction pathway of
CO2 hydration as mediated by the carbonic anhydrase mimics N4 and N3 (see Fig. 1) with either
Co2+ or Zn2+. Since cobalt in the CAII form of carbonic anhydrase is experimentally known to
be a high spin quartet (S=3/2)11-12, calculations on the cobalt-containing mimics are also carried
out with a fixed quartet multiplicity. Optimizations were performed for the free reactants,
encounter complex and transition states of the nucleophilic attack on the CO2 (see Fig. 2). The
starting reagents included a free CO2 molecule and either N3 or N4 with an OH- ligand in the
proximal axial position. The presence of the negatively charged axial ligand and positively
charged metal gives our starting complexes a net charge of +1.
DFT calculations were performed using the hybrid functional B3LYP within the Gaussian 03
suite of programs13. Geometry optimizations were performed using the 6-31G* basis-set and a
conductor-like polarizable continuum model (CPCM)14-15 to approximate solvent effects. Single
point energies were calculated the B3LYP/6-311+G* level of theory on the geometries from the
optimizations with a smaller basis set.

Results and discussion

Zn2+ complexes of 1,5,9-triazacyclododecane (N3) and 1,4,7,10-tetraazacyclododecane (N4)


hydrate CO2 to carbonic acid by a mechanism believed to mimic CAII. We considered the
nucleophilic attack by the catalyst on the CO2 carbon (reaction 1 in Figure 1). Kinetic rate data8-
9
and estimates of activation energies16 are available for the catalysis of CO2 hydration by N3-
Zn2+ and N4-Zn2+ in the gas phase as well as the condensed phase. Here, we focused on the
condensed phase as that is the relevant environment for industrial applications.

As a first metric
for validation, we
note that the
geometries
obtained for the
Zn2+ containing
catalysts are
comparable to
crystallographic
coordinates of the
catalysts17-18
(Tables 3 and 4).
The N-Zn
distances are
nearly identical,
Figure 3. Transition state geometries for the Zn containing catalysts
820 S.E. Wong et al. / Energy Procedia 4 (2011) 817–823
4 Author name / Energy Procedia 00 (2010) 000–000

and the angles between atoms are fairly close (within about 5 degrees of the crystal structure).
The crystallographic unit cell included 3 unique molecules that interact with each other in a
hydrogen bond network that join the oxygen atoms that chelate the metal (3 oxygens, one from
each catalyst). This is likely the reason for the apparent disagreement with experiment. We also
include the geometries calculated in gas phase by Brauer et al16 on these complex transition
states.

Table1. Calculated energies of the encounter complex (EC), transition state (TS), product 1 (P1)
and product 2 (P2) for the 1,4,7,10-tetraazacyclododecane (N4) and 1,5,9-triazacyclododecane
(N3) catalysts at the B3LYP/6-311+G(d)/CPCM//B3LYP/6-31G(d)/CPCM. All energies are
relative to the free reactants.
Molecule/Metal EC 'E TS 'E P1 'E P2 'E
(kcal/mol) (kcal/mol) (kcal/mol) (kcal/mol)

N4/Zn CPCM 1.5 1.8 -11.7 -20.5

N3/Zn CPCM 2.6 3.3 -10.6 -8.9

N4/Co CPCM 3.0 3.2 -11.6 -12.3

N3/Co CPCM 2.1 3.5 -8.9 -8.8

Table 2. Transition state geometries resulting from gas phase and CPCM optimizations of Zn2+
and Co2+complexes.
Zn2+ Co2+
Source Catalyst O1-C O2-Zn O1-Zn O1-C O2-Co O1-Co
Brauer et al N4 1.716 2.579 1.936 - - -
B3LYP/6-31G(d) N4 2.13 3.29 1.90 2.20 3.75 1.90
CPCM
Brauer et al N3 1.754 2.384 1.934 - - -
B3LYP/6-31G(d) N3 2.10 3.47 1.90 2.08 3.47 1.89
CPCM

Table 3. Comparison of experimental and calculated geometric data of azacyclododecane


complexes. N(1)-N(3) label the nitrogen atoms on the cycloamine sequentially.
Experiment17 B3LYP/6-
31G(d)/PCM
Zn-O 1.944 1.850
Zn-N(1) 2.004 2.046
Zn-N(2) 2.007 2.048
Zn-N(3) 2.042 2.036
N(1)-Zn- 104.3 104.9
S.E. Wong et al. / Energy Procedia 4 (2011) 817–823 821
Author name / Energy Procedia 00 (2010) 000–000 5

N(2)
N(1)-Zn-O 127.2 116.5
N(2)-Zn-O 109.2 116.5
N(1)-Zn- 105.2 105.8
N(3)
N(2)-Zn- 107.0 105.8
N(3)
N(3)-Zn-O 102.5 106.2

Table 4. Calculated geometric properties of N4-Zn2+ compared with crystallographic data18.


Atom numbers refer to Figure 3.
Experiment 18 B3LYP/6-
(N4/carbonate) 31+G(d)/CPCM

Zn-O 1.950 1.875


Zn-N(1) 2.144 2.179
Zn-N(2) 2.1773 2.186
Zn-N(3) 2.154 2.175
Zn-N(4) 2.111 2.185
O-Zn-N(1) 113.3 117.8
O-Zn-N(2) 103.2 113.4
O-Zn-N(3) 108.9 109.1
O-Zn-N(4) 119.1 113.8
N(1)-Zn-N(4) 82.6 80.8
N(2)-Zn-N(4) 137.6 132.7
N(3)-Zn-N(4) 82.9 80.9
N(1)-Zn-N(3) 137.3 133.1
N(1)-Zn-N(2) 82.4 80.7
N(2)-Zn-N(3) 81.9 80.9

Calculations of N3 and N4 were also performed with Co2+ as the metal center. While cobalt
mutants of CAII have been experimentally studied, no kinetic data exists for the N3-Co2+
catalyst. Table 1 shows activation energies (transition state energies) calculated in this work for
the Co2+-containing N3 and N4 mimetics. Activation energies for N3-Zn2+ and N3-Co2+ are
nearly the same (3.3 versus 3.5 kcal/mol, respectively) which suggests N3-Co2+ likely has
comparable catalytic efficiency as N3-Zn2+. The same comparison with the N4 species (Co2+ vs
Zn2+) suggests the zinc species, not the cobalt one, will be fastest.

Conclusion

DFT/CPCM calculations were used to study the catalyzed reaction mechanism of CO2
hydration. We examined the reaction energetic of the N3 vs N4 catalyst and the effect of using
cobalt, instead of zinc, as the metal center. The results show qualitative agreement with
822 S.E. Wong et al. / Energy Procedia 4 (2011) 817–823
6 Author name / Energy Procedia 00 (2010) 000–000

experiment in two ways: 1) they show that the N4-Zn2+ has a lower transition state energy (and
thus is a more efficient catalyst) than the N3-Zn2+ catalyst and 2) optimized structures agree with
crystallographic coordinates for the Zinc-bearing catalysts show. Substituting Cobalt for Zinc as
the metal-center yields comparable activation energies for N3-Co2+ and N3-Zn2+ which suggests
their catalytic efficiency is likely similar. An interesting contrast with the Zinc-bearing
compounds, is that our calculations yield nearly the same activation energy for N4-Co2+ and N3-
Co2+ which is not the case for the N4-Zn2+/N3-Zn2+ pair.

Acknowledgements
We thank the Laboratory Directed Research and Development Program at Lawrence Livermore
National Laboratory for funding 10-ERD-035 and Livermore Computing for the computer time.
This work was performed under the auspices of the U.S. Department of Energy by Lawrence
Livermore National Laboratory under Contract DE-AC52-07NA27344. Release # LLNL-PROC-
451972.

REFERENCES

1. Pachauri R, Reisinger A. Climate Change 2007: Synthesis Report. Geneva, Switzerland: Intergovernmental
Panel on Climate Change;2007.
2. Figueroa J, Fout T, Plasynski S, Mclvried H, Srivastava R. Advances in CO2 capture technology - The U.S.
Department of Energy's Carbon Sequestration Program. International Journal of Greenhouse Gas Control.
2008;2:9-20.
3. Stolaroff J. Carbon dioxide capture from atmopheric air using sodium hydroxide spray. Environmental
Science and Technology. 2008;4(8):2728-2735.
4. Stolaroff J. Capturing CO2 from ambient air, a feasibility assessment. Pittsburgh, PA, Carnegie Mellon
University; 2006.
5. Cullinane JT, Rochelle GT. Kinetics of carbon dioxide absorption into aqueous potassium carbonate and
piperazine. Ind Eng Chem Res. Apr 12 2006;45(8):2531-2545.
6. Khalifah R. The carbon dioxide hydration activity of carbonic anhydrase. I. Stop-flow kinetic studies on the
native human isoenzymes B and C. Journal of Biological Chemistry. 1971;246(8):2561-2573.
7. Silverman DN, Tu CK, Roessler N. Diffusion-Limited Exchange of O-18 between Co2 and Water in Red-
Cell Suspensions. Resp Physiol. 1981;44(3):285-298.
8. Zhang XP, Vaneldik R, Koike T, Kimura E. Kinetics and Mechanism of the Hydration of Co2 and
Dehydration of Hco3- Catalyzed by a Zn(Ii) Complex of 1,5,9-Triazacyclododecane as a Model for
Carbonic-Anhydrase. Inorg Chem. Dec 8 1993;32(25):5749-5755.
9. Zhang XP, Vaneldik R. A Functional-Model for Carbonic-Anhydrase - Thermodynamic and Kinetic-Study
of a Tetraazacyclododecane Complex of Zinc(Ii). Inorg Chem. Oct 25 1995;34(22):5606-5614.
10. Kogut K, Rowlett R. A comparison of the mechanisms of CO2 hydration by native and Co2+-substitued
carbonic anhydrase. Journal of Biological Chemistry. 1987;262:16417-16424.
11. Haffner P, Colema J. High Spin and Low Spin Forms of Co(II) Carbonic Anhydrase. The Journal of
Biological Chemistry. 1973;248(19):6630-6636.
12. Taylor J, Mushak P, Colema J. Electron Spin Resonance Studies of Carbonic Anhydrase: Transition Metal
Ions and Spin-Labeled Sulfonamides. Proc Natl Acad Sci U S A. 1979;67:1410-1416.
13. Gaussian 03, Revision C.02 [computer program]. Walingford, CT: Gaussian Inc; 2004.
14. Cances E, Mennucci B, Tomasi J. A new integral equation formalism for the polarizable continuum model:
Theoretical background and applications to isotropic and anisotropic dielectrics. J Chem Phys. Aug 22
1997;107(8):3032-3041.
15. Cossi M, Barone V. Analytical second derivatives of the free energy in solution by polarizable continuum
models. J Chem Phys. Oct 15 1998;109(15):6246-6254.
16. Brauer M, Perez-Lustres JL, Weston J, Anders E. Quantitative reactivity model for the hydration of carbon
dioxide by biomimetic zinc complexes. Inorg Chem. Mar 25 2002;41(6):1454-1463.
S.E. Wong et al. / Energy Procedia 4 (2011) 817–823 823
Author name / Energy Procedia 00 (2010) 000–000 7

17. Kimura E, Shiota T, Koike T, Shiro M, Kodama M. A Zinc(Ii) Complex of 1,5,9-Triazacyclododecane


([12]Anen3) as a Model for Carbonic-Anhydrase. J Am Chem Soc. Jul 18 1990;112(15):5805-5811.
18. Schrodt A, Neubrand A, van Eldik R. Fixation of CO 2 by Zinc(II) Chelates i nAlcoholic Medium. X-ray
Structures of {[Zn(cyclen)]3(m3-CO3)}(ClO4)4 and [Zn(cyclen)EtOH](ClO4)2. Inorg Chem. 1997;36:4579-
4584.

Vous aimerez peut-être aussi