Eric E. Benson, Clifford P. Kubiak*, Aaron J. Sathrum and Jonathan M. Smieja
Department of Chemistry and Biochemistry, University of California San Diego, 9500 Gilman Drive MC 0358, La Jolla, CA 92093, USA. E-mail: ckubiak@ucsd.edu; Tel: 1-858-822-2477
First published on 24th October 2008
Research in the field of catalytic reduction of carbon dioxide to liquid fuels has grown rapidly in the past few decades. This is due to the increasing amount of carbon dioxide in the atmosphere and a steady climb in global fuel demand. This tutorial review will present much of the significant work that has been done in the field of electrocatalytic and homogeneous reduction of carbon dioxide over the past three decades. It will then extend the discussion to the important conclusions from previous work and recommendations for future directions to develop a catalytic system that will convert carbon dioxide to liquid fuels with high efficiencies.
Eric Benson, Cliff Kubiak, Aaron Sathrum, Jon Smieja, and Bear | Eric Benson (left) received his BS degree in chemistry from the University of Arizona in 2005. His research is on biomimetic catalysts for the reduction of CO2. |
Cliff Kubiak received the Sc. B. degree with honors in chemistry from Brown University in 1975. He was awarded the PhD (1980) in inorganic chemistry by the University of Rochester, where he worked with Richard Eisenberg. He was a postdoctoral associate with Mark S. Wrighton at M. I. T. Kubiak was a faculty member at Purdue University from 1982–1998. He moved to UCSD in 1998 as Harold C. Urey Professor. His current research interests include ultra-fast electron transfer and molecular electronics, as well as atom transfer of organometallic complexes focusing on the homogeneous reduction of CO2 to liquid fuels. |
Aaron Sathrum (third from left) received a BS in electrical engineering and materials science from UC Davis in 2002. He is currently researching semiconductor photoelectrochemistry. |
Jon Smieja (right) received his BS degree in chemistry from the University of St. Thomas (MN) in 2005. His research is in ligand and catalyst design for the homogeneous reduction of CO2. |
CO2 + 2H+ + 2e−→ CO + H2O E0 = −0.53 V | (E1) |
CO2 + 2H+ + 2e−→ HCO2H E0 = −0.61 V | (E2) |
CO2 + 4H+ + 4e−→ HCHO + H2O E0 = −0.48 V | (E3) |
CO2 + 6H+ + 6e−→ CH3OH + H2O E0 = −0.38 V | (E4) |
CO2 + 8H+ + 8e−→ CH4 + 2H2O E0 = −0.24 V | (E5) |
CO2 + e−→ CO2˙− E0 = −1.90 V | (E6) |
In the general sense, electrocatalysts are electron transfer agents that ideally operate near the thermodynamic potential of the reaction to be driven, E0(products/substrates). Direct electrochemical reduction of carbon dioxide on most electrode surfaces requires large overvoltages which consequently lowers the conversion efficiency. The overvoltage can be considered to be the difference between the applied electrode potential, Vapplied, and E0(products/substrates), at a given current density. Both thermodynamic and kinetic considerations are important here. Clearly, in order to minimize overvoltages, catalysts need to be developed that have formal potentials, E0(Catn+/0) well matched to E0(products/substrates), and appreciable rate constants, kcat, for the chemical reduction of substrates to products at this potential. In addition, the heterogeneous rate constant, kh, for reduction of the electrocatalyst at the electrode must be high for Vapplied near E0(Catn+/0). A general approach for an electrocatalytic system is given in Scheme 1.
Scheme 1 Electrocatalysis with electron source. |
Reaction rates for these processes can be estimated from the steady-state limiting current in cyclic voltammetry, or by rotating disk voltammetry studies of the heterogeneous electron transfer kinetics. Identification of electrocatalytic activity can be seen easily in cyclic voltammetry (CV) (Fig. 1). In a CV under a dry inert atmosphere, an electrocatalyst should show a reversible redox couple. Upon addition of CO2, the diffusion limited current should increase significantly, while the potential shifts anodically, and the reversibility in the return oxidation wave is lost due to the chemical reaction between CO2 and the electrocatalyst. Electrocatalysts offer critical solutions to lowering the overpotentials, improving selectivity, and increasing the reaction kinetics of carbon dioxide conversion.
Fig. 1 Example cyclic voltammogram (CV) under (a) N2 and (b) CO2. Under a CO2 environment is readily observed: (1) anodic potential shift, (2) large increase in current, (3) non-reversible waveform. Figure reproduced from ref. 11. Copyright 1986 J. Amer. Chem. Soc. USA. |
Before we review the important studies in the electrocatalytic reduction of CO2, we must consider several molecular chemistry studies of the reactivity of CO2 toward transition metal complexes that provide important insights concerning the activation and reduction of CO2. Aresta and Nobile first published the crystal structure of CO2 bound to a transition metal complex in 1975 and reported an η2-bidentate binding mode involving the carbon atom and one oxygen atom, with significant bending in the CO2 structure.5 Another important study came in 1981 when Darensbourg and co-workers reported that anionic group 6B metal hydrides would react readily with CO2 to form the metal formates.6 This reaction proceeds according to eqn (7).
HM(CO)5− + CO2↔ HC(O)OM(CO)5− | (E7) |
In subsequent years, many reports of apparently homogeneous electrocatalysts for the reduction of CO2 appeared in the literature. We have divided the catalysts into three major categories that depend on ligand type: (1) metal catalysts with macrocyclic ligands, (2) metal catalysts with bipyridine ligands, and (3) metal catalysts with phosphine ligands. These seminal studies are summarized in the following subsections. It is important to note that this review is concerned with the electrocatalytic and homogeneous reduction of carbon dioxide and therefore may omit other types of carbon dioxide activation such as photocatalytic reduction, heterogeneous catalysis, and various others.
The first reported transition metal catalysts with high current efficiencies and turnover numbers were demonstrated by Eisenberg and co-workers in 1980.9 In this work, tetraazomacrocyclic complexes of cobalt and nickel were employed as shown in Fig. 2. These complexes were able to reduce carbon dioxide to carbon monoxide or a combination of CO and molecular hydrogen at potentials ranging from −1.3 to −1.6 V vs. SCE. These catalysts were also able to provide high current efficiencies, up to 98% (complex 3), but displayed low turnover frequencies between 2 and 9 per hour at 23 °C.
Fig. 2 Eisenberg catalysts for the reduction of CO2 to CO.9 All complexes (1–5) were effective for the reduction with varying degrees of success. System suffered from a requirement of high overpotentials for the reduction as well as the coincidental and competing production of H2. |
Sauvage and co-workers have extensively studied the reaction of CO2 with NiII(cyclam) complexes.10–12 The Sauvage complexes are extremely stable, highly selective, and can show faradaic efficiencies up to 96% in the production of CO at −0.86 V vs. SCE, even under aqueous conditions. The ligand geometry allows for a highly accessible metal center. It was noted that unsaturated or open chain complexes of similar moieties showed poor catalytic activity. The nickel macrocycle complexes were shown to be very sensitive to the pH and required an Hg electrode surface to turnover. The anion in the supporting electrolyte was shown to affect the selectivity, with KNO3 and KClO4 showing the fastest observed rates. A subsequent study on a binuclear transition-metal centered nickel complex using bicyclam showed similar reactivity towards CO2 but no coupling products were observed. As shown in later studies by Balazs the high electrocatalytic activity originates with Ni(cyclam)+ strongly absorbed on a mercury electrode surface.13 It was also found that CO limits the long-term stability of the catalyst in an unstirred solution by the deposition of an insoluble precipitate believed to be Ni(cyclam)(CO).
Iron(0) porphyrins were reported by the Savéant group in 1991 to reduce CO2 to CO at −1.8 V vs. SCE in DMF.14 However the porphyrins were unstable and degraded after a few catalytic cycles. With the addition of a hard electrophile, such as Mg2+, the stability and reactivity of the catalyst improved noticeably. It is believed through mechanistic studies that the Mg2+ ion assists in the breaking of the CO2 bound to iron creating Fe(II)CO and MgCO3. This is an example where the electron rich iron center initiates the reduction while the Lewis acidic Mg2+ helps complete the process, by sequestering an oxygen atom in MgCO3.
Later Savéant found that iron(0) porphyrins (Fig. 3) were able to catalyze the reduction of CO2 to CO in the presence of weak Brønsted acids.15 For the system where weak Brønsted acids such as 1-propanol, 2-pyrrolidine, and CF3CH2OH were added, Savéant found that catalysis was significantly improved in terms of both the efficiency and lifetime, without significant formation of H2. They were also able to reach turnover numbers as high as 350 h−1 at a catalyst decay rate of 1% per catalytic cycle. This system does, however, require reduction potentials that are still too negative for practical use (approximately −1.5 V vs. SCE) and the necessity of a mercury electrode, which we have noted above can display significant activity in the reduction of CO2sans catalyst.
Fig. 3 Savéant iron porphyrin catalyst structure shown to reduce CO2 to CO in the presence of weak Brønsted acids at a potential of −1.5 V vs. SCE.15 Porphyrins of this nature reduce CO2 to CO with high turnover frequency and low catalyst degradation, but require potentials too high for practical applications. |
In 2002 Fujita and co-workers published their findings on cobalt and iron corroles and their ability to catalyze the reduction of CO2 to CO.16 These complexes of the structure shown in Fig. 4 displayed a catalytic current with onset at approximately −1.7 V vs. SCE in the presence of CO2. The majority of analytical studies of the products for this reaction were done during photochemical reduction, but it was found that the major product of reduction was CO. Photochemical reduction of the complexes was achieved in deoxygenated solvents containing “sacrificial” triethylamine (Et3N) reducing agent and p-terphenyl (TP) as a sensitizer. The catalysts would produce varying amounts of CO and H2 (depending on catalyst) for up to 10 hours. Problems with this catalyst system include high overpotentials for reduction of CO2 and catalyst breakdown after long periods of irradiation.
It was later reported by Tanaka and co-workers that bipyridine complexes of ruthenium could catalyze the reduction of CO2.18 Ru(bipy)(CO)22+ and Ru(bipy)(CO)Cl+ were found to electrocatalytically reduce CO2 to CO, H2, and HCOO−. Both complexes were shown to reduce CO2 at −1.40 V vs. SCE. Following a two electron reduction of the complex, CO is lost to form a five coordinate neutral complex. In the presence of CO2 the complex forms an η1-CO2 adduct of Ru(0). This species can also be formed by addition of two equivalents of OH− to Ru(bipy)(CO)22+. Addition of a proton forms the LRu(CO)(COOH) species which under acidic conditions (pH 6.0) gains another proton to lose water and regenerate the catalyst as shown in Scheme 2. Under basic conditions (pH 9.5) the catalyst may undergo a two electron reduction with the participation of a proton to create HCOO− and regenerate the five coordinate Ru(0) complex. Even with the limitations of this system, such as low turnover numbers and low selectivity, it helped to elucidate several of the key intermediates in the reduction of CO2.
Scheme 2 Proposed mechanism for the Tanaka catalyst. |
In a similar system studied by the Meyer group, it was found that 2,2′-bipyridine complexes of rhodium and iridium act as electrocatalysts for the reduction of CO2.19 They found that cis-[Rh(bpy)2X2]+ (X is Cl or OTf) reduces CO2 at −1.55 V vs. SCE to predominantly form formate. It is interesting to note that no CO was detected in any of the electrolysis experiments; however, H2 was formed presumably by the degradation of the supporting electrolyte. This system was found to have both low turnovers (6.8 to 12.3) and poor current efficiencies (64% for formate, and 12% for H2). Meyer and co-workers also found that [M(bpy)2(CO)H]+ (M = Os, Ru) were electrocatalysts for the reduction of CO2.20 Under anhydrous conditions the major product was CO, and with the addition of H2O up to 25% formate was observed.
Palladium complexes using polydentate phosphine ligands represent some of the most extensively studied CO2 reduction catalysts to date. First reported in 1991 by the Dubois group22 these systems have shown significant development over the past 15 years.23,24 The typical catalyst system is based on a tridentate phosphine ligand that was initially coordinated to Co, Fe, or Ni. While the iron system showed catalysis of CO2 reduction, the overpotentials were high and the rates were slow. However, the Ni system displayed two one-electron reductions in the area of interest for CO2 reduction. This observation ultimately led this group to study the palladium triphos complexes as catalysts. The first reported complex, [Pd(triphos)(PR3)](BF4)2 (Fig. 5), would catalyze the reduction of CO2 to CO in acidic acetonitrile solutions. The active species was later found to be the phosphine-dissociated solvent complex [Pd(triphos)(solvent)](BF4)2.
Fig. 5 DuBois tridentate phosphine catalyst. |
In later mechanistic studies it was determined that as the Pd(II) complex gained an electron to form a Pd(I) intermediate, a reaction with CO2 occurred to form a five-coordinate CO2 adduct. Upon transfer of the second electron, the Pd(0) intermediate dissociated the solvent. Protonation of one of the oxygen atoms of coordinated CO2 affords a metallocarboxylic acid intermediate, Pd–COOH. It is believed that the metallocarboxylic acid is protonated again to form a “dihydroxy carbene”, and that CO is then formed by dehydration of the dihydroxy carbene. CO dissociation and solvent association regenerates the initial complex. This proposed catalytic cycle is shown in Scheme 3. In solutions of high acid concentration the rate determining step was found to be the reaction of the Pd(I) intermediate with CO2. However, in solutions of low acid concentrations the cleavage of the C–O bond to form carbon monoxide and water limits the rate of the catalytic cycle.
The triphosphine ligand system allows for variation of both electronic and steric effects. Changing the substituent on the central phosphorus to a mesityl group effectively blocked one of the open coordination sites cutting the rate of reaction in half, excluding the possibility of a six-coordinate intermediate that has been proposed for some Ni(I) macrocyclic complexes.13,25 The central donating atom of the tridentate ligand was varied from P to C, N, S, and As as shown in Fig. 6. None were as effective as the original triphosphine ligand, however mechanistic insight into the production of molecular hydrogen was gained. With the [Pd(PCP)(CH3CN)](BF4)2 (11) complex, CO2 was found to be a cofactor in the production of hydrogen suggesting that H2 production goes through the same intermediate that produces CO. From these studies DuBois hypothesized that the preference for forming hydrogen or carbon monoxide depends on the basicity/redox potential of the catalyst. The more negative redox potential favors the protonation of the Pd to form a hydride, the less negative redox potential favors protonation of the coordinated CO2 oxygen to form CO and H2O.
Fig. 6 Varied tridentate phosphine ligands. |
Recent studies have focused on complexes incorporating two or more independent Pd triphosphine units.23 Early studies with dendrimers of the Pd catalyst showed decreased activity and selectivity, however with the methylene bridged monomers of the catalysts rates were found to increase by three orders of magnitude, suggesting a cooperative binding of CO2. Along with this increase in catalytic activity came an increase in the formation of Pd(I)–Pd(I) bonds, thus decreasing catalyst lifetimes.
These classes of Pd phosphine complexes have shown catalytic rates in the range of 10 to 300 M−1 s−1 and with >90% current efficiencies for CO production. Overpotentials were in the range of 100–300 mV, yet turnover numbers were low (ca. 10–100) and the decomposition to Pd(I) dimers and hydrides eventually causes cessation of catalytic activity.
Catalytic CO2 reduction has been reported by our group as early as 1987 with a binuclear Ni(0) “cradle” complex, [Ni2(CNMe)3(dppm)2][PF6]2 operating around −0.87 V vs. SCE.26 However over extended time periods the complete carbonylation of the complex occurs to give Ni2(CO)3(dppm)2. More recently, we have returned to a binuclear [Ni2(μ-dppa)2(μ-CNR)(CNR)2] system.27 This system also suffers from the CO produced being trapped by the catalyst.
A key finding was made in 1992 when a new trinuclear nickel cluster [Ni3(μ3-I)(μ3-CNMe)(μ2-dppm)3]+ (dppm = bis(diphenylphosphino)methane) (15) was found to be an electrocatalyst for the reduction of CO2. We have extended this class of nickel cluster electrocatalysts significantly over the past several years to include other isocyanide capped clusters (16–20),28,29 the CO capped cluster 21 (Table 1),28 and chalcogenide capped clusters.30 Results of studies of the electrochemical kinetics of the reduction of CO2 by clusters 15–21 have been reported.29 These results are summarized here briefly. Under an atmosphere of CO2 in dry acetonitrile, the reduction of 15 affords CO and CO32− as the only products observed by GC, HPLC, and IR spectroscopy. The use of 13CO2 results in 13CO and 13CO32−, and no oxalate was observed. The relative rates of reaction of the alkyl or aryl substituted isocyanide or carbonyl capped clusters with CO2 follow the order: CNCH3 (15) ∼ CN(i-C3H7) (16) > CNC6H11 (17) > CNCH2C6H5 (18) > CO (21) > CN(t-C4H9) (19) ∼ CN(2,6-Me2C6H3) (20) (Table 2). It is important to note that although the values of E1/2(+/0) fall into a relatively narrow range of −1.18 (15) to −1.08 V (20) vs. SCE, small differences in E1/2(+/0) dramatically affect the rates of the reaction with CO2. The data indicate that the reaction rates of the isocyanide capped nickel clusters 16–20 with CO2 are primarily influenced by the reduction potentials of the clusters and that the size and geometry of the substituents of the capping ligand play a secondary role. The interaction of CO2 with the reduced forms of 16–20 can occur either at the isocyanide ligand or at the nickel core. In these clusters, we have favored a metal-based mechanism as molecular orbital studies have shown that the LUMO which becomes singly occupied in the reduction of this class of clusters is almost entirely metal-centered.28 The electrochemical studies also suggest that the energy of this orbital correlates with reactivity toward CO2. The secondary steric effect observed in the electrochemical kinetics study further suggests that CO2 activation is occurring on the isocyanide capped face of the clusters.
L | E1/2(+/0)a | FT-IRb | 31P NMRc | UV-Visd | |
---|---|---|---|---|---|
(V vs. SCE) | ν(CN) | δ (ppm) | λmax (ε) | ||
a Cyclic voltammograms recorded in CH3CN.b Recorded as KBr pellets.c Recorded at 121.6 MHz in CD3CN.d Recorded in CH3CN, λmax in nm, ε (M−1 cm−1× 103) given in parentheses.e Broad.f ν(CO). | |||||
15 | CNCH3 | −1.18 | 1927, 1871 | 0.4 s | 527.0 (3.8) |
16 | CN(i-C3H7) | −1.18 | 1876, 1815 | −0.2 s | 534.4 (4.5) |
17 | CNC6H11 | −1.17 | 1885, 1832 | 0.3 s | 536.5 (5.0) |
18 | CNCH2C6H5 | −1.11 | 1887, 1801 | 0.4 s | 529.3 (4.3) |
19 | CN(t-C4H9) | −1.12 | 1780e | −1.4 s | 539.2 (3.8) |
20 | CN(2,6-Me2C6H3) | −1.08 | 1849, 1822 | −1.9 s | 542.2 (3.4) |
21 | CO | −1.12 | 1726f | 3.1 s | 520.0 (6.3) |
L | E1/2(+/0)a | kCO2b | |
---|---|---|---|
(V vs. SCE) | (M−1 s−1) | ||
a Cyclic voltammograms recorded in CH3CN.b Rates determined by rotating disk voltammetry. | |||
15 | CNCH3 | −1.18 | 1.6 ± 0.3 |
16 | CN(i-C3H7) | −1.18 | 1.4 ± 0.3 |
17 | CNC6H11 | −1.17 | 0.5 ± 0.1 |
18 | CNCH2C6H5 | −1.11 | 0.2 ± 0.05 |
19 | CN(t-C4H9) | −1.12 | 0.0 ± 0.05 |
20 | CN(2,6-Me2C6H3) | −1.08 | 0.0 ± 0.05 |
21 | CO | −1.12 | 0.1 ± 0.1 |
The binuclear copper complex, [Cu2(μ-PPh2bipy)2(MeCN)2][PF6]2, 22, (PPh2bipy = 6-diphenylphosphino-2,2′-bipyridyl), and its pyridine analog, [Cu2(μ-PPh2bipy)2(py)2][PF6]2, 23 (Fig. 7), were also found to be efficient electrocatalysts for the reduction of CO2.31 Two sequential single electron transfers to 22 are observed at E1/2(2+/+) = −1.35 V and E1/2(+/0) = −1.53 V vs. SCE in MeCN. Both are required to effect CO2 reduction. Again, the CO2 derived products correspond mainly to the reductive disproportionation to CO and CO32−. The only gaseous product was found to be CO. A turn-over frequency of >2 h−1 was maintained over the course of a 24 hour experiment. The catalyst was still active at the end of this experiment, and 22 was recovered quantitatively in its original form. The homogeneous electron transfer kinetics for the reduction of CO2 by complex 22 were studied by chronoamperometry. The limiting rate constant, kCO2, for the reaction of the doubly reduced state of 22 with CO2, was determined to be 0.7 M−1 s−1 in acetonitrile. The rate constant, kCO2, for 22 in methylene chloride solvent is comparable, 0.6 M−1 s−1. However, the rate constant, kCO2, for the pyridine adduct, 23, in methylene chloride solvent is significantly less, 0.1 M−1 s−1. These data suggest that substitution of the labile acetonitrile or pyridine ligands of 22 and 23 respectively is required for CO2 reduction. Significantly, 22 is a 2e− electrocatalyst for the reduction of carbon dioxide. The 2e− redox cycle of 22 appears to lead to at least an order of magnitude increase in the steady state catalytic currents for CO2 reduction compared to our nickel cluster electrocatalysts described above. The difference in overall rates is significant since the limiting [CO2] dependent derived rate constants are comparable. This suggests that the nickel cluster catalysts are slow because they operate via a single electron redox cycle while the overall reduction of CO2 is a 2e− process. On the other hand, the potentials required by the copper catalysts to reduce CO2 are between 350 and 450 mV, more negative than those of the trinuclear nickel catalysts.
Fig. 7 Kubiak dinuclear copper complexes that reduce CO2 to CO and CO32− at −1.53 V vs. SCE. Catalyst shows turnover numbers of greater than 2 h−1 over 24 hour periods and is still intact after catalytic period. Catalysts are effective, but suffer from low turnover and high overpotentials. |
CO + H2O ↔ CO2 + 2H+ + 2e− | (E8) |
Upon addition of CO2 to the reduced state, CO2 binds to both the Ni and Fe. This binding causes minimal geometry changes and occupies the fourth position around Ni completing the square planar geometry. In the coordination of CO2, nickel acts as the Lewis base, while the iron acts as the Lewis acid, and the partial negative charge on the oxygen is stabilized through hydrogen bonding provided by the protein surroundings. The positions of the Ni and Fe are held in place by the Fe3S4 framework and are essentially unchanged by the presence or absence of CO2. The cluster also serves to act as an electronic buffer stabilizing the electronic charges on Fe and Ni during the catalytic cycle (Scheme 4). It is this low reorganization energy that allows for a catalyst with a turnover rate of 31000 s−1.32
Scheme 4 Proposed catalytic cycle for anaerobic CO dehydrogenases. |
Fig. 8 Active site of aerobic CO dehydrogenase. |
Scheme 5 Proposed catalytic pathway for aerobic CO dehydrogenase. |
The active site consists of distorted square pyramidal molybdenum with an apical oxo ligand, a hydroxyl group, bridging sulfur to the copper and bound to the protein through molybdopterin cytosine dinucleotide cofactor. The copper is attached to the protein through the Sγ atom of Cys-388. Within the second coordination sphere of the active site there are several groups within hydrogen bonding distance, allowing for stabilization of charges. In the CO reduced, air oxidized, cyanide inactivated, and n-butyl isocyanide bound states the hydrogen bonding distances remain essentially the same. The active site lies 17 Å deep within the protein with a hydrophobic channel that averages 7 Å in diameter.
In the refined structures of the reduced state, the active site remains structurally similar to the air oxidized state. Several of the bond lengths increase (Mo–Cu 3.74 Å to 3.93 Å and Mo–OH 1.87 Å to 2.03 Å) yet the overall geometry remains the same. The enzyme has a catalytic rate of 107 s−1. The active site of O. Carboxidovorans can be inhibited by the use of isocyanides. When inhibited by n-butyl isocyanide, an “intermediate” in the reduction of CO2 can be observed.
From the crystal structures reported to date the requirements for efficient reduction of carbon dioxide are becoming evident. Each of these structures exhibit late first row transition metals in low oxidation states, Cu(I) and Ni(0), and a redox reservoir (Mo(IV) and Fe3S4) to supply electrons for the reduction of carbon dioxide. A key feature of both enzymes is the minimal change in energy between the reduced and oxidized states.
Although most of the work mimicking active sites of proteins has focused on the hydrogenases,35 some research has been done on “bio-inspired” reduction of CO2. Tezuka and co-workers first reported the reduction of CO2 by using a Fe4S4(SR)42− cluster in DMF.36 With this cluster they showed that they were able to reduce CO2 to formate at a potential of −1.7 V vs. SCE. They also report considerable amounts of C3 hydrocarbons in the gas analyses after controlled potential electrolysis, however this may be due to the reduction of the electrolyte.
More recently, Tatsumi and co-workers reported the synthesis of sulfide bridged molybdenum-copper complexes related to the active site of O. carboxidovorans.37 Several compounds were synthesized mimicking the active site, however none exhibited reactivity towards either carbon monoxide or tert-butyl isocyanide. This emphasizes the difficulty in creating a working mimic of biological systems.
In the case of CO2, slow kinetics and high overpotentials for electrochemical reduction result from a large nuclear reorganization energy. The winning strategy for efficient reduction of CO2 must involve simultaneous multi-electron transfers and catalytic sites that direct nuclear configurations of reactants favorably for product formation. The development of catalysts to enable two-electron transfer from the same molecule will allow for kinetically more efficient reduction of CO2.
Past examples of carbon dioxide reduction catalysts have cleared some paths for understanding what remains to be done in order to solve the problems summarized above. In most of these past studies the major products have been carbon monoxide and formic acid, and on relatively small scales. In only a few studies has methane or methanol been the primary product of the reductions.
In order to do a multi-electron reduction of CO2 to a usable liquid fuel we will have to develop new methods for activating the carbon dioxide molecule and it will probably be necessary to develop a catalyst that can perform several different kinds of reduction or a series of catalysts that can each work together to perform the required steps.
As an example, consider the first logical step in a CO2 reduction scheme: the hydrogenation of CO2 to formic acid, HCOOH. CO2 is an “amphoteric” molecule (possessing both acidic and basic properties). The carbon atom is susceptible to attack by nucleophiles and the oxygen atoms are susceptible to attack by electrophiles. In the hydrogenation of CO2, therefore, one can think of activation of CO2 being initiated by attack of a nucleophilic metal hydride (H−) at the CO2 carbon atom. The transfer of charge from the hydride to the carbon atom in turn causes negative charge to develop on the oxygen atoms. This charge can be stabilized by a Brønsted acid (H+). In the limit of these interactions, one can consider the hydrogenation of CO2 to result from a heterolytic process that adds H− to the carbon and H+ to the oxygen as shown in Fig. 9. The next logical step in the reduction of CO2 is the deoxygenation of formic acid (HCOOH) to formaldehyde (H2CO). In photosynthesis, for example, CO2 is reduced to H2CO equivalents that are combined to form saccharides (H2CO)6. The reduction of formic acid to a formaldehyde equivalent can again be thought to proceed by H− addition to the carbon atom and H+ addition to the OH group. However, it is important to note that the details of this next step in the reduction (strength of hydride and proton equivalents, geometry of addition, etc.) can be expected to be quite different from those required in the production of formic acid in the first place. Similarly, the third step, namely reduction of formaldehyde to methanol (CH3OH), may proceed by a completely different mechanism such as the direct addition of H2 across the CO double bond of formaldehyde. This illustrates the need for detailed mechanistic and theoretical knowledge in the development of catalysts for the conversion of CO2 to liquid fuels.
Fig. 9 One way to envision proton-couple electron transfer to CO2 is from a metal hydride and a neighboring acid. |
One important lesson from previous work is that the fundamental reorganization energies of both the CO2 molecule and the catalysts that reduce CO2 are extremely important considerations. An important distinction between the biological CODHs and the synthetic transition metal complex catalysts for CO2 reduction is that the CODHs can equilibrate CO2 and CO. In other words; they can catalyze the reduction of CO2 in both directions. This implies that the CODHs function at the thermodynamic potential for CO2 reduction. It further implies that the CODHs operate at low barriers that result from active sites that direct the linear CO2 molecular substrate toward a necessarily bent CO2 configuration in the reduction intermediates.
A second feature which prior studies also suggest to be important in guiding future work is a multi-electron transfer capacity of the electrocatalysts. For example, of all of the known synthetic electrocatalysts for CO2 reduction, only the [Ni3(μ3-I)(μ3-CNR)(μ2-dppm)3]+ can equilibrate CO2 with CO and CO32−. The rates, however, are quite slow and this has been attributed to the fact that the [Ni3(μ3-I)(μ3-CNR)(μ2-dppm)3]+ catalysts function only by a single-electron redox cycle, while the reduction of CO2 to CO and CO32− is a two-electron process. The development of CO2 reduction catalysts that have highly reversible electrochemistry, indicative of low catalyst reorganization energies, but that also possess multiple electron redox capacities appears to be a very promising area of research.
A third point which has been largely overlooked in previous work, but should figure importantly in future work is proton coupled electron transfer. Proton coupled electron transfer (PCET) is frequently discussed in mechanisms for water splitting. PCET may well be even more important in CO2 reduction catalysis. The redox potentials summarized in eqn (1)–(5) show that the coupling of C–H bond forming reactions with CO2 reduction can lead to very reasonable overall thermodynamics. The key challenge here then is not just one proton coupled electron transfer, but multiple proton coupled electron transfers to produce molecules such as methanol or methane. One observation that we make from this is that it is difficult to imagine one catalyst that can do it all. As we noted in the previous section, in the hydrogenation of CO2, it appears that the most effective initial hydrogenation of CO2 to formic acid would occur by the heterolytic hydride (H−)/proton (H+) addition of H2. The second step in the reduction of CO2 is the deoxygenation of formic acid (HCOOH) to formaldehyde (H2CO). The reduction of formic acid to a formaldehyde equivalent may again be thought to proceed by H− addition to the carbon atom and H+ addition to the OH group, but by a mechanism that could be quite different from that which led to the production of formic acid in the first place.
Finally, the fact that the reduction of formaldehyde to methanol (CH3OH) may proceed by a completely different mechanism still illustrates that either a multi-functional single catalyst or panel of catalysts would be needed for efficient conversion of CO2 to methanol. Presently, we favor the second approach as shown in Scheme 6. This approach includes a panel of three catalysts that each contributes to the overall reduction of CO2 to methanol in optimized single steps. There are various platforms for such catalysts to be developed. These include immobilization on beads, anchoring to separate or single supports or surfaces, etc.
Scheme 6 Proposed path for the tandem catalytic reduction of CO2 to methanol. A series of three catalysts that each contributes to the overall reduction of CO2 to methanol in optimized single steps. |
Footnote |
† Part of the renewable energy theme issue. |
This journal is © The Royal Society of Chemistry 2009 |