Significance
Enzymes achieve rapid and reversible H2 oxidation catalysis by cooperative behavior between the active site and the protein scaffold. To better understand the role of the enzyme scaffold, we have attached amino acids (glycine, arginine, and arginine methyl ester) to an active functional mimic of hydrogenase to give . The resulting complexes are fully reversible catalysts with the arginine complex exhibiting high activity for both H2 oxidation/production, functionality achieved by the addition of an outer coordination sphere.
Keywords: hydrogenase mimics, reversible catalysis, amino acid catalysts, outer coordination sphere, homogeneous electrocatalysis
Abstract
Hydrogenases interconvert H2 and protons at high rates and with high energy efficiencies, providing inspiration for the development of molecular catalysts. Studies designed to determine how the protein scaffold can influence a catalytically active site have led to the synthesis of amino acid derivatives of complexes, (CyAA). It is shown that these CyAA derivatives can catalyze fully reversible H2 production/oxidation at rates approaching those of hydrogenase enzymes. The reversibility is achieved in acidic aqueous solutions (pH = 0–6), 1 atm 25% H2/Ar, and elevated temperatures (tested from 298 to 348 K) for the glycine (CyGly), arginine (CyArg), and arginine methyl ester (CyArgOMe) derivatives. As expected for a reversible process, the catalytic activity is dependent upon H2 and proton concentrations. CyArg is significantly faster in both directions (∼300 s−1 H2 production and 20 s−1 H2 oxidation; pH = 1, 348 K, 1 atm 25% H2/Ar) than the other two derivatives. The slower turnover frequencies for CyArgOMe (35 s−1 production and 7 s−1 oxidation under the same conditions) compared with CyArg suggests an important role for the COOH group during catalysis. That CyArg is faster than CyGly (3 s−1 production and 4 s−1 oxidation) suggests that the additional structural features imparted by the guanidinium groups facilitate fast and reversible H2 addition/release. These observations demonstrate that outer coordination sphere amino acids work in synergy with the active site and can play an important role for synthetic molecular electrocatalysts, as has been observed for the protein scaffold of redox active enzymes.
Hydrogenases are metalloenzymes that interconvert H2 and protons (Eq. 1), reactions necessary for biological processes within certain organisms to provide energy by splitting hydrogen, as well as to balance the redox potential in the cell (1–4). Their reversible catalytic behavior is a demonstration of their energy efficiency, indicating that they are operating at the equilibrium potential of the H2/H+ couple. They are also very active under conditions optimized for a particular direction, operating at up to 20,000 s−1 for H2 production (5) and 10,000 s−1 for H2 oxidation (6). Under conditions where reversibility is optimized (3, 7–10), net turnover frequencies (TOFs) are typically slower. For hydrogenases attached to electrode surfaces, the TOFs are not always quantitated due to uncertainty in surface coverage (9), but in one example, TOFs for a series of mutants of Desulfovibrio fructosovorans [NiFe]-hydrogenase were reported, ranging from 3 to 500 s−1 for H2 production and 600 to 1,000 s−1 for H2 oxidation under one set of conditions (8).
| [1] |
The ability of hydrogenases to function reversibly requiring no applied overpotential and with high catalytic TOFs is a direct reflection of their biological optimization, and a hallmark of enzymes. Reversibility requires enzymes that have fast and reversible proton and electron transfer, and that are nearly thermoneutral for H2 addition/elimination, so that little excess energy is needed for either the production or oxidation of hydrogen.
The complexes are hydrogenase mimics that can be tuned to either produce or oxidize H2. A key feature of the complexes is a positioned pendant amine in the second coordination sphere, inspired by [FeFe]-hydrogenase, to assist the metal in forming or heterolytically cleaving H2 (11, 12). We have shown that tuning for H2 oxidation or production catalysis depends on the free energy of H2 addition (ΔG°H2) to these complexes. Complexes with large positive ΔG°H2 values are generally fast H2 production catalysts with TOFs of 104 to 105 s−1 at overpotentials (OPs) of 0.5–0.6 V, and those with large negative ΔG°H2 values are fast H2 oxidation catalysts with TOFs of ∼50 s−1 at 1.0 atm H2 and an OP of 0.4 V using triethylamine as a base (13). The large positive or negative values of ΔG°H2 are associated with moderate to large overpotentials, where overpotential is the extra energy over the equilibrium potential needed to drive the reaction, determined at Ecat/2 (SI Appendix). Using thermoneutral H2 addition/elimination as a design principle, reversibility was achieved for a complex related to those studied here, , a complex that has a free energy of hydrogen addition of ΔGH2 = 0.8 kcal/mol (14). This complex operated with a TOF of <1 s−1 in both directions, the slow TOFs being consistent with what has typically been observed for this family of complexes as the thermodynamic driving force for H2 addition (ΔG°H2) approaches 0 kcal/mol. These observations point to features in hydrogenase enzymes that are not typically present in the catalysts.
To investigate how the protein scaffold can influence catalysis, we have used the complex as a probe molecule upon which to build an enzyme-inspired outer coordination sphere (15, 16). We recently reported the incorporation of the amino acids arginine or glycine into the outer coordination sphere of this class of catalysts to give (CyArg) or (CyGly), where the amine of the amino acid is a member of the P2N2 ring (Fig. 1) (17, 18). Both of these complexes are active, water-soluble H2 oxidation catalysts with TOFs up to 200 s−1 at 1 atm H2 and 298 K, operating at low overpotentials (less than 180 mV) (17, 18). Unexpectedly, catalysis was fastest under acidic conditions (from pH = 0 to pH = 1) (17, 18). Both complexes also demonstrated impressive TOFs for H2 oxidation as a function of pressure, up to 144,000 s−1 at 133 atm H2, 298 K, and pH = 0.5, albeit at reduced energy efficiency (overpotential = 460 mV) (18). The carboxyl functional group in both complexes is thought to play a key role by coupling proton and electron transfer, facilitated by the ability of the COOH group to deprotonate the pendant amine (17, 18). The CyArg derivative operates at similar overpotentials but with TOFs ∼6 times faster than those of the CyGly catalyst. Thus, both the carboxyl and guanidinium substituents play important roles in modifying catalytic activity and demonstrate the dramatic influence that the amino acid functional groups can have on the active site.
Fig. 1.
Complexes studied for reversible catalysis; (CyGly), (CyArg), and (CyArgOMe).
The low overpotential of the CyGly and CyArg catalysts suggested that they might also function as reversible catalysts for both H2 production and oxidation. To probe this proposed reversibility in more detail we report in this work the effect of increased temperatures and different hydrogen pressures on the performance of these catalysts. It is shown that both complexes can reversibly interconvert H+ and H2, which requires reversible proton and electron transfers as well as reversible H2 binding and release. A third complex, CyArgOMe, where ArgOMe is the methyl ester of arginine, has also been synthesized and studied to directly probe the role of the COOH group. Our results demonstrate that amino acids incorporated in the outer coordination sphere can facilitate fast, reversible enzyme-like catalysis for small synthetic molecular electrocatalysts.
Results
H2 Oxidation at Elevated Temperatures.
The H2 oxidation activity of CyArg was studied as a function of temperature under 1 atm of H2 at pH = 1.55. The plateau currents of the waves (icat) were scan-rate independent from 293 to 337 K (SI Appendix, Fig. S1) and showed a linear dependence on catalyst concentration (SI Appendix, Fig. S1) as expected for a process that is first order in catalyst. As illustrated in Fig. 2A, the catalytic current increased with increasing temperature. The TOFs and overpotentials were calculated from this electrochemical data (Materials and Methods), with TOFs ranging from 45 s−1 (293 K) to 205 s−1 (337 K). It can be seen from Fig. 2B that the overpotentials, the excess energy beyond the equilibrium potential needed for the reaction, decreased from 95 mV at 293 K to 50 mV at 337 K.† These TOFs and overpotentials are similar to those observed for hydrogenase enzymes (8).
Fig. 2.
Electrocatalytic data shows enhanced TOFs and lower overpotentials at increasing temperatures. (A) Electrocatalytic behavior of 45 μM CyArg under 1 atm H2 at temperatures from 293 to 337 K at pH = 1.55, 0.1 M Na2SO4. The data collected for the same sample under N2 at 293 K is shown in light gray. The vertical dotted line denotes the equilibrium potential. The horizontal black arrow indicates the initial scanning direction. Data were collected at a scan rate of 0.1 V/s. (B) The TOF (black squares) and overpotential (blue circles) for H2 oxidation for CyArg as a function of temperature shows an increase in TOF with a corresponding decrease in overpotential.
Reversible Catalysis: Effect of Decreased H2 Pressure and Elevated Temperatures on H2 Oxidation and Production.
Under 1 atm H2, the catalytic response is dominated by H2 oxidation (Fig. 2). The low observed overpotentials suggest the complex can operate reversibly, and that catalysis of proton reduction to H2 should be observed at lower H2 concentrations. To provide more favorable conditions under which to observe reversible H2 oxidation/H2 production, a mixture of 25% H2/Ar was used. Fig. 3A shows the electrocatalytic behavior at five pH values at 348 K under 1.0 atm of a 25% H2/Ar mixture. Both H2 production and H2 oxidation are observed at all pH values operating at the equilibrium potential of the H2/H+ couple, consistent with a truly reversible system. The fastest TOFs in both directions are observed for pH values under 2. Scan rate independence was confirmed for the catalytic waves in both directions (SI Appendix, Fig. S2). Reversible catalysis was not observed below ∼320 K.
Fig. 3.
Reversible H2 oxidation/production catalysis is observed for CyArg at various pH conditions under 1 atm 25% H2/Ar. (A) Cyclic voltammograms for 50 μM CyArg at 348 K in 0.1 M Na2SO4: pH 0 (black trace), pH 1.0 (red trace), pH 2.5 (blue trace), pH 5.0 (magenta trace), and pH 6.0 (green trace). The corresponding dotted vertical lines correspond to the equilibrium potential for each pH condition. The horizontal dotted line demarcates the zero current line and the horizontal black arrow indicates the initial scanning direction. Data were collected at a scan rate of 0.1 V/s. (B) TOFs for H2 production (black) and H2 oxidation (blue) as a function of pH at 348 K.
The fastest TOFs (largest currents) were observed for CyArg at pH 1 and below: 300 s−1 for H2 production and 20 s−1 for H2 oxidation, with little enhancement at lower pH values (Fig. 3B, Table 1). The H2 oxidation TOFs showed a counterintuitive increase of ∼3 times as the pH decreased from 6.0 to 0.0, similar to previous observations at room temperature (7). For H2 production, a ∼50-fold increase in TOF is observed as the pH decreases from 6.0 to 1.0. The observed half wave potentials for H2 oxidation and H2 production were within 15 mV of each other in all cases, as expected for a fully reversible catalytic process.
Table 1.
Maximum rates of H2 oxidation and production under reversible conditions (348 K, 1 atm)
| H2 production TOF ± 5% (s−1) | ||||
| Complex | pH | H2 oxidation TOF ± 5% (s−1) Under 25% H2/Ar | Under 25% H2/Ar | Under N2 |
| CyArg | 6.0 | 6 | 7 | 21 |
| 5.0 | 11 | 10 | 32 | |
| 2.5 | 14 | 15 | 36 | |
| 1.0 | 20 | 300 | 403 | |
| 0 | 20 | 304 | 403 | |
| CyArg(OMe) | 0 | 7 | 35 | 124 |
| CyGly | 1.0 | 4 | 3 | 12 |
To ensure that the observed reactivity was due to the CyArg complex, two control experiments were performed. In one experiment (SI Appendix, Fig. S3A) no complex was used and no catalysis was observed as expected. In a second experiment (SI Appendix, Fig. S3B), [Ni(CH3CN)6](BF4)2 was tested for catalytic activity at pH = 0 and 348 K. No significant current was observed, indicating that the observed response was due to the CyArg complex.
To evaluate the importance of the guanidinium groups in CyArg for catalytic reversibility, the analogous water-soluble H2 oxidation electrocatalyst, CyGly, was studied, where the arginine in the complex shown in Fig. 1 is replaced with a glycine. Under similar conditions, 348 K, 1 atm 25% H2/Ar, and pH = 1.0, the optimal pH conditions for the CyGly catalyst, reversible catalytic behavior is also observed but at lower TOFs (Table 1 and SI Appendix, Fig. S4). The TOF for H2 production was 3 and 4 s−1 for H2 oxidation, significantly slower in both directions compared with CyArg.
A second control complex, the methyl ester of CyArg, (CyArgOMe), was prepared to test the importance of the COOH groups. The TOFs for this complex were also significantly reduced compared with CyArg, with TOFs at 348 K, 1 atm 25% H2/Ar, and pH = 1.0 of 7 s−1 for H2 oxidation and 35 s−1 for H2 production (Table 1), but faster than those observed for CyGly.
Reversibility under an N2 Atmosphere: Studies of H2 Production in the Absence of H2.
To enhance H2 production at high temperature, the electrochemical behavior for CyArg was studied under an N2 atmosphere from pH = 0.0 to pH = 6.0, and temperatures up to 348 K. Under these conditions the catalyst exhibited predominantly H2 production (Table 1, Fig. 4, and SI Appendix, Fig. S5). The small contribution from H2 oxidation (Fig. 4, asterisk) is a result of the H2 produced in the reverse reaction. This wave is not observed in anodic sweeps originating at −0.05 V, providing further evidence of reversibility. Similar results were observed for CyGly at pH = 1.0 (SI Appendix, Fig. S4).
Fig. 4.
CyArg under N2 shows reversible catalysis dominated by H2 production. Cyclic voltammograms for 50 μM CyArg under 1 atm N2 in aqueous 0.1 M Na2SO4 at pH 0.0 and temperatures from 293 to 348 K. The vertical gray dotted line denotes the equilibrium potential for reversible catalysis. The horizontal black arrow indicates the initial scanning direction. Scan rates were 0.1 V/s. The asterisk denotes H2 oxidation observed as a result of the H2 produced from the H+ reduction.
In the electrochemistry for all three complexes, a second wave is observed negative of the reversible electrocatalytic response (Hrev). This wave (Hirr) is most easily seen for CyArg in Fig. 3 and is also scan-rate independent. Based on the lack of catalytic activity for a glassy carbon electrode following a rinse test, this wave appears to represent a second homogeneous process for catalytic H2 production (SI Appendix, Fig. S6) (18). At lower pH values, Hrev increases relative to Hirr suggesting either enhanced catalytic TOFs for Hrev vs. Hirr, or that the species resulting in Hirr is transforming into the species responsible for Hrev. The mechanism for Hirr is not yet clear and is the subject of further investigation.
H2 Oxidation at High Pressure and High Temperature.
To investigate the role of temperature on overpotential, catalysis for CyArg was run at 348 K, at 67 atm H2. Although high pressure alone results in an increase in overpotential (18), operating at elevated temperatures and pressure (SI Appendix, Fig. S7) results in a reduction in the overpotential, from 350 mV at 298 K to 200 mV at 350 K.
Reversibility of H2 Addition.
Increasing TOFs of H2 production as a function of decreased H2 pressure (Fig. 3) suggests that H2 addition becomes reversible. To further evaluate the reversibility of the addition of H2 to CyArg, two additional experiments were performed. In one experiment, hydrogen addition is shown to be reversible by purging the hydrogen addition product with N2, resulting in the loss of H2 based on the recovery of the characteristic red color of the NiII species. In a second experiment, 31P NMR experiments were performed for CyArg at temperatures from 298 to 348 K at pH values of 1.0 and 4.0 at 1 atm H2. At both pH values, complete conversion to the hydrogen addition products was confirmed. Upon heating, the Ni(II) species was regenerated. Cycling the temperature between 313 and 348 K resulted in the reversible addition (313 K) and loss (348 K) of H2 as seen in Fig. 5 and SI Appendix, Figs. S8 and S9.
Fig. 5.
H2 binding to CyArg is reversible above 313 K. 31P{1H} NMR spectra for CyArg at pH = 4.0 in water. (A) The starting Ni(II) (under N2) looks identical from 298 to 348 K (shown); (B) exposed to 1 atm 25% H2 at 298 K, conversion to the H2 addition products, [Ni(0)(PCy2NArg2H)2], is observed, including the e/e (15.5 ppm; 60%) and e/x (e/x: 14 ppm and e/x: −9 ppm; 40%) isomers. These isomers are the only observed species up to 313 K (shown) and the e/e and e/x peaks merge at 323 K (SI Appendix, Fig. S9). (C) Heating to 348 K, Ni(II) was observed (11 ppm; 10%), demonstrating H2 release. (D and E) Cycling the temperature of the solution between 313 and 348 K demonstrates reversible H2 addition from the H2 addition products. Integrals are shown.
Stability of CyArg at High Temperature and Low pH.
The stability of CyArg was tested by cyclic voltammetry and NMR spectroscopy. Sequential cyclic voltammograms showed no degradation in catalytic peak current over 20 min at 348 K (SI Appendix, Fig. S10). The 31P{1H} NMR spectra at elevated temperature also showed minimal degradation (∼5%) after 90 min under similar solution conditions (SI Appendix, Figs. S8 and S9).
Protonation State of the Complex.
The extent of protonation of CyGly was examined in MeCN by 31P{1H} NMR. Under N2, no pendant amines were protonated and no hydrides were observed, as expected (Fig. 6, bottom spectrum). Introducing 1 atm H2 generated the hydrogen addition product (15.5 ppm) in the 1H NMR spectrum (Fig. 6, middle spectrum).‡ Addition of acid results in a shift in the hydrogen addition product to 13.8 ppm, and a hydride species at −11.4 ppm (Fig. 6, top spectrum). The integrations are consistent with a species with three protons, instead of only two as expected if the only source was the addition of H2.
Fig. 6.
The 1H NMR spectroscopy indicates protonation of the complex CyGly in MeCN. (Bottom) Under N2 at neutral pH, no protonated amines or hydrides are observed in the 1H NMR spectrum. (Middle) Adding H2 at neutral pH results in the hydrogen addition product, x/x in MeCN (15.6 ppm). (Top) The addition of three equivalents of HTFSI results in an additional resonance at −11.4 ppm, consistent with a hydride, along with an observed upfield shift in the x/x resonance to 13.8 ppm. The integrations are consistent with a species with three protons.
Discussion
The TOFs and lack of overpotential of the CyArg complex presented here are comparable to those of the [NiFe]-hydrogenase enzymes. In the following discussion we will examine some of the features responsible for this enzymelike performance, including the specific roles of H2 pressure, highly acidic solutions, ligand structure, and temperature.
The Effect of H2 Pressure.
The catalytic TOFs for H2 oxidation for CyArg depend linearly on the H2 pressure (18), and H2 pressures of 1.0 atm or greater inhibit H2 production (Fig. 2). For lower partial pressures of H2 (0.25 or 0.0 atm, Figs. 3 and 4, respectively) reversible catalysis is observed by cyclic voltammetry experiments with large currents for both H2 oxidation and production. These results indicate the reversibility of H2 binding, and also imply that the TOFs for these complexes are determined by the rate of H2 addition, and by extension of microscopic reversibility, hydrogen elimination.
The Role of Aqueous, Acidic Solutions.
As can be seen from Fig. 3, the catalytic TOFs increase dramatically below pH 2. As demonstrated by NMR experiments, the catalyst is more highly protonated under acidic conditions (Fig. 6), likely resulting in a different catalytic mechanism. A proposed mechanism is shown in Fig. 7 (clockwise for H2 oxidation, counterclockwise for H2 production), with the standard mechanism for the nonprotonated species shown in SI Appendix, Fig. S11. This mechanism is initiated by the protonation of the doubly charged Ni(II) complex, 1, to form the triply charged Ni(II) complex, 2. Protonation produces a positive shift of the Ni(II/I) couple (Fig. 3), indicating a lower-energy lowest unoccupied molecular orbital (LUMO). The lower-energy LUMO of 2 will likely also lower the associated barriers for H2 addition/elimination (19), contributing to faster and more reversible H2 addition for complex 2 compared with complex 1. We hypothesize that this proton is positioned endo, based on the observation from 31P{1H} NMR that the catalyst preferentially positions protons endo to the Ni in aqueous solution (Fig. 5), the necessary location for H2 release.
Fig. 7.
Altered mechanism due to protonation. The first step in this mechanism is the protonation of the Ni(II). The carboxylic acids, as well as the R′ groups could contribute in several hydrogen bonding and protonation/deprotonations, one of which is shown. The additional amino acids are not shown for simplicity. Note that for simplicity, the charges shown do not account for charges on the amino acid side chains, hence the overall charge will depend on the nature of the amino acid.
H2 addition generates a hydride (Figs. 6 and 7) and protonates one pendant amine, likely on the ligand that was not initially protonated. This ligand will have more basic pendant amines and will thus be the favored site for heterolytic H2 cleavage (16, 20) as shown in Fig. 7. The overall mechanism is supported by the hydride intermediate observed using 31P{1H} NMR spectroscopy (Fig. 6) that is consistent with complex 4, an intermediate not observed for nearly all other complexes (16). Of interest is that once complex 4 is formed, there are two equivalent deprotonation sites (i.e., the molecule is now symmetric, assuming fast mobility of the hydride, which has been observed before for the nonprotonated complex) (21). This effectively provides two sites for deprotonation, and may contribute to the high efficiency for this reversible process. Multiple proton channels have been proposed for both [FeFe]- and [NiFe]-hydrogenases (22–26), and it may be that the enzymes use multiple channels to achieve maximum efficiency. Structural studies and studies evaluating the movement of protons are ongoing to provide additional details of the mechanism for the protonated complex.
The Effect of the Carboxyl Group.
The carboxyl group of the amino acid plays an important role in the catalytic mechanism. As discussed previously (17, 18), the introduction of a carboxylic acid in the outer coordination sphere results in rapid proton transfer, a critical component in eliminating the overpotentials required for reversibility. The endo positioned carboxyl group may also play a role in stabilizing the dihydrogen intermediate, 3, of Fig. 7. The observation that the CyArg catalyst is 3 times faster than CyArgOMe for H2 oxidation and 8 times faster for H2 production, coupled with the observation that H2 addition is rate limiting, indicates that the OH group of the carboxyl substituent facilitates H–H bond cleavage as shown in structure 3. This interaction is less likely for CyArgOMe, because the larger Me group of the ester will interact more strongly with the R substituent on P, thereby reducing the interaction between the carboxyl group and the dihydrogen ligand. This possibility is currently under investigation using computational methods.
The Effect of the Amino Acid Side Chain.
A comparison of the TOFs of CyGly with CyArg for H2 oxidation and production at pH 1 (Table 1) indicates that the Arg substituent enhances H2 oxidation by a factor of five and H2 production by a factor of 100. This demonstrates the essential influence of the amino acid side chain on the activity, and because H2 addition is rate limiting, on the rate of H2 addition and elimination in particular. This has been attributed previously to the CyArg complex exhibiting Arg–Arg interactions that alter the Ni…N distance to facilitate H2 addition (18). Another possible contribution is a fine tuning of the positioning of the carboxyl group due to interactions between the amino acid side chains, optimizing its role. These postulates will be aided by ongoing structural studies.
The Effect of Temperature.
As shown in Figs. 2 and 4, increasing the temperature results in faster TOFs for H2 oxidation combined with lower overpotentials. In contrast, previous studies have shown that when H2 pressure alone was used to increase the TOFs, i.e., in the absence of elevated temperature, the overpotential also increased, from 180 to 460 mV (17, 18). The higher overpotential under high pressure conditions in the absence of elevated temperature is consistent with slow electron and/or proton transfer, because a more positive potential is required for the electron transfer rate to match the rate of addition of H2. Therefore, the observed decrease in overpotential as a function of temperature suggests that temperature is facilitating the electron and/or proton transfers shown in Fig. 7. At sufficiently high temperatures, a Nerstian shape is observed for the reversible electrocatalytic waves and the potential shifts linearly as a function of pH (60 mV per pH unit), all of which are consistent with rapid, reversible electron and proton transfers under these conditions. For efficient electrocatalysis (i.e., operating at low or no overpotential), electron-transfer and proton-transfer rates must be fast and reversible, and at least as fast as the H2 addition and elimination steps. CyArg achieves this with a combination of factors both internal to and external to the complex.
The temperature also affects H2 addition and elimination. As shown in Fig. 5, and SI Appendix, Figs. S8 and S9, H2 addition is reversible above 313 K. This is consistent with the temperature at which reversible catalysis is initially observed (323 K), consistent with the need for reversible H2 binding and release before reversible activity can be observed.
Summary and Conclusions
Achieving reversibility for a catalytic cycle necessitates achieving reversibility for all of the steps that occur during catalysis. These steps can be grouped into three categories: (i) e− transfer steps, (ii) H+ transfer steps, and (iii) H2 addition and elimination. For the complexes described above, fully reversible catalysis of H2 oxidation and production has been achieved by an interrelated combination of solution conditions (pH, H2 pressure, and temperature) and careful catalyst design.
CyArg displayed reversible catalytic H2 oxidation and production behavior at a broad pH range (0–6) at elevated temperatures (>323 K) and subatmospheric H2 pressure. The complex showed the maximal activity for both directions at low pH conditions (pH < 1), and H2 production was inhibited under higher pressures of H2. Catalyst design is also critical to the achieved catalytic functionality and our studies demonstrate that reversible H2 addition and release from the catalyst at pressures close to 1.0 atm H2 begins with a suitable choice of the pendant amine (a component of the second coordination sphere) and the electronic properties of the Ni center (determined by the first coordination sphere). In addition, it is also clear that the carboxylates and side chains introduced into the outer coordination sphere by incorporation of amino acids exert an essential influence on the catalytic properties of the active center in much the same way that the protein matrix is thought to influence the catalytic activity of enzymes. In this sense, the amino acid substituted complexes described here, that are beginning to rival the TOFs and overpotentials of enzymes, represent a remarkable opportunity to probe in detail how amino acids and the protein matrix can influence the activity of catalytically active sites in synthetic electrocatalysts.
Although the mechanistic details controlling fast, reversible catalysis in hydrogenases are likely different, the Léger group demonstrated that catalytic bias in [NiFe]-hydrogenase is controlled not by the redox properties of the active site, but by residues remote from the active site that control either H2 release (H2 production) or electron transfer (H2 oxidation) (8). In this model system, we demonstrate that we are controlling H2 addition and proton transfer with groups remote from the active site, providing a conceptual parallel between the role of the outer coordination sphere in the molecular model and the role of the protein scaffold in the enzymatic system.
It is notable that the absence of overpotentials and the high catalytic TOFs observed for both H2 production and oxidation at 348 K for CyArg are comparable to those of the [NiFe]-hydrogenase enzymes, but offer significant advantages in stability under a broader set of reaction conditions. CyArg is not inhibited by CO (18), is stable for extended periods at high acid concentrations (18), and as shown in this work, is reasonably stable under high acid at elevated temperatures. These conditions are optimal for functioning in fuel cells, and were achieved by using an enzyme-inspired design, including first, second, and outer coordination spheres.
Materials and Methods
General Procedures.
All samples were prepared under an N2 atmosphere either using a standard single manifold Schlenk line or glove box. Ultra-pure water, 18.2 MΩ • cm, was obtained from a Millipore unit.
Preparation of complexes.
The (CyGly) (17) and (CyArg) (18) were prepared as previously reported. (CyArgOMe) is a new complex and was prepared by the same synthetic procedure, using arginine methyl ester HCl as the amine precursor.
: 1H NMR (CD3OD): δ 1.15–1.97 (H-Cy, Arg-Hδ, and Arg-Hε, 30H, m); 3.17–3.41 ppm (-PCH2N, Arg-Hα, and Arg-Hγ, 14H, m), 3.72 (Arg-COOCH3, 3H, s). 31P{1H} NMR (CD3OD): δ −41.6 ppm (broad), -45.2 ppm (sharp). ESI MS (+ve mode) (in methanol): m/z : 657.42 (calcd: 656.78).
: 1HNMR (D2O): δ 1.18–1.99 (H-Cy, Arg-Hδ, and Arg-Hε, 30H, m); 3.15–3.32 ppm (-PCH2N, Arg-Hα, and Arg-Hγ, 14H, m), 3.75 (Arg-COOCH3, 3H, s). 31P{1H} NMR (D2O): 15.6 ppm, broad. ESI MS (+ve mode) (in methanol): m/z : 1489.78 (calcd: 1489.13).
Electrochemistry.
Cyclic voltammetry was performed with solutions of 0.1 M Na2SO4 electrolyte in water using a glassy-carbon electrode (1 mm diameter), polished with 0.25 μm MetaDi diamond polishing paste (Buehler). The buffer solutions with pH values of 5.0 and 6.0 were prepared using 0.1 M 2-morpholinoethanesulfonic acid (Mes) buffer followed by adjusting the pH using dilute aqueous NaOH and dilute acetic acid solutions. Solutions with pH 0.0 and 1.0 were prepared by adding appropriate amounts of 1.16 M HClO4, and pH 1.55 were prepared by adding bis(trifluoromethane)sulfonamide acid (HTFSI). All solution pHs were constant throughout the experiment as monitored by pH meter. Cyclic voltammetry experiments were performed from 298 to 348 K on a CH Instruments 600D electrochemical analyzer using a standard three-electrode configuration. A glassy carbon rod was used as the counterelectrode and an AgCl-coated Ag wire (in saturated KCl) separated from the analyte solution by a Vycor frit was used as the reference electrode. All couples were referenced to the equilibrium potential observed at each pH condition along with the internal standard hydroxymethyl ferrocenium/ferrocene couple at 380.0 mV vs. SHE (see SI Appendix for determination of referencing using the open circuit potential method, SI Appendix, Fig. S12).
Turnover frequencies were determined with Eq. 2, as previously reported (17, 18):
| [2] |
In Eq. 2, icat is the catalytic current (measured from the zero current to the plateau of the wave for either H2 oxidation or H2 production); n is the number of electrons involved in catalysis (two in this case); F is Faraday’s constant; A is the surface area of the electrode (17); [cat] is the catalyst concentration; and D is the diffusion coefficient for the catalyst. The diffusion coefficients were determined previously (17, 18) and extrapolated to higher temperatures as described in the SI Appendix. The overpotentials were calculated by taking the difference between the Ecat/2 (the potential corresponding to 1/2 the catalytic current, icat/2) and the equilibrium potential (SI Appendix, Fig. S7), measured using the open circuit potential method (SI Appendix) (27).
NMR Studies.
H2 addition/elimination.
The 50 μM solution of CyArg complex was prepared in pH 4.0 solution and purged with 25% H2 in Ar for 15 min; a 31P{1H} NMR spectrum of the resulting sample was recorded at 313 K. Then the sample was heated and cooled between 348 and 313 K 2 times, recording 31P{1H} NMR spectra at each temperature. In a separate experiment, a 31P{1H} NMR spectrum was recorded for 50-μM CyArg under an N2 atmosphere and similar conditions (pH 4.0 solution at 348 K).
Protonation state of CyGly.
The 1H NMR spectrum of 1.0 mM solution of CyGly* (Gly*: 13C/15N labeled glycine) complex in CD3CN was recorded under N2. The sample was purged with H2 for 15 min and the changes in 1H NMR spectra were monitored. Another spectrum was measured following the addition of three equivalents of HTFSI to the resulting solution under H2 atmosphere. All of the measurements were performed at room temperature (298 K).
Supplementary Material
Acknowledgments
This work was funded by the Office of Science Early Career Research Program through the US Department of Energy (DOE), Basic Energy Sciences (A.D. and W.J.S.), and the Center for Molecular Electrocatalysis, an Energy Frontier Research Center funded by the US DOE Office of Science, Basic Energy Sciences (D.L.D. and J.A.S.R.). Pacific Northwest National Laboratory is operated by Battelle for the US DOE.
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1416381111/-/DCSupplemental.
The numbers reported previously (14) were based on a reference to the hydroxymethyl ferrocene/hydroxymethyl ferrocenium couple, which has a reported potential of 440 mV relative to standard hydrogen electrode (SHE). The reversible behavior of the catalyst as well as the open circuit potential measurements reported in this work allow us to establish the equilibrium potential with confidence and without an external reference, allowing us to adjust the previously reported numbers, vida infra.
In MeCN, the exo–exo (x/x, 15.5 ppm) isomer is observed, contradictory to what is observed in water, where only endo species are present, previously demonstrated to be stabilized in water (17) and also seen in SI Appendix, Fig. S9.
References
- 1.Fontecilla-Camps JC, Volbeda A, Cavazza C, Nicolet Y. Structure/function relationships of [NiFe]- and [FeFe]-hydrogenases. Chem Rev. 2007;107(10):4273–4303. doi: 10.1021/cr050195z. [DOI] [PubMed] [Google Scholar]
- 2.Frey M. Hydrogenases: Hydrogen-activating enzymes. ChemBioChem. 2002;3(2-3):153–160. doi: 10.1002/1439-7633(20020301)3:2/3<153::AID-CBIC153>3.0.CO;2-B. [DOI] [PubMed] [Google Scholar]
- 3.Vincent KA, Parkin A, Armstrong FA. Investigating and exploiting the electrocatalytic properties of hydrogenases. Chem Rev. 2007;107(10):4366–4413. doi: 10.1021/cr050191u. [DOI] [PubMed] [Google Scholar]
- 4.Lubitz W, Ogata H, Rüdiger O, Reijerse E. Hydrogenases. Chem Rev. 2014;114(8):4081–4148. doi: 10.1021/cr4005814. [DOI] [PubMed] [Google Scholar]
- 5.Madden C, et al. Catalytic turnover of [FeFe]-hydrogenase based on single-molecule imaging. J Am Chem Soc. 2012;134(3):1577–1582. doi: 10.1021/ja207461t. [DOI] [PubMed] [Google Scholar]
- 6.Jones AK, Sillery E, Albracht SP, Armstrong FA. Direct comparison of the electrocatalytic oxidation of hydrogen by an enzyme and a platinum catalyst. Chem Commun (Camb) 2002;(8):866–867. doi: 10.1039/b201337a. [DOI] [PubMed] [Google Scholar]
- 7.Fourmond V, et al. Steady-state catalytic wave-shapes for 2-electron reversible electrocatalysts and enzymes. J Am Chem Soc. 2013;135(10):3926–3938. doi: 10.1021/ja311607s. [DOI] [PubMed] [Google Scholar]
- 8.Abou Hamdan A, et al. Understanding and tuning the catalytic bias of hydrogenase. J Am Chem Soc. 2012;134(20):8368–8371. doi: 10.1021/ja301802r. [DOI] [PubMed] [Google Scholar]
- 9.Lukey MJ, et al. How Escherichia coli is equipped to oxidize hydrogen under different redox conditions. J Biol Chem. 2010;285(6):3928–3938. doi: 10.1074/jbc.M109.067751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.McIntosh CL, Germer F, Schulz R, Appel J, Jones AK. The [NiFe]-hydrogenase of the cyanobacterium Synechocystis sp. PCC 6803 works bidirectionally with a bias to H2 production. J Am Chem Soc. 2011;133(29):11308–11319. doi: 10.1021/ja203376y. [DOI] [PubMed] [Google Scholar]
- 11.Rakowski DuBois M, DuBois DL. The role of pendant bases in molecular catalysts for H2 oxidation and production. C R Chim. 2008;11:805–817. [Google Scholar]
- 12.Rakowski DuBois M, DuBois DL. The roles of the first and second coordination spheres in the design of molecular catalysts for H2 production and oxidation. Chem Soc Rev. 2009;38(1):62–72. doi: 10.1039/b801197b. [DOI] [PubMed] [Google Scholar]
- 13.DuBois DL. Development of molecular electrocatalysts for energy storage. Inorg Chem. 2014;53(8):3935–3960. doi: 10.1021/ic4026969. [DOI] [PubMed] [Google Scholar]
- 14.Smith SE, Yang JY, DuBois DL, Bullock RM. Reversible electrocatalytic production and oxidation of hydrogen at low overpotentials by a functional hydrogenase mimic. Angew Chem Int Ed Engl. 2012;51(13):3152–3155. doi: 10.1002/anie.201108461. [DOI] [PubMed] [Google Scholar]
- 15.Ginovska-Pangovska B, Dutta A, Reback ML, Linehan JC, Shaw WJ. Beyond the active site: The impact of the outer coordination sphere on electrocatalysts for hydrogen production and oxidation. Acc Chem Res. 2014;47(8):2621–2630. doi: 10.1021/ar5001742. [DOI] [PubMed] [Google Scholar]
- 16.Shaw WJ, Helm ML, DuBois DL. A modular, energy-based approach to the development of nickel containing molecular electrocatalysts for hydrogen production and oxidation. Biochim Biophys Acta. 2013;1827(8-9):1123–1139. doi: 10.1016/j.bbabio.2013.01.003. [DOI] [PubMed] [Google Scholar]
- 17.Dutta A, et al. Minimal proton channel enables H2 oxidation and production with a water-soluble nickel-based catalyst. J Am Chem Soc. 2013;135(49):18490–18496. doi: 10.1021/ja407826d. [DOI] [PubMed] [Google Scholar]
- 18.Dutta A, Roberts JAS, Shaw WJ. Arg-Arg pairing enhances H2 oxidation catalyst performance. Angew Chem Int Ed. 2014;53:6487–6491. doi: 10.1002/anie.201402304. [DOI] [PubMed] [Google Scholar]
- 19.Raugei S, et al. The role of pendant amines in the breaking and forming of molecular hydrogen catalyzed by nickel complexes. Chemistry. 2012;18(21):6493–6506. doi: 10.1002/chem.201103346. [DOI] [PubMed] [Google Scholar]
- 20.Yang JY, et al. Hydrogen oxidation catalysis by a nickel diphosphine complex with pendant tert-butyl amines. Chem Commun (Camb) 2010;46(45):8618–8620. doi: 10.1039/c0cc03246h. [DOI] [PubMed] [Google Scholar]
- 21.O’Hagan M, et al. Moving protons with pendant amines: Proton mobility in a nickel catalyst for oxidation of hydrogen. J Am Chem Soc. 2011;133(36):14301–14312. doi: 10.1021/ja201838x. [DOI] [PubMed] [Google Scholar]
- 22.Fdez Galván I, Volbeda A, Fontecilla-Camps JC, Field MJ. A QM/MM study of proton transport pathways in a [NiFe] hydrogenase. Proteins. 2008;73(1):195–203. doi: 10.1002/prot.22045. [DOI] [PubMed] [Google Scholar]
- 23. Ginovska-Pangovska B, Ho M-H, Linehan JC, Cheng Y, Dupuis M, Raugei S, Shaw, WJ (2014) Molecular dynamics study of the proposed proton transport pathways in [FeFe]-hydrogenase. Biochim Biophys Acta Bioenergetics 1837(1):131-138. [DOI] [PubMed]
- 24.McCullagh M, Voth GA. Unraveling the role of the protein environment for [FeFe]-hydrogenase: A new application of coarse-graining. J Phys Chem B. 2013;117(15):4062–4071. doi: 10.1021/jp402441s. [DOI] [PubMed] [Google Scholar]
- 25.Sumner I, Voth GA. Proton transport pathways in [NiFe]-hydrogenase. J Phys Chem B. 2012;116(9):2917–2926. doi: 10.1021/jp208512y. [DOI] [PubMed] [Google Scholar]
- 26.Teixeira VH, Soares CM, Baptista AM. Proton pathways in a [NiFe]-hydrogenase: A theoretical study. Proteins. 2008;70(3):1010–1022. doi: 10.1002/prot.21588. [DOI] [PubMed] [Google Scholar]
- 27.Kilgore UJ, et al. Studies of a series of Ni(PPh2NR2)2](BF4)2 complexes as electrocatalysts for H2 production: Effect of substituents, acids, and water on catalytic rates. J Am Chem Soc. 2011;133(15):5861–5872. doi: 10.1021/ja109755f. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







