This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes. Research Article Cite This: ACS Catal. 2019, 9, 5837−5846 pubs.acs.org/acscatalysis Improved Electro- and Photocatalytic Water Reduction by Confined Cobalt Catalysts in Streptavidin Arnau Call,†,‡ Carla Casadevall,†,‡ Adrian Romero-Rivera,§ Vlad Martin-Diaconescu,† Dayn J. Sommer,∥ Sílvia Osuna,§,⊥ Giovanna Ghirlanda,*,∥ and Julio Lloret-Fillol*,†,⊥ Downloaded via 84.88.84.34 on March 9, 2020 at 12:37:01 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles. † Institute of Chemical Research of Catalonia (ICIQ), The Barcelona Institute of Science and Technology, Avinguda Països Catalans 16, 43007 Tarragona, Spain § Institut de Química Computacional i Catàlisi (IQCC) and Departament de Química, Universitat de Girona, Carrer Maria Aurèlia Capmany 69, 17003 Girona, Spain ∥ School of Molecular Sciences, Arizona State University, Tempe, Arizona 85287-1604, United States ⊥ Catalan Institution for Research and Advanced Studies (ICREA), Passeig Lluïs Companys, 23, 08010, Barcelona, Spain S * Supporting Information ABSTRACT: Incorporation of biotinylated aminopyridine cobalt complexes derived from the triazacyclononane scaffold into the streptavidin protein leads to formation of artificial metalloenzymes for water reduction to hydrogen. The synthesized artificial metalloenzymes have lower overpotential (at the half-peak up to 100 mV) and higher photocatalytic hydrogen evolution activity (up to 14and 10-fold increase in TOF and TON, respectively, at pH 12.5) than the free biotinylated cobalt complexes. 1H-NMR, EPR and XAS highlight the presence of the metal complexes upon supramolecular attachment to the streptavidin. pHdependent catalytic studies and molecular dynamics (MD) simulations suggest that the increase in the catalytic activity could be induced by the protein residues positioned close to the metal centers. These findings illustrate the ability of the biotin−streptavidin technology to produce artificial metalloproteins for photo- and electrocatalytic hydrogen evolution reaction. KEYWORDS: water reduction to hydrogen, photocatalysis, electrocatalysis, cobalt complexes, streptavidin−biotin 1. INTRODUCTION coordination sphere suffers from significant synthetic challenges, although it is crucial to improve the TOF.6,7 In this regard, hybrid natural−artificial systems could serve as testing grounds to study the effect of protein scaffold encapsulation and second-sphere interactions in the catalytic performance of coordination complexes.3b,8 Several artificial protein−catalysts hybrids have been developed either by assembling synthetic catalysts into the protein pocket9 or by creating biomimetic systems made by metal substitution and/ or mutation.10 One of the most convenient synthetic approaches builds on supramolecular host−guest interactions, for example, using streptavidin (SA) as a host protein and exploiting its high affinity for biotin (also known as vitamin B7 or vitamin H). SA binds biotin with one of the strongest noncovalent interactions known in nature (dissociation constant (Kd) on the order of ∼10−14), ensuring quantitative incorporation of biotinylated molecules.11 In addition, the biotin−SA complex is resistant to organic solvent, temperature, and pH. On the basis of a concept first reported by Whitesides12 and further developed by the Ward group, the biotin−SA technology has been applied to create artificial Among the possible alternative fuels, molecular hydrogen obtained from light-driven water splitting is ideal from an environmental point of view.1 However, sustainable generation of hydrogen is extraordinarily challenging and requires a profound understanding of the different reactions involved. In this context, rational design of catalytic systems serves as a tool to interrogate strategies to maximize efficiency and robustness as well as to unravel the mechanistic details.2 Likewise, nature has served as inspiration by studying and mimicking hydrogenase enzymes, which reversibly catalyze the reduction of protons to molecular hydrogen with high efficiency (turnover frequencies of up to 21 000 s−1).3 The high activity of hydrogenases is ascribed in part to the control of the geometry of the active site, the second coordination sphere, and the protection of the catalytic center.4 However, their characterization and further development for practical applications is a demanding task. Conversely, catalysts active in H2 evolution based on metal complexes are tunable5 for improving the catalytic activity and overpotentials and increasing the oxygen tolerance. However, most organometallic catalysts are only efficient with strong acids in organic solvents or they are not sufficiently robust in water to be catalytically active. Moreover, control of the second © 2019 American Chemical Society Received: December 13, 2018 Revised: April 25, 2019 Published: May 13, 2019 5837 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis metalloenzymes for a broad variety of catalytic applications.13 Very recently, this strategy has also been explored in hydrogen evolution catalysis. It has been observed that the resulting artificial hydrogenases are able to drive H2 evolution when using [Ru(bpy)3]Cl2 as a photosensitizer and ascorbate as a sacrificial electron donor, albeit with no significant improvement over the “naked” catalyst.14 We recently described a new family of robust, well-defined, and highly tunable aminopyridine cobalt complexes derived from the triazacyclononane (tacn) scaffold for the electro- and photocatalytic H2 evolution in water.15 In view of these unique properties and inspired by the previous studies, here we study the encapsulation of two new biotinylated aminopyridine cobalt complexes in SA for photo- and electrocatalytic H2 production. We found that encapsulation increases the activity significantly under both sets of conditions. Detailed electroand photochemical pH-dependent hydrogen evolution studies and molecular dynamics (MD) simulations have been performed to elucidate the enhanced catalytic activity observed. ligands AmidePy2tacn and Amide‑SPPy2tacn was carried out by acylating the free nitrogen atoms of H Py 2 tacn and NH2‑SP Py2tacn, respectively (section S.1.3). 1H NMR spectroscopy, high-resolution electrospray ionization mass spectrometry (HR-MS-ESI), magnetic measurements, and elemental analysis confirm formation of d7 high-spin (S = 3/2) CoII complexes (Figures S1−S7). The cyclic voltammogram (CV) of 1Bio and 1Amide in CH3CN shows a reversible peak ascribable to the CoII/I couple at −1.09 V (all redox processes are reported vs SCE unless otherwise notified) (Figures S10 and S12). Moreover, the reduction potential for the CoII/I transition in 2Bio (−1.06 V) is close to that of 2Amide (−1.15 V) (Figures S11 and S13). We note that complex 2Bio exhibits an additional irreversible wave at −0.71 V, which can be attributed to reduction of the biotin moiety since it is also observed in the CV of the free Bio‑SPPy2tacn ligand and in the biotin (Figures S14−S17). Incorporation of 1Bio and 2Bio into SA is quantitative, as indicated by the HABA (2-(4′-hydroxyazobenzene)benzoic acid) displacement assay.16 The linear profile of the HABA displacement by the biotinylated complexes indicates that their affinity for SA is similar to previously reported systems and in the range of that obtained with D-biotin (i.e., >1010 M−1, section S.1.5).16 The stoichiometry of the complex incorporated in the synthesized and isolated metalloenzymes 1Bio C SA and 2Bio C SA was evaluated by measuring the Co concentration by ICP-MS and the protein concentration by UV−vis. These results are in concordance with previous reports, indicating saturation of the binding sites of the SA employed (SA is a tetramer containing a binding site per unit of monomer). Moreover, the paramagnetic 1H NMR studies performed with free SA, 2Bio, and 2Bio C SA indicate that the main proton resonance features of the cobalt complex are present with slight modifications after encapsulation into the SA and that some of the 1H NMR features of the SA protein are also altered in the 2Bio C SA sample, consistent with the supramolecular assembly (Figures S22−S25). We also studied cobalt K-edge X-ray absorption spectroscopy (XAS) and EPR to provide information about the electronic and geometric structure of the metal coordination environment. The Co K-edge XAS spectra of 2Bio and 2Bio C SA have a comparable profile, suggesting a similar coordination environment and oxidation state. The rising edges centered around 7719.3 eV are consistent with a CoII center (Figure S26).18 Furthermore, the intensity of the pre-edges at 7710.4 eV, of 0.06 normalized units, suggests a centrosymmetric pseudo-octahedral environment.18a This is further supported by EPR experiments of 2Bio and 2Bio C SA, which at liquid nitrogen temperatures show low-spin CoII signals with gx and gy values centered around a g of 2.2 and gz values approaching 2.0, similar to previously reported values for elongated distorted pseudo-octahedral CoII complexes (Figure S27.B).6,19 On the other hand, the EPR spectra of 1Bio and 1Bio C SA suggest an axially compressed pseudo-octahedral environment at the CoII center with gz values centered around a g of 2.3 and gx and gy values close to 2.0 (Figure S27.A).6,19 At 15 K the EPR spectra are more informative. The basic EPR features observed in the [Co(OTf)(TsPy2tacn)](OTf) (2), [Co(Cl)(TsPy2tacn)](Cl) 2Cl, 1Bio and 2Bio series guide one to understand the EPR of 1Bio C SA and 2Bio C SA (Table S2). Interestingly, while the EPR spectrum of 2Bio C SA is similar to that of 2Bio, in the case of 1Bio C SA it is essentially different than 1Bio. The EPR spectrum of 2Bio C SA recorded at 15 K presents three 2. RESULTS AND DISCUSSION We synthesized and studied two new biotinylated cobalt complexes (CatBio) based on the 1,4-di(picolyl)-1,4,7triazacyclononane (HPy2tacn) moiety by employing two different rigid spacer groups for the covalent bond between the biotin and the catalytic center (Figure 1)15a in order to investigate the catalytic activity of the cobalt center when positioned inside the protein pocket.13f,16 The biotin moiety was covalently attached to the tacn backbone either via the secondary nitrogen atom on the tacn (BioPy2tacn) or via a free amine on the spacer group directly linked to the tacn (Bio‑SPPy2tacn) through reaction with biotin N-hydroxysuccinimide active ester (NHS-biotin) (Scheme S1). Figure 1. Biotinylated cobalt complexes and artificial metalloenzymes employed in this work. The biotin-functionalized aminopyridine ligands and complexes [Co(BioPy2tacn)(MeCN)](OTf)2 (1Bio) and [Co(Bio‑SPPy2tacn)(MeCN)](OTf)2 (2Bio) (Figure 1) were synthetized adapting a general procedure described in the literature (sections S.1.3 and 1.4).15a,17 The homologous nonbiotinylated [Co(AmidePy2tacn)(MeCN)](OTf)2 (1Amide) and [Co(Amide‑SPPy2tacn)(MeCN)](OTf)2 (2Amide) complexes were also synthesized with the aim to study how the absence of biotin affects their behavior (Scheme S1). The synthesis of the 5838 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis different Co center environments as for 2Bio (Figure S33). The predominant one corresponds to a CoII high-spin center (S = 3/2), and two minor signals can be assigned to low-spin CoII centers coordinated to a different sixth ligand (CH3CN and H2O). In contrast, the main feature in the EPR spectrum of 1Bio C SA correspond to the rhombic signal with g values centered at 2.37, 2.07, and 2.03, which is in agreement with coordination of a water molecule. This correlates with the fact that the cobalt complex in 1Bio C SA is more internalized into the protein cavity as judged by the molecular dynamics (MD) simulations (dicussed below). Therefore, the presence of mainly a low-spin CoII rhombic signal may suggest that the protein pocket environment may regulate the binding ligand at the metal center. In the case of the 2Bio C SA, MD simulations show that the cobalt complex is exposed to the protein surface, but it is much more exposed to the reaction media than 1Bio C SA. For both series of complexes the XAS and EPR spectroscopic data indicate only minor electronic and geometric environment modification of the CatBio after inclusion into the SA protein. Lastly, incorporation of the cobalt catalyst proceeds, conserving the SA secondary structure, as assessed by CD spectroscopy (Figure S21). We compared the catalytic activity of these artificial metalloenzymes with the naked biotinylated cobalt complexes 1Bio and 2Bio in the electrochemical reduction of water to H2. The electrochemical measurements were performed at pH ranging from 3.5 to 6 (citrate/NaCl buffer, 100 mM). At pH 3.5 the CV of 1Bio showed a high current intensity irreversible peak with an onset potential at −1.21 V and a maximum at −1.36 V (overpotential (η) = 0.78 V)20 (Figure 2A, black trace), indicative of a catalytic process. Under the same conditions, complex 2Bio showed similar electrochemical features but shifted to a more negative potential (onset potential about −1.23 V and a maximum at −1.47 V, η = 0.79 V, Figure 2C). The electrocatalytic wave in both complexes (1Bio and 2Bio) displays a pH dependence. The currents appear to follow a nonstandard behavior in the case of 2Bio with a change in slope as the potential is increased. The electrocatalytic peak current (i c) increases with the proton concentration (Figures S35 and S36), suggesting that the rate-determining step of the catalytic process relies on a proton transfer from water.21 The onset, the peak potential for proton reduction (Ep), and the half peak (E1/2) of 1Bio shift toward less negative potentials when increasing the proton concentration, following a PCET (proton-coupled electron transfer) reaction (Figure 2 and Figure S46).22 The Nernstian response was found to be −37 mV per pH unit (Figure 2B), close to expected for a 2-electron and 1-proton reduction process (−30 mV per pH). This value is qualitatively similar to the −26 mV per pH unit reported for a Co(Py5) system21b or to the −24 mV per pH unit observed in cobalt diamine systems.21a Complex 2Bio reveals only a minor dependence of the Ep and onset potential with the pH (Figure 2D). The molecular behavior of 1Bio and 2Bio was confirmed by rinsing tests (Figure S36), which indicates the absence of active species attached to the electrode surface.23 Furthermore, the catalytic current shows a first-order dependence on the catalyst concentration24 (Figure S37) similarly to what was observed for the parent complex 2.15a The electrochemical properties of CatBio C SA were interrogated as a thin film,3a,9c,10a,25 prepared by drop casting over a glassy carbon electrode and carbon cloth (see sections S.1.7 and S.1.8), which yields an intense electrocatalytic wave Figure 2. CV of (A) 1Bio (1 mM, black trace) and 1Bio C SA (thin film, red trace) and (C) 2Bio (1 mM, black trace) and 2Bio C SA (thin film, red trace) (see section S.1.7 for detailed experimental conditions). Plot of the half intensity of the catalytic peak potential (Ep) as a function of pH of (B) 1Bio (1 mM) and 1Bio C SA (thin film) and (D) 2Bio (1 mM) and 2Bio C SA (thin film). (Figure S38). For comparison, we also analyzed and compared the CV of 2Bio C SA (73 μM) in solution, which gives similar results under the same experimental conditions (Figure S39). When the cathodic scan was performed in the presence of naked SA the observed current was negligible at this redox potential (Figure S38), indicating that the catalytic wave observed for CatBio C SA is due to the hosted CatBio. The reduction potential of the reduction peak did not change with the amount of protein deposited on a carbon cloth electrode, suggesting that protein−protein interactions possible in the film do not interfere in the redox process (Figure S40). The catalytic activity of the CatBio C SA adducts display a pH dependence (Figures S42 and S43) comparable to that obtained with the CatBio and CatAmide catalysts (Figures S35, S36, S44−S48, S55, and S56). A clear shift in the redox potential of the catalytic value (ΔEp, ΔEp/2, and onset potential) was observed when the CatBio were incorporated into the host SA. The magnitude of the redox potential shift when the current intensity is one-half of the value regarding the current intensity at the maximum of the catalytic peak (ΔEp/2) varies with the pH value and ranges between 20 and 30 mV for 1Bio and 1Bio C SA (Figure 2B) and between 60 and 100 mV for 2Bio and 2Bio C SA (Figure 2D) (Figures S44− S46). This result is in sharp contrast with those previously reported for some artificial metalloenzymes based on cobalt complexes.9c,10a,e For instance, incorporating cobaloxime derivatives into the heme pocket of apo sperm-whale myoglobin results in a negative shift of about 100 mV, increasing the overpotential.9c Conversely, the positive direction of the shift observed for our systems suggests beneficial interactions of the SA pocket with the catalyst (presumably the secondary coordination sphere) that reduces the water reduction overpotential. However, a direct comparison of current density is not straightforward since the protein concentration on the electrode is unknown. To verify the formation of H2 we performed controlled potential electrolysis (CPE) at −1.4 V during 3 h under static conditions for 2Bio and 2Bio C SA (Figures S49−S51). In both 5839 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis Figure 3. Photoinduced H2 production at different pH values for reaction of CatBio and CatBio C SA (5 μM in Co), PSIr (250 μM), TEOA (0.4 M) in 0.4 mL of CH3CN:H2O (3:7 v/v) containing phosphate buffer (100 mM) and NaCl (100 mM) after 4 h irradiation (1100 W·m−2, λ > 400 nm). Time traces of H2 evolved using (A) 2Bio and (B) 2Bio C SA. (C) TON for 2Bio (gray) and 2Bio C SA (black) and TON2Bio C SA/TON2Bio (red). (D) Initial TOF (h−1) for 2Bio (gray), 2Bio C SA (black), and TOF2Bio C SA/TOF2Bio (red). (E) TON for 1Bio (gray) and 1Bio C SA (black) and TON1Bio C SA/TON1Bio (red). (F) Initial TOF (h−1) for 1Bio (gray), 1Bio C SA (black), and TOF1Bio C SA/TOF1Bio (red). Error bars are average of two duplicates. biotin interactions appear negligible as there is a linear correlation between the current intensity and the CatBio concentration (vide supra). All of these observations together with the XAS, EPR and catalysis show no coordination of biotin to the cobalt center and indicate that the ΔEp observed between CatBio and CatBio C SA arises form second coordination sphere interactions between the catalytic center and the protein pocket. We also investigated whether the catalytic activity of 1Bio and Bio 2 could be influenced by protein encapsulation in lightdriven H2 evolution. The catalytic activity of 2Bio and 2Bio C SA was studied by irradiation (1100 W·m−2 of visible light, λ > 400 nm) of a solution containing the catalysts (5 μM), [Ir(ppy)2(bpy)]PF6 (PSIr) (250 μM) and TEOA (0.4 M).26,15a Due to the low solubility of PSIr in water, 30% in volume of MeCN was used as a cosolvent. We verified the stability of 2Bio C SA by monitoring its secondary structure by CD at increasing concentrations of CH3CN. No changes in the CD were observed up to 50% of cosolvent (Figures S57 and S58) nor in all of the pH range studied (Figures S59−S62). Hydrogen evolution was monitored as a function of time by analysis of aliquots from the reaction cuvette headspace by gas chromatography at different pH values adjusted with phosphate buffer (100 mM). The pH of the buffer solutions was measured and adjusted (when needed) after addition of TEOA using NaOH or HCl. The catalytic activity of 2Bio raises from pH 7 to a maximum at about pH 10.5 (TON, 89; TOF, 49·h−1; Figure 3).27 For 2Bio C SA the catalytic activity was higher than for 2Bio at all pH values, and the maximum in catalytic activity changes from pH 10.5 to 11.5 (Table S4). A pH bell-shaped profile is observed for the catalytic activity of both 2Bio and 2Bio C SA (Figure 3). This bell shape has been previously observed in photocatalytic systems such as Ru/ ascorbate28 and fluorescein/Et3N.29 The stablished reason for this catalytic activity pH behavior is that at lower pH the cases H2 was detected by sampling the headspace of the electrochemical cell after the CPE and subjecting the sample to chromatographic analysis (gas chromatography-thermal conductivity detector, GC-TCD). In the case of 2Bio C SA, after an initial current decrease it remains about constant for the rest of the experiment with good Faraday yield. In contrast, under the same CPE conditions, 2Bio yields about 560 TON of H2 with good Faraday yield, but suffers a fast and intense activation process during the CPE, leading to a higher current density (Figure S51). This is indicative of the formation of a more active species. A CV after the CPE with a clean GC electrode did not show any wave in the CV at the CoII/I redox couple 2Bio. In contrast, after 3 h of CPE the CV of the 2Bio C SA solution presents a wave, although shifted and significantly reduced. We hypothesized that the observed difference in activity between CatBio and CatBio C SA is due to secondary sphere interactions with the protein scaffold. To rule out other possibilities, such as coordination of biotin to the cobalt center, we also studied the electrochemical behavior of model complexes 1Amide and 2Amide. We found that 1Amide and 2Amide exhibit similar electrochemical behaviors (CoII/I redox couples and pH behavior) to the CatBio homologous complexes. In addition, the cyclic voltammograms of 1Amide and 2Amide in CH3CN shows negligible changes upon addition of large amounts of esterified biotin (up to 20 equiv, biotin was esterified for solubility reasons, Figures S53 and S54). We also studied computationally (B3LYP-3D/6−31+G** level of theory) the possible coordination modes for the N, O, and S atoms of biotin to the cobalt center in oxidation states I and II. All optimized structures showed that intramolecular coordination of biotin to the cobalt center is largely endergonic. In addition, an exchange between a coordinated pyridine (or tacn backbone) and a N, O, or S atom of biotin is also highly endergonic (Figures S79−S81 and S84−S86). The bimolecular 5840 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis effective electron donor concentration (TEOA) is reduced, while at higher pH values the catalytic activity decreases due to the lower concentration of protons available, resulting in the catalytic activity vs pH in a bell-shaped profile. The difference in the catalytic activity between 2Bio and 2Bio C SA increases at higher pH values. At pH 12.5 the catalytic activity of 2Bio C SA is 10- and 14-fold higher in terms of TON and TOF, respectively, than the values obtained for 2Bio. Moreover, at pH 12.5 the catalytic activity of 2Bio is negligible. This behavior may be rationalized if it is considered that the protein protects the cobalt center and alters the effective available protons for the metal protonation regarding the pH. Nevertheless, this is a very complex system, and other possible contributions to the observed behavior cannot be discarded. It is important to note that protonation of the CoI species to give CoIII−H is thought to be the rate-limiting step in cobalt-based catalysts at basic pH values.15a Although 1Bio shows lower catalytic activity than 2Bio, the catalytic enhancement by the protein environment is also remarkable with an enhancement factor in TON at pH 11.5 of 7.3 (1Bio C SA) and 3.6 (2Bio C SA), suggesting a better protecting effect by the protein pocket due to the absence of the spacer (Figure 3E and 3F and Table 1). Moreover, control Figure 4. Traces of the photoinduced H2 production by (from top to bottom) 2Bio C SA, 1Bio C SA, 2Bio, and 1Bio (5 μM) together with PSCu (250 μM), Et3N (0.15 M) in 0.4 mL of CH3CN:H2O (3:7 v/v) after 9 h irradiation (1100 W·m−2, λ > 400 nm). a high affinity of the protein toward 1Bio and 2Bio. The presence of a spacer group between the biotin and the catalytic center in 2Bio confers additional flexibility to the ligand, leading to two main different conformations (Figure S70). As in biotin−SA complex, tryptophan residues (Trp79, Trp92, Trp108) bind the ligand through CH··· π interactions (the mean distance between the center of mass of Trp108 and biotin rings is ca. 4.0 Å for both 1Bio and 2Bio, see Figure 5).30 To gain insights into the origin of the improvement in catalytic activity, we evaluated the conformational dynamics of the binding pocket in the presence of 1Bio and 2Bio (Figure 5). Polar residues around the cobalt complex together with the assistance of water could promote the catalysis. Both 1Bio and 2Bio are exposed to several solvent molecules (ca. 3−5 waters in 1Bio and 2−4 waters in 2Bio, Figures S69 and S72). In addition, many polar residues that could act as a proton donor or activate a water molecule to promote metal protonation are found close to the ligand (Figure 5). 1Bio shows low flexibility during the MD simulations, sampling most of the time the same conformation, which positions the cobalt center near polar Ser112, Ser88, Glu51, and Asn49 residues (S67 and S68). In contrast, 2Bio is more flexible and adopts two different conformations during the MD simulations (the bend conformation highlighted in the square of Figure 5 and extended conformation of Mon3 in S34A). In the bend conformation, the metal center is situated close to polar residues Ser112 and Leu124 (Figure 5). For most of the simulation time the hydrophobic residue Leu124 in the bend conformation is near the metal center (Figures 5 and S71), suggesting that this residue could be mutated to enhance activity. In contrast, in the extended conformation, the metal center is close to the polar residues Asn49 and His87 (Figures S70E and S71). We investigated whether the flexibility of the metal complexes is restricted upon encapsulation by performing MD simulations for 1Bio and 2Bio in solution. In the absence of protein, both complexes have a certain flexibility. However, complex 2Bio, with the longer link, is the most flexible one, as indicated by the computed root-mean square deviations (RMSD, see Figure S74). The higher flexibility of 2Bio is also retained upon protein encapsulation (compare RMSD of 1Bio and 2Bio inside the protein, Figures S75 and S76), which allows a higher water accessibility, consistent with the water shell and g(r) analysis (Figure S73). Table 1. Photocatalytic Activities of 2Bio and 2Bio C SA at pH 11.5a 1Bio TON TONCatBio C SA/TONCatBio TOF TOFCatBio C SA/TOFCatBio 7 7.3 1.8 5.7 1Bio C SA 51 10.3 2Bio 2Bio C SA 46 3.6 16 5.5 166 88 Cobalt catalyst (5 μM in Co) PSIr (250 μM), TEOA (0.4 M) in 0.4 mL of CH3CN:H2O (3:7 v/v) containing phosphate buffer (100 mM) and NaCl (100 mM) (pH 11.5) after 4 h irradiation (1100 W· m−2, λ > 400 nm). a experiments performed generating the metalloenzyme in situ before catalysis (by mixing 2Bio with SA ∼30 min before the catalysis) show the same catalytic activity as the presynthesized and purified 2Bio C SA (Figure S65). In contrast, addition of SA to complex 2Amide does not affect its photocatalytic activity (Figure S66), pointing toward an enhancement in the H2 production activity due to supramolecular interactions in 2Bio C SA. We also performed photochemical studies using [Cu(xantphos)(bathocuproine)]PF6 (PSCu) as photoredox catalyst and triethylamine (Et3N) as electron donor in nonbuffered solutions (Figure 4). Under these catalytic conditions we found that the H2 evolution catalytic performance for the metalloproteins (CatBio C SA) is higher than for the CatBio in terms of both TOF and TON. Also, a remarkably higher quantity of hydrogen is formed than in the case of using the Ir photosensitizer (PSIr) in combination with TEOA. We performed molecular dynamics (MD) simulations of streptavidin in complex with 1Bio and 2Bio to evaluate their binding modes and identify which residues play a crucial role on catalytic enhancement (see SI for computational details). These calculations do not capture protein−protein interactions and therefore are only relevant for isolated proteins such as the conditions employed in this study. Representative conformations sampled along the accumulated 750 ns of MD simulations (i.e., three replicas of 250 ns for each case) reveal 5841 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis Figure 5. Overlay of representative snapshots visited along the MD simulations for 1Bio and 2Bio. Conformational dynamics of the binding pockets are also shown, where the most important residues surrounding the biotin complexes are represented together with some selected distances (in Angstroms). Distances have been computed from the metal center and either the side chain of Ser112, Ser88 (oxygen atom, in purple and green, respectively), and Leu25 (terminal carbon of isobutyl side chain, orange). Trp108−biotin distance is computed between the center of mass of the indole ring of Trp108 and the tetrahydrothiophene ring of biotin (in blue). Mean distances and standard deviations are also included (in Angstroms). Gas Chromatography. Gas chromatography was carried out on a SRI instruments, model 310C, GC using a 5 Å molecular sieve column with a thermal conductivity detector and argon carrier gas. Nuclear Magnetic Resonance. All NMR spectra were recorded in CDCl3 on a Varian 400 or 500 MHz instrument, as noted. Circular Dichroism. Spectra measurements were carried out on an Applied Photophysics Chirascan Circular Dichroism spectrometer equipped with a photomultiplier detector, dualpolarizing prism design monochromator, photoelastic modulator (PEM), and 150 W xenon light source. An average of 3 scans per sample were measured. X-ray Diffraction. Crystals of 1Amide were grown by slow diffusion of Et2O to a CH2Cl2 solution of the complex. The crystals for these samples were selected using a Zeiss stereomicroscope using polarized light and prepared under inert conditions immersed in perfluoropolyether as protecting oil for manipulation. Crystal structure determinations for 1Amide were carried out using a Apex DUO Kappa 4-axis goniometer equipped with an APPEX 2 4K CCD area detector, a Microfocus Source E025 IuS using Mo Kα radiation, Quazar MX multilayer optics as monochromator, and an Oxford Cryosystems low-temperature device Cryostream 700 plus (T = −173 °C). On the Bruker device, the following programs were used: Data collection APEX-2,32 data reduction Bruker Saint33 V/.60A, and absorption correction SADABS34 or TWINABS.35 The crystal structure solution was achieved using the computer program SHELXT.36 Visualization was performed with the program SHELXle.37 Missing atoms were subsequently located from difference Fourier synthesis and added to the atom list. Least-squares refinement on F2 using all measured intensities was carried out using the 3. CONCLUSIONS In conclusion, we prepared new cobalt artificial metalloenzymes based on the biotin−streptavidin technology and able to reduce water to hydrogen under electro- and photochemical conditions. We found that incorporation of the cobalt catalysts into the streptavidin scaffold results in several beneficial effects, especially in the 2Bio catalyst. It reduces the overpotential for electrocatalytic hydrogen production and increases the light-driven hydrogen production rate and turnover numbers. This work demonstrates a straightforward and successful strategy to enhance the catalytic activity for water reduction to H2. Future studies will be performed to extend such cobalt artificial metalloenzymes to various other photocatalytic chemical transformations. 4. EXPERIMENTAL SECTION Materials and Reagents. Reagents and solvents were purchased from commercial sources and used as received unless otherwise stated. Triethylamine (Et3N, ≥99% purity) and triethanolamine (TEOA, ≥99% purity) were purchased from Sigma-Aldrich and used without further purification. Photosensitizer [Ir(bpy)(ppy)2]PF6 (PSIr) was synthesized according to literature procedures.26a Ethyl [(3aS,4S,6aR)-2oxo-hexahydrothieno[3,4-d]imidazole-4-yl]pentanoate (COOEtBiotin) was synthesized following the reported method from the literature.31 Anhydrous acetonitrile was purchased from Scharlab. Water (18.2 MΩ·cm) was purified with a MilliQ Millipore Gradient AIS system. All solvents were thoroughly degassed and stored under anaerobic conditions. Physical Methods. UV−vis Spectroscopy. UV−vis spectra were acquired on a Varian Cary 50 Bio Spectrophotometer. 5842 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis program SHELXL 2015.38 All non-hydrogen atoms were refined including anisotropic displacement parameters. Cyclic Voltammetry. Electrochemical experiments were carried out on a CH instruments 1242B potentiostat and with a VSP potentiostat from Bio-Logic, equipped with the EC-Lab software. The experiments were performed under inert (Ar) atmosphere in a custom double-wall jacketed singlecompartment cell. For all electrochemical measurements, a three-electrode system was used: a 3 mm diameter glassy carbon working electrode with a surface area of 0.28 cm2, a platinum mesh counter electrode, and a saturated calomel reference electrode, unless otherwise specified. Working electrodes were polished with 1 μM alumina for 5 min, followed by 10 min of sonication, prior to use. Cyclic voltammetry was carried out on samples prepared in 100 mM citrate and 25 mM NaCl buffer of desired pH (3.5, 4.0, 4.5, 5.0, 5.5, and 6.0); scans were performed at a 100 mV·s−1 scan rate. All potentials are quoted versus SCE. All electrolyte solutions were degassed by freeze and pump techniques. Samples were prepared by diluting the catalyst from a stock solution (10 mM) in water containing 5% DMSO to solubilize the catalyst. The final percentage of DMSO in the electrochemical cell was 0.5%. Electrochemical experiments CatBio C SA were done by preparing a thin film of CatBio C SA over the glassy carbon electrodes (0.28 cm2 surface area) by drop casting of 10 μL of a stock solution of metalloprotein of 10 μM and drying, otherwise indicated. After drying, electrochemical measurements were performed by directly submerging the glassy carbon electrode into the solution. Electron Paramagnetic Resonance. An EMX Micro Xband EPR spectrometer from Bruker was used to collect data on 0.35 mM solution samples in a mixture of aqueous buffer and MeCN (7:3, buffer = Tris·HCl 100 mM, pH 7.5) using a finger Dewar at 77 K and He cryostat from Oxford Instruments at 50 and 15 K. Data was acquired in perpendicular mode with a modulation frequency of 100 KHz, a modulation amplitude of 10 G, a 10.24 ms time constant, and 19.02 ms conversion time with a microwave power was 0.1851 mW. Spectra were simulated using the EasySpin software package.39 X-ray Absorption Spectroscopy. All NM Samples were recorded as 0.35 mM solutions in a mixture of aqueous buffer and MeCN (7:3, buffer = Tris·HCl 100 mM, pH 7.5) at the SOLEIL synchrotron (2.75 GeV, 400mA storage ring) SAMBA Beamline equipped with a helium cryostat (20 K) and using a Si(220) double-crystal monochromator together with a 36 channel Ge detector. Data calibration and normalization was carried out using the Athena software package.40 Energies were calibrated to the first inflection point of Co foil spectra set at 7709.5 eV. A linear pre-edge function and a quadratic polynomial for the postedge were used for background subtraction and normalization of the edge jump. Photoinduced H2 Evolution. Photoinduced H2 production irradiation was performed using a 450W xenon lamp with a 400 nm cutoff filter, irradiating at a constant 1100 W/m2 throughout the experiment. For each experiment, 400 μL of a reaction mixture containing all of the reagents was added to an airtight 1 mm cuvette and degassed extensively with argon prior to illumination. During irradiation time course 100 μL samples of the headspace were removed with a gastight syringe and injected directly for analysis by GC. A GC calibration curve was obtained by injecting various volumes of a 1% H2, 99% N2 gas mixture. The pH of the buffer solutions was measured and adjusted (if needed) after addition of TEOA by addition NaOH or HCl. Theoretical Calculations. DFT Calculations. We studied the coordination ability of the biotinated ligands BioPy2tacn and Bio‑SPPy2tacn to the metal center and the possibility of binding the biotin fragment to the metal center in complexes 1Bio and 2Bio computationally by density functional theory (DFT) using the Gaussian09 program.41 Geometry optimizations were performed in the unrestricted spin formalism, with the B3LYP hybrid exchange-correlation functional42 and the standard 6-31+G** 6d basis set for all atoms. An extra quadratic convergent SCF step was added when the first-order SCF did not converge (“scf = xqc” keyword). The solvation effect of acetonitrile was introduced in geometry optimizations and energy through the IEFPCM-SMD polarizable continuum model.43 Dispersion effects were also included using the Grimme D3 correction.44 The geometries have been edited with the Chemcraft program.45 The located stationary points were characterized by analytical frequency calculations at the same level of theory as geometry optimizations. Gibbs energy values (G) were obtained by including thermal, solvation, and Grimme corrections to the potential energy computed with the 6-31+G** 6d basis set on equilibrium geometries. Molecular Dynamics Simulations. MD simulations in explicit water were performed using the AMBER 16 package46 on our in-house GPU cluster Galatea. 1Bio and 2Bio parameters for the MD simulations were generated within the antechamber module of AMBER 16 using the general AMBER force field (GAFF).47 We used the bonded model for the Co (quadruplet state as the lowest in energy) and ligand moiety; in particular, we used the Seminario method approach to obtain the metal parameters needed for the simulation as implemented in Prof. Ryde’s program.48 The optimization, frequencies, and partial charge were set to fit the electrostatic potential generated at the UB3LYP/6-31G(d) level by the restrained electrostatic potential (RESP) model using Gaussian 09. The charges were calculated according to the Merz−Singh−Kollman scheme. We docked 1Bio and 2Bio complexes in the biotin pocket using the pair-fitting tool in Pymol (http://www.pymol.org) using the biotin moiety of the crystal structure (PDB 1MK5) as a reference and the optimized structure of the complex creating the tetramer enzyme. Amino acid protonation states and the data needed for the electrostatic map were set up using the Propka3.0 server (http://nbcr-222.ucsd.edu/pdb2pqr_2.0.0/ ).49Then the enzyme was solvated in a pre-equilibrated truncated cuboid box with a 10 Å buffer of TIP3P50 water molecules using the AMBER16 leap module, resulting in the addition of ca. 9000 solvent molecules. The systems were neutralized by addition of explicit counterions (Na+ and Cl−). All subsequent calculations were carried out using the widely tested Stony Brook modification of the Amber 99 force field (ff14SB).51 The structure used was 1MK5 from the protein data bank (PDB). A two-stage geometry optimization was performed. The first stage minimizes the positions of solvent molecules and ions imposing positional restraints on solute by a harmonic potential with a force constant of 500 kcal mol−1 Å−2, and the second stage is an unrestrained minimization of all of the atoms in the simulation cell. The systems are gently heated using six 50 ps steps, incrementing the temperature 50 K each step (0−300 K) under constant volume and periodic boundary conditions. Water molecules were treated with the SHAKE algorithm such that the angle between the hydrogen atoms is 5843 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis kept fixed. Long-range electrostatic effects were modeled using the particle-mesh-Ewald method.52 A 8 Å cutoff was applied to Lennard−Jones and electrostatic interactions. Harmonic restraints of 10 kcal·mol−1 were applied to the solute, and the Langevin equilibration scheme was used to control and equalize the temperature. The time step was kept at 1 fs during the heating stages, allowing potential inhomogeneities to self-adjust. Each system was then equilibrated without restrains for 2 ns with a 2 fs time step at a constant pressure of 1 atm and temperature of 300 K. After the systems were equilibrated in the NPT ensemble, 3 independent 250 ns MD simulations for 1Bio and 2Bio (i.e., 750 μs accumulated) were performed under the NVT ensemble and periodic-boundary conditions. ■ Supercomputing Center-Centro Nacional de Supercomputación. ■ (1) (a) Barber, J.; Tran, P. D. From Natural to Artificial Photosynthesis. J. R. Soc., Interface 2013, 10, 20120984. (b) Lewis, N. S.; Nocera, D. G. Powering the Planet: Chemical Challenges in Solar Energy Utilization. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 15729−15735. (2) (a) McKone, J. R.; Marinescu, S. C.; Brunschwig, B. S.; Winkler, J. R.; Gray, H. B. Earth-abundant Hydrogen Evolution Electrocatalysts. Chem. Sci. 2014, 5, 865−878. (b) Du, P.; Eisenberg, R. Catalysts Made of Earth-abundant Elements (Co, Ni, Fe) for Water Splitting: Recent Progress and Future Challenges. Energy Environ. Sci. 2012, 5, 6012−6021. (c) Artero, V.; Chavarot-Kerlidou, M.; Fontecave, M. Splitting Water with Cobalt. Angew. Chem., Int. Ed. 2011, 50, 7238−7266. (d) Le Goff, A.; Artero, V.; Jousselme, B.; Tran, P. D.; Guillet, N.; Metaye, R.; Fihri, A.; Palacin, S.; Fontecave, M. From Hydrogenases to Noble Metal-free Catalytic Nanomaterials for H2 Production and Uptake. Science 2009, 326, 1384−1387. (3) (a) Lubitz, W.; Ogata, H.; Rudiger, O.; Reijerse, E. Hydrogenases. Chem. Rev. 2014, 114, 4081−4148. (b) Simmons, T. R.; Berggren, G.; Bacchi, M.; Fontecave, M.; Artero, V. Mimicking Hydrogenases: From Biomimetics to Artificial Enzymes. Coord. Chem. Rev. 2014, 270−271, 127−150. (c) Tard, C.; Pickett, C. J. Structural and Functional Analogues of the Active Sites of the [Fe]-, [NiFe]-, and [FeFe]-Hydrogenases. Chem. Rev. 2009, 109, 2245−2274. (4) (a) Mulder, D. W.; Shepard, E. M.; Meuser, J. E.; Joshi, N.; King, P. W.; Posewitz, M. C.; Broderick, J. B.; Peters, J. W. Insights into [FeFe]-hydrogenase Structure, Mechanism, and Maturation. Structure 2011, 19, 1038−1052. (b) Knörzer, P.; Silakov, A.; Foster, C. E.; Armstrong, F. A.; Lubitz, W.; Happe, T. Importance of the Protein Framework for Catalytic Activity of [FeFe]-hydrogenases. J. Biol. Chem. 2012, 287, 1489−1499. (c) Berggren, G.; Adamska, A.; Lambertz, C.; Simmons, T. R.; Esselborn, J.; Atta, M.; Gambarelli, S.; Mouesca, J. M.; Reijerse, E.; Lubitz, W.; Happe, T.; Artero, V.; Fontecave, M. Aiomimetic Assembly and Activation of [FeFe]hydrogenases. Nature 2013, 499, 66−69. (d) Esselborn, J.; Lambertz, C.; Adamska-Venkatesh, A.; Simmons, T.; Berggren, G.; Noth, J.; Siebel, J.; Hemschemeier, A.; Artero, V.; Reijerse, E.; Fontecave, M.; Lubitz, W.; Happe, T. Spontaneous Activation of [FeFe]-hydrogenases by an Inorganic [2Fe] Active Site Mimic. Nat. Chem. Biol. 2013, 9, 607−609. (5) (a) Zee, D. Z.; Chantarojsiri, T.; Long, J. R.; Chang, C. J. Metalpolypyridyl Catalysts for Electro- and Photochemical Reduction of Water to Hydrogen. Acc. Chem. Res. 2015, 48, 2027−2036. (b) Queyriaux, N.; Jane, R. T.; Massin, J.; Artero, V.; ChavarotKerlidou, M. Recent Developments in Hydrogen Evolving Molecular Cobalt(II)-polypyridyl Catalysts. Coord. Chem. Rev. 2015, 304−305, 3−19. (6) (a) Jian, J. X.; Liu, Q.; Li, Z. J.; Wang, F.; Li, X. B.; Li, C. B.; Liu, B.; Meng, Q. Y.; Chen, B.; Feng, K.; Tung, C. H.; Wu, L. Z. Chitosan Confinement Enhances Hydrogen Photogeneration from a Mimic of the Diiron Subsite of [FeFe]-hydrogenase. Nat. Commun. 2013, 4, 1− 9. (b) Singleton, M. L.; Reibenspies, J. H.; Darensbourg, M. Y. A Cyclodextrin Host/Guest Approach to a Hydrogenase Active Site Biomimetic Cavity. J. Am. Chem. Soc. 2010, 132, 8870−8871. (7) (a) Orain, C.; Quentel, F.; Gloaguen, F. Photocatalytic Hydrogen Production Using Models of the Iron-iron Hydrogenase Active Site Dispersed in Micellar Solution. ChemSusChem 2014, 7, 638−643. (b) Pullen, S.; Fei, H.; Orthaber, A.; Cohen, S. M.; Ott, S. Enhanced Photochemical Hydrogen Production by a Molecular Diiron Catalyst Incorporated into a Metal-organic Framework. J. Am. Chem. Soc. 2013, 135, 16997−17003. (c) Yu, T.; Zeng, Y.; Chen, J.; Li, Y. Y.; Yang, G.; Li, Y. Exceptional Dendrimer-based Mimics of Diiron Hydrogenase for the Photochemical Production of Hydrogen. Angew. Chem., Int. Ed. 2013, 52, 5631−5635. (8) (a) Faiella, M.; Roy, A.; Sommer, D.; Ghirlanda, G. De NovoDesign of Functional Proteins: Toward Artificial Hydrogenases. ASSOCIATED CONTENT S * Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscatal.8b04981. ■ REFERENCES Synthesis and characterization of ligands, catalysts, and SA adducts; determination of electro- and photocatalytic performance; DFT and molecular dynamic simulations (PDF) AUTHOR INFORMATION Corresponding Authors *E-mail: gghirlanda@asu.edu *E-mail: jlloret@iciq.es ORCID Arnau Call: 0000-0003-0970-7195 Sílvia Osuna: 0000-0003-3657-6469 Julio Lloret-Fillol: 0000-0002-4240-9512 Author Contributions ‡ A.C. and C.C.: These authors ontributed equally. Notes The authors declare no competing financial interest. ■ ACKNOWLEDGMENTS We thank the European Commission for the ERC-CG-2014648304 (J.Ll.-F) and ERC-2015-StG-679001 (S.O) projects. The Spanish Ministry of Science is acknowledged for a FPU fellowship to A.C. (AP2012-6436) and C.C. and the Generalitat de Catalunya for a Ph.D. fellowship to A.R.R. (2015-FI-B-00165). We also thank Catexel for a generous gift of tritosyl-1,4,7-triazacyclononane. Financial support from the ICIQ Foundation and CELLEX Foundation through the CELLEX-ICIQ high-throughput experimentation platform and the Starting Career Program is gratefully acknowledged. We also thank the CERCA Programme (Generalitat de Catalunya) for financial support and MINECO (Severo Ochoa Excellence Accreditation 2014−2018; SEV-2013-0319) and MINECO project CTQ2016-80038-R. S.O. thanks the Spanish MINECO CTQ2016-80038-R, Ramón y Cajal contract (RYC-201416846), the European Community for CIG project (PCIG14GA-2013-630978). This work was supported in part by NSF award 1508301 to G.G. We acknowledge SOLEIL and DIAMOND for provision of synchrotron radiation facilities, and we thank Dr. Gautier Landrot for assistance in using beamline SAMBA. We are grateful for the computer resources, technical expertise, and assistance provided by the Barcelona 5844 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis Biopolymers 2013, 100, 558−571. (b) Bos, J.; Browne, W. R.; Driessen, A. J. M.; Roelfes, G. Supramolecular Assembly of Artificial Metalloenzymes Based on the Dimeric Protein LmrR as Promiscuous Scaffold. J. Am. Chem. Soc. 2015, 137, 9796−9799. (9) (a) Soltau, S. R.; Niklas, J.; Dahlberg, P. D.; Poluektov, O. G.; Tiede, D. M.; Mulfort, K. L.; Utschig, L. M. Aqueous Light Driven Hydrogen Production by a Ru−ferredoxin−Co Biohybrid. Chem. Commun. 2015, 51, 10628−10631. (b) Onoda, A.; Kihara, Y.; Fukumoto, K.; Sano, Y.; Hayashi, T. Photoinduced Hydrogen Evolution Catalyzed by a Synthetic Diiron Dithiolate Complex Embedded within a Protein Matrix. ACS Catal. 2014, 4, 2645−2648. (c) Bacchi, M.; Berggren, G.; Niklas, J.; Veinberg, E.; Mara, M. W.; Shelby, M. L.; Poluektov, O. G.; Chen, L. X.; Tiede, D. M.; Cavazza, C.; Field, M. J.; Fontecave, M.; Artero, V. Cobaloxime-based Artificial Hydrogenases. Inorg. Chem. 2014, 53, 8071−8082. (d) Silver, S. C.; Niklas, J.; Du, P.; Poluektov, O. G.; Tiede, D. M.; Utschig, L. M. Protein Delivery of a Ni Catalyst to Photosystem I for Light-driven Hydrogen Production. J. Am. Chem. Soc. 2013, 135, 13246−13249. (e) Berggren, G.; Adamska, A.; Lambertz, C.; Simmons, T. R.; Esselborn, J.; Atta, M.; Gambarelli, S.; Mouesca, J. M.; Reijerse, E.; Lubitz, W.; Happe, T.; Artero, V.; Fontecave, M. Biomimetic Assembly and Activation of [FeFe]-hydrogenases. Nature 2013, 499, 66−69. (f) Utschig, L. M.; Silver, S. C.; Mulfort, K. L.; Tiede, D. M. Nature-driven Photochemistry for Catalytic Solar Hydrogen Production: a Photosystem I-transition Metal Catalyst Hybrid. J. Am. Chem. Soc. 2011, 133, 16334−16337. (10) (a) Kandemir, B.; Chakraborty, S.; Guo, Y.; Bren, K. L. Semisynthetic and Biomolecular Hydrogen Evolution Catalysts. Inorg. Chem. 2016, 55, 467−477. (b) Slater, J. W.; Shafaat, H. S. Nickelsubstituted Rubredoxin as a Minimal Enzyme Model for Hydrogenase. J. Phys. Chem. Lett. 2015, 6, 3731−3736. (c) Sommer, D. J.; Vaughn, M. D.; Clark, B. C.; Tomlin, J.; Roy, A.; Ghirlanda, G. Reengineering Cyt b for Hydrogen Production: A facile Route to Artificial Hydrogenases. Biochim. Biophys. Acta, Bioenerg. 2016, 1857, 598−603. (d) Sommer, D. J.; Vaughn, M. D.; Ghirlanda, G. Protein Secondary-shell Interactions Enhance the Photoinduced Hydrogen Production of Cobalt Protoporphyrin IX. Chem. Commun. 2014, 50, 15852−15855. (e) Kleingardner, J. G.; Kandemir, B.; Bren, K. L. Hydrogen Evolution from Neutral Water under Aerobic Conditions Catalyzed by Cobalt Microperoxidase-11. J. Am. Chem. Soc. 2014, 136, 4−7. (f) Saint-Martin, P.; Lespinat, P. A.; Fauque, G.; Berlier, Y.; Legall, J.; Moura, I.; Teixeira, M.; Xavier, A. V.; Moura, J. J. G. Hydrogen Production and Deuterium-proton Exchange Reactions Catalyzed by Desulfovibrio Nickel(II)-substituted Rubredoxins. Proc. Natl. Acad. Sci. U. S. A. 1988, 85, 9378−9380. (11) Weber, P. C.; Ohlendorf, D. H.; Wendoloski, J. J.; Salemme, F. R. Structural Origins of High-affinity Biotin Binding to Streptavidin. Science 1989, 243, 85−88. (12) Wilson, M. E.; Whitesides, G. M. Conversion of a Protein to a Homogeneous Asymmetric Hydrogenation Catalyst by Site-specific Modification with a Diphosphinerhodium(I) Moiety. J. Am. Chem. Soc. 1978, 100, 306−307. (13) (a) Liang, A. D.; Serrano-Plana, J.; Peterson, R. L.; Ward, T. R. Artificial Metalloenzymes Based on the Biotin−Streptavidin Technology: Enzymatic Cascades and Directed Evolution. Acc. Chem. Res. 2019, 52, 585. (b) Schwizer, F.; Okamoto, Y.; Heinisch, T.; Gu, Y.; Pellizzoni, M. M.; Lebrun, V.; Reuter, R.; Kohler, V.; Lewis, J. C.; Ward, T. R. Artificial Metalloenzymes: Reaction Scope and Optimization Strategies. Chem. Rev. 2018, 118, 142−231. (c) Heinisch, T.; Ward, T. R. Artificial Metalloenzymes Based on the BiotinStreptavidin Technology: Challenges and Opportunities. Acc. Chem. Res. 2016, 49, 1711−1721. (d) Chatterjee, A.; Mallin, H.; Klehr, J.; Vallapurackal, J.; Finke, A. D.; Vera, L.; Marsh, M.; Ward, T. R. An Enantioselective Artificial Suzukiase Based on the Biotin−streptavidin Technology. Chem. Sci. 2016, 7, 673−677. (e) Dundas, C. M.; Demonte, D.; Park, S. Streptavidin-biotin Technology: Improvements and Innovations in Chemical and Biological Applications. Appl. Microbiol. Biotechnol. 2013, 97, 9343−9353. (f) Ward, T. R. Artificial Metalloenzymes Based on the Biotin-Avidin Technology: Enantiose- lective Catalysis and Beyond. Acc. Chem. Res. 2011, 44, 47−57. (g) Letondor, C.; Ward, T. R. Artificial Metalloenzymes for Enantioselective Catalysis: Recent Advances. ChemBioChem 2006, 7, 1845−1852. (14) Keller, S. G.; Probst, B.; Heinisch, T.; Alberto, R.; Ward, T. R. Photo-driven Hydrogen Evolution by an Artificial Hydrogenase Utilizing the Biotin-Streptavidin Technology. Helv. Chim. Acta 2018, 101, e1800036. (15) (a) Call, A.; Codolà, Z.; Acuña-Parés, F.; Lloret-Fillol, J. Photoand Electrocatalytic H2 Production by New First-row Transitionmetal Complexes Based on an Aminopyridine Pentadentate Ligand. Chem. - Eur. J. 2014, 20, 6171−6183. (b) Call, A.; Franco, F.; Kandoth, N.; Fernández, S.; González-Béjar, M.; Pérez-Prieto, J.; Luis, J. M.; Lloret-Fillol, J. Understanding Light-driven H2 Evolution through the Electronic Tuning of Aminopyridine Cobalt Complexes. Chem. Sci. 2018, 9, 2609−2619. (16) Skander, M.; Humbert, N.; Collot, J.; Gradinaru, J.; Klein, G.; Loosli, A.; Sauser, J.; Zocchi, A.; Gilardoni, F.; Ward, T. R. Artificial Metalloenzymes: (Strept)avidin as Host for Enantioselective Hydrogenation by Achiral Biotinylated Rhodium-diphosphine Complexes. J. Am. Chem. Soc. 2004, 126, 14411−14418. (17) Wilson, M. E.; Nuzzo, R. G.; Whitesides, G. M. Bis(2diphenylphosphinoethyl)amine. A flexible Synthesis of Functionalized Chelating Diphosphines. J. Am. Chem. Soc. 1978, 100, 2269−2270. (18) (a) Planas, O.; Whiteoak, C. J.; Martin-Diaconescu, V.; Gamba, I.; Luis, J. M.; Parella, T.; Company, A.; Ribas, X. Isolation of Key Organometallic Aryl-Co(III) Intermediates in Cobalt-Catalyzed C(sp2)−H Functionalizations and New Insights into Alkyne Annulation Reaction Mechanisms. J. Am. Chem. Soc. 2016, 138, 14388−14397. (b) Planas, O.; Roldán-Gómez, S.; Martin-Diaconescu, V.; Luis, J. M.; Company, A.; Ribas, X. Mechanistic Insights into the SN2-type Reactivity of Aryl-Co(III) Masked-carbenes for C−C Bond Forming Transformations. Chem. Sci. 2018, 9, 5736−5746. (c) Sarangi, R.; Cho, J.; Nam, W.; Solomon, E. I. XAS and DFT Investigation of Mononuclear Cobalt(III) Peroxo Complexes: Electronic Control of the Geometric Structure in CoO2 versus NiO2 Systems. Inorg. Chem. 2011, 50, 614−620. (d) Fujita, E.; Furenlid, L. R.; Renner, M. W. Direct XANES Evidence for Charge Transfer in Co−CO2 Complexes. J. Am. Chem. Soc. 1997, 119, 4549−4550. (19) Nishida, Y.; Ida, K.; Kida, S. Investigation on Low-spin Cobalt(II) Complexes.X. [1] An ESR Study on Six-coordinate Complexes. Inorg. Chim. Acta 1980, 38, 113−116. (20) (a) The overpotential is defined as the difference between the applied potential and the thermodynamic potential for the reduction of protons at given conditions. The overpotential is defined as the difference between the applied potential and the thermodinamic potential for the reduction of protons at given conditions. (b) The thermodyinamic potential for the electrocatalytic water reduction to H2 at given pH (SCE, 1 atm of H2) is E(pH) = −0.241 V − 0.059 V × pH. The thermodynamic potential for the electrocatalytic water reduction to H2 at given pH (SCE, 1 atm of H2) is E(pH) = −0.241 V − 0.059 V × pH. (21) (a) McCrory, C. C.; Uyeda, C.; Peters, J. C. Electrocatalytic Hydrogen Evolution in Acidic Water with Molecular Cobalt Tetraazamacrocycles. J. Am. Chem. Soc. 2012, 134, 3164−3170. (b) Sun, Y.; Bigi, J. P.; Piro, N. A.; Tang, M. L.; Long, J. R.; Chang, C. J. Molecular Cobalt Pentapyridine Catalysts for Generating Hydrogen from Water. J. Am. Chem. Soc. 2011, 133, 9212−9215. (c) McNamara, W. R.; Han, Z.; Yin, C.-J. M.; Brennessel, W. W.; Holland, P. L.; Eisenberg, R. Cobalt-dithiolene Complexes for the Photocatalytic and Electrocatalytic Reduction of Protons in Aqueous Solutions. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 15594−15599. (d) Abe, T.; Kaneko, M. pH-dependent Electrocatalysis for Proton Reduction by Bis(2,2′:6′,2″-terpyridine) Cobalt(II) Complex Embedded in Nafion® Membrane. J. Mol. Catal. A: Chem. 2001, 169, 177−183. (e) Slater, J. W.; Marguet, S. C.; Monaco, H. A.; Shafaat, H. S. Going beyond Structure: Nickel-Substituted Rubredoxin as a Mechanistic Model for the [NiFe] Hydrogenases. J. Am. Chem. Soc. 2018, 140, 10250−10262. 5845 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846 Research Article ACS Catalysis (22) (a) Slattery, S. J.; Blaho, J. K.; Lehnes, J.; Goldsby, K. A. pHDependent Metal-based Redox Couples as Models for Protoncoupled Electron Transfer Reactions. Coord. Chem. Rev. 1998, 174, 391−416. (b) Save′ant, J.-M. Elements of Molecular and Biomolecular Electrochemistry: An Electrochemical Approach to Electron Transfer Chemistry; Wiley-Interscience: Hoboken, NJ, 2006. (c) Horvath, S.; Fernandez, L. E.; Appel, A. M.; Hammes-Schiffer, S. pH-dependent Reduction Potentials and Proton-coupled Electron Transfer Mechanisms in Hydrogen-producing Nickel Molecular Electrocatalysts. Inorg. Chem. 2013, 52, 3643−3652. (23) Anxolabéhère-Mallart, E.; Costentin, C.; Fournier, M.; Nowak, S.; Robert, M.; Saveant, J. M. Boron-capped Tris(glyoximato) Cobalt Clathrochelate as a Precursor for the Electrodeposition of Nanoparticles Catalyzing H2 Evolution in Water. J. Am. Chem. Soc. 2012, 134, 6104−6107. (24) Bigi, J. P.; Hanna, T. E.; Harman, W. H.; Chang, A.; Chang, C. J. Electrocatalytic Reduction of Protons to Hydrogen by a Watercompatible Cobalt Polypyridyl Platform. Chem. Commun. 2010, 46, 958−960. (25) Léger, C.; Elliott, S. J.; Hoke, K. R.; Jeuken, L. J. C.; Jones, A. K.; Armstrong, F. A. Enzyme Electrokinetics: using Protein Film Voltammetry to Investigate Redox Enzymes and their Mechanisms. Biochemistry 2003, 42, 8653−8662. (26) (a) Cline, E. D.; Adamson, S. E.; Bernhard, S. Homogeneous Catalytic System for Photoinduced Hydrogen Production Utilizing Iridium and Rhodium Complexes. Inorg. Chem. 2008, 47, 10378− 10388. (b) Curtin, P. N.; Tinker, L. L.; Burgess, C. M.; Cline, E. D.; Bernhard, S. Structure-activity Correlations among Iridium(III) Photosensitizers in a Robust Water-reducing System. Inorg. Chem. 2009, 48, 10498−10506. (c) Hanss, D.; Freys, J. C.; Bernardinelli, G.; Wenger, O. S. Cyclometalated Iridium(III) Complexes as Photosensitizers for Long-range Electron Transfer: Occurrence of a Coulomb Barrier. Eur. J. Inorg. Chem. 2009, 2009, 4850−4859. (d) Ladouceur, S.; Fortin, D.; Zysman-Colman, E. Role of Substitution on the Photophysical Properties of 5,5′-diaryl-2,2′bipyridine (bpy*) in [Ir(ppy)2(bpy*)]PF6 Complexes: a Combined Experimental and Theoretical Study. Inorg. Chem. 2010, 49, 5625− 5641. (27) Simond, M. R.; Ballerat-Busserolles, K.; Coulier, Y.; Rodier, L.; Coxam, J. Y. Dissociation Constants of Protonated Amines in Water at Temperatures from 293.15 to 343.15 K. J. Solution Chem. 2012, 41, 130−142. (28) Natali, M. Elucidating the Key Role of pH on Light-Driven Hydrogen Evolution by a Molecular Cobalt Catalyst. ACS Catal. 2017, 7, 1330−1339. (29) Han, Z.; McNamara, W. R.; Eum, M. S.; Holland, P. L.; Eisenberg, R. A nickel thiolate catalyst for the long-lived photocatalytic production of hydrogen in a noble-metal-free system. Angew. Chem., Int. Ed. 2012, 51, 1667−1670. (30) (a) Stayton, P. S.; Freitag, S.; Klumb, L. A.; Chilkoti, A.; Chu, V.; Penzotti, J. E.; To, R.; Hyre, D.; Le Trong, I.; Lybrand, T. P.; Stenkamp, R. E. Streptavidin−biotin Binding Energetics. Biomol. Eng. 1999, 16, 39−44. (b) DeChancie, J.; Houk, K. N. The Origins of Femtomolar Protein-ligand Binding: Hydrogen-bond Cooperativity and Desolvation Energetics in the Biotin-(Strept)avidin Binding Site. J. Am. Chem. Soc. 2007, 129, 5419−5429. (c) The structure used was taken from the protein data bank (PDB); 1MK5. (31) Schiller, T. L.; Keddie, D. J.; Blakey, I.; Fredericks, P. M. Surface-enhanced Raman Encoded Polymer Stabilized Gold Nanoparticles: Demonstration of Potential for Use in Bioassays. Eur. Polym. J. 2017, 87, 508−518. (32) Data collection with APEX II, version v2013, 4-1; Bruker AXS Inc.: Madison, WI, 2007. (33) Data reduction with Bruker SAINT, version V8.30c; Bruker AXS Inc.: Madison, WI, 2007. (34) SADABS, V2012/1; Bruker AXS Inc.: Madison, WI, 2001.Blessing, R. H. An empirical correction for absorption anisotropy. Acta Crystallogr. 1995, A51, 33-38. (35) TWINABS, Version 2012/1; Bruker AXS scaling for twinned crystals: Blessing, R. H. An empirical correction for absorption anisotropy. Acta Crystallogr., Sect. A: Found. Crystallogr. 1995, A51, 33−38. (36) SHELXT; V2014/4, 2014: Sheldrick, G. M. SHELXTintegrated space-group and crystal determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, A71, 3−8. (37) SHELXle: Hübschle, C. B.; Sheldrick, G. M.; Dittrich, B. ShelXle: a Qt graphical user interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281−1284. (38) SHELXL; SHELXL-2014/7, 2014: Sheldrick, G. M. SHELXTintegrated space-group and crystal determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, A71, 3−8. (39) Stoll, S.; Schweiger, A. EasySpin, a Comprehensive Software Package for Spectral Simulation and Analysis in EPR. J. Magn. Reson. 2006, 178, 42−55. (40) (a) Newville, M. IFEFFIT: Interactive XAFS Analysis and FEFF Fitting. J. Synchrotron Radiat. 2001, 8, 322−324. (b) Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data Analysis for X-ray Absorption Spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537−541. (41) Gaussian 09, Revision E.01; Gaussian, Inc.: Wallingford, CT, 2009. (42) (a) Becke, A. D. J. Chem. Phys. 1993, 98, 1372−1377. (b) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (c) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785. (43) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378−6396. (44) Grimme, S. Semiempirical GGA-type Density Functional Constructed with a Long-range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787−1799. (45) www.chemcraftprog.com. (46) Case, D. A.; Darden, T. A.; Cheatham, T. E.; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Crowley, M.; Walker, R. C.; Zhang, W.; Merz, K. M.; Wang, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Yang, L.; Tan, C.; Mongan, J.; Hornak, V.; Cui, G.; Mathews, D. H.; Seetin, M. G.; Sagui, C.; Babin, V.; Kollman, P. A. AMBER 16; University of California: San Francisco, CA, 2016. (47) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25, 1157−1174. (48) Hu, L.; Ryde, U. Comparison of Methods to Obtain ForceField Parameters for Metal Sites. J. Chem. Theory Comput. 2011, 7, 2452−2463. (49) Dolinsky, T. J.; Nielsen, J. E.; McCammon, J. A.; Baker, N. A. PDB2PQR: an Automated Pipeline for the Setup of PoissonBoltzmann Electrostatics Calculations. Nucleic Acids Res. 2004, 32, W665−667. (50) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926−935. (51) Wang, J.; Cieplak, P.; Kollman, P. A. How Well Does a Restrained ElectrostaticPotential (RESP) Model Perform in Calculating Conformational Energies of Organic and Biological Molecules? J. Comput. Chem. 2000, 21, 1049−1074. (52) Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An Nlog(N) Method for Ewald Sums in Large Systems. J. Chem. Phys. 1993, 98, 10089−10092. 5846 DOI: 10.1021/acscatal.8b04981 ACS Catal. 2019, 9, 5837−5846