Introduction

Hydrogen turnover (\({\mathrm{H}}_{2}\rightleftharpoons 2 {\mathrm{H}}^{+}+2 {\mathrm{e}}^{-}\)) is an important reaction in microorganisms that thrive under reducing, anoxic conditions and plays a role in certain aerobes [1]. Trace amounts of H2 provide electrons to power the anabolism of numerous archaea and bacteria in the soil, aqueous environments, or host tissue [2,3,4]. Under fermentative or ‘micro aerobic’ conditions, proton reduction and H2 release has been shown to be a key in the redox regulation of autotrophs like photosynthetic bacteria and algae [5,6,7]. The enzymes responsible for proton reduction and H2 oxidation are referred to as ‘hydrogenases’ [8,9,10].

In archaea, Hmd [Fe]-hydrogenase is involved in the conversion of CO2 to CH4 (methanogenesis), catalyzing H2 splitting and hydride transfer at a cofactor with a monometallic iron center [11,12,13]. Unlike Hmd [Fe]-hydrogenase, [NiFe]- and [FeFe]-hydrogenase are iron-sulfur enzymes and carry a bimetallic active site cofactor. [NiFe]-hydrogenases have been found in archaea and bacteria, and are typically involved in heterotrophic H2 uptake [13,14,15]. The soluble ‘standard’ [NiFe]-hydrogenases are distinguished from membrane-bound, multimeric, as well as O2-tolerant and bidirectional [NiFe]-hydrogenases [15,16,17]. Although highly diverse, all [NiFe]-hydrogenases share the same active site cofactor. It is comprised of a nickel and an iron ion, covalently attached to the enzyme by either four cysteine residues, or three cysteine and one selenocysteine. The iron ion is coordinated by two cyanide (CN) and one carbon monoxide ligand (CO) ligand, the latter being a common feature of hydrogenases [18].

[FeFe]-hydrogenases have been found in bacteria, lower eukaryotes, and green algae, representing the phylogenetically most recent class [8,9,10]. [FeFe]-hydrogenases are involved in H2 uptake as well as redox regulation (H2 release upon proton reduction), but in complex with other metalloproteins, additional roles in CO2 reduction or electron bifurcation are suggested [19]. Figure 1 depicts cyclic voltammograms of two different [FeFe]-hydrogenases that demonstrate bidirectional catalysis [20]. Typically, H2 oxidation is observed up until 0 V vs SHE, while many [FeFe]-hydrogenase are inhibited at more positive potentials, i.e., due to an interaction with halide anions like chloride [21]. At potentials more negative than -0.3 V, a proton reduction current is observed resulting in H2 evolution while the current at higher potentials indicates H2 oxidation. Near the equilibrium or midpoint potential (Em) the ‘clean cut’ through the zero-current X-axis indicates minimum electric overpotential, a key property of hydrogenase-catalyzed hydrogen turnover [22,23,24]. The catalytic mechanism of [FeFe]-hydrogenase is discussed controversially [25,26,27] as we will see in Sect. “The Spectroscopic Characterization of [FeFe]-hydrogenase”.

Fig. 1
figure 1

Reproduced with permission from the ACS

Cyclic voltammograms of [FeFe]-hydrogenases A CrHydA1 and B CaHydA indicate bidirectional catalysis. Recorded between −0.5 V and +0.3 V vs. SHE at pH 6 and in the presence of 1 atm H2, (i) positive currents and (ii) negative currents results from H2 oxidation and proton reduction (i.e., H2 evolution), respectively. The midpoint potentials (Em) are highlighted. The decrease of current above 0 V (particularly visible for CrHydA1 in panel A) is due to ‘anaerobic inactivation’. [20]

While all [FeFe]-hydrogenases carry the same catalytic cofactor, a classification into four groups A–D has been suggested among which ‘Group A’ represents the well-studied ‘standard’ [FeFe]-hydrogenases [28]. Figure 2 shows the crystal structure of standard [FeFe]-hydrogenase CpI including the four ferredoxin-type iron-sulfur clusters (‘F-clusters’) that connect the hydrogen-forming active site cofactor (‘H-cluster’) with the surface. The H-cluster comprises a [4Fe-4S] cluster (‘[4Fe]H’), connected to a homobimetalic iron site via a bridging cysteine ligand. The metal ions of this diiron site (Fep, Fed) are modified with a bridging carbonyl ligand (µCO) and two terminal CN and CO ligands, while the apical-binding site of Fed (X) may be vacant or occupied with different ligands. A unique S2(CH2)NH ligand (azadithiolate, ADT) connects the two iron ions. The biosynthesis of the H-cluster is well understood [29].

Fig. 2
figure 2

Crystal structure of the [FeFe]-hydrogenase CpI (PDB ID 4XDC). The accessory, ferredoxin-type iron-sulfur clusters (F-clusters) are highlighted in blue, the active site cofactor (H-cluster) is highlighted in red. On the right side, the H-cluster is shown as a chemical model. Position ‘X’ may be vacant or occupied with ligands like CO, CN, H, S, or O2. [4Fe]H, [4Fe-4S] cluster; µCys, bridging cysteine; CO, carbon monoxide; CN, cyanide; ADT, azadithiolate; Fep/Fed proximal/distal iron ion (relative to [4Fe]H)

The active site niche of the H-cluster provides a specialized environment beyond the first coordination sphere [30]. The [4Fe]H cluster is in electron tunneling distance to the proximal [4Fe-4S] cluster, facilitating electron exchange between the diiron site and soluble redox partners like ferredoxin [31] The charge of the F-clusters affects the electron density distribution across the H-cluster and has been suggested to influence catalysis [32], together with protein structural fluctuations [33] and/or protonation changes near the [4Fe]H. [34] Conserved amino acid residues facilitate proton transfer between bulk solvent and H-cluster [35] Here, the amine head group of the ADT ligand receives a proton from a nearby cysteine (Fig. 2). The ADT ligand does not only function as a proton relay but has been discussed to stabilize various H-cluster states by ‘intramolecular’ hydrogen bonding [36] The CO and CN ligands have be modeled to face hydrophobic and hydrophilic pockets of the active site niche, and for Fep-CN, hydrogen-bonding interactions with the protein fold have been confirmed [37] While the H-cluster is the same in all [FeFe]-hydrogenases, significant diversity is observed among the primary structure of Group A–D enzymes [28].

In this review, we will establish a historical and topical overview on the literature of [FeFe]-hydrogenase since the late 1990s. We focus on structural biology and biophysical aspects of hydrogenase research. The timeline in Scheme 1 suggests four phases: between 1998 and 2002 (phase I), the first crystal structures were solved by X-ray diffraction protein crystallography and three H-cluster states (Hox, Hox-CO, Hred) were characterized by electron paramagnetic resonance spectroscopy and Fourier-transform infrared spectroscopy. In a second phase (ca. 2004–2010), different bacterial expression hosts were introduced that allowed producing and modifying [FeFe]-hydrogenases in preparative amounts. Moreover, protein film electrochemistry became a go-to technique following hydrogenase turnover, CO inhibition, and O2 deactivation. The understanding of in vivo H-cluster synthesis and the development of ‘artificial maturation’ characterize the third phase (ca. 2010–2017). The characterization of cofactor states with a reduced [4Fe]H cluster (Hsred, Hred´) falls in this phase, too. The current and ongoing phase IV (since 2017) is dominated by the discussion about proton transfer, electron transfer, and the catalytic mechanism of [FeFe]-hydrogenase. Additionally, we saw a ‘revival’ of the inactive H-cluster state (Hinact) that protects [FeFe]-hydrogenase from O2 deactivation.

Scheme 1.
scheme 1

Timeline of important events and key findings in [FeFe]-hydrogenase research. Phases I–IV are marked black, blue, red, and white, respectively. Hox, Hox-CO, Hred, Hsred, Hred´, and Hinact are H-cluster states

This review is organized in two main sections, introducing structural aspects of [FeFe]-hydrogenase in Sect. “The Structural Characterization of [FeFe]-hydrogenase” and spectroscopic properties in Sect. “The Spectroscopic Characterization of [FeFe]-hydrogenase”. Following each section or sub-section, ‘further reading’ suggestions will serve as starting point for a topical deep dive into the literature on [FeFe]-hydrogenase.

figure b

The structural characterization of [FeFe]-hydrogenase

We will start this chapter with a brief summary of key findings between 1931 and 1997; for a more detailed review on the history of [FeFe]-hydrogenase until 1990, see Ref. [38]. Hydrogenases were first described by Stephenson and Stickland in 1931 [39] and demonstrated to catalyze H2 oxidation and H/D exchange with the solvent by Yudkin et al. in 1934 [40]. Throughout the 1940’s and 1950’s, hydrogenases were reported to interact with O2, CO, CN, and visible light [41] and although the prosthetic groups were unknown, non-heme iron was identified to play a key role [42]. In 1963, Sadana and Rittberg noted a peculiar dependence of hydrogenase activity on sulfur, which led them to postulate ferrous-hydroxyl complexes [43]. Improving the purity of enzyme preparations in the early 1970’s, Nakos and Mortenson eventually reported an iron-sulfur ratio of 1:1 in the hydrogenase from C. pasteurianum [44], whose electron paramagnetic resonance (EPR) spectra resembled the [4Fe-4S] clusters of ferredoxin [45] While this settled the identity of hydrogenases as iron-sulfur enzymes, EPR also suggested the presence of a unique metal site, tentatively assigned to the catalytic active site in an important review article by Mortenson et al. [46]. In the 1980’s, certain hydrogenase preparations showed evidence of [3Fe-4S] clusters [47], selenocysteine [48], and a dependence on nickel [49], which inspired the distinction between “Fe-hydrogenase” and “Ni-hydrogenase” (Hmd [Fe]-hydrogenase was identified by Thauer et al. no earlier than 1992 [50]).

The first structure of a [NiFe]-hydrogenase was published by Fontecilla-Camps in 1995, showing the active site iron ion with three non-protein ligands, tentatively assigned to water [51] Based on the unique infrared signature of [NiFe]-hydrogenase, in 1997, Albracht et al. suggested these ligands to be CO and CN [52], and comparing the infrared spectra of different hydrogenases, a somewhat similar architecture was predicted for [FeFe]-hydrogenase [53], the later which was yet to be crystalized.

In December 1998 and January 1999, the first cryogenic X-ray diffraction (XRD) crystal structures of [FeFe]-hydrogenase were published independently by two groups. Peters et al. solved the structure of the oxidized, monomeric [FeFe]-hydrogenase from C. pasteurianum (CpI) with a resolution of 1.8 Å [54] The authors identified the binding motives of the accessory iron-sulfur F-clusters that facilitate electron transfer between bulk solvent and active site cofactor. Furthermore, Peters et al. were the first to describe atomistic details of the H-cluster, including a discussion of amino acid residues that line the active site niche and may serve as hydrogen-bonding partner to the CN ligands, shield the H-cluster from unwanted solvent access, or enable catalytic proton transfer. It took 20 more years to validate the functionality of the proton transfer pathway in [FeFe]-hydrogenase CpI [55,56,57]. Here, protein crystallography helped uncovering the role of water molecules that substitute for certain amino acid sidechains when replaced by site-directed mutagenesis. In their pioneering study, Peters et al. found the H-cluster to bind a putative water molecule in the apical position of the distal iron ion, Fed (Fig. 3A). Fontecilla-Camps et al. published the crystal structure of the oxidized, dimeric [FeFe]-hydrogenase from D. desulfuricans (DdHydAB) with a resolution of 1.6 Å only one month later [58] Their model of the H-cluster shows no indications of a water molecule in the apical position of Fed. Furthermore, the structure comprises a fully resolved dithiolate ligand, which was assigned to S2(CH2)3 back then (propanedithiolate, PDT, see Fig. 3B). Fontecilla-Camps et al. observed a hydrophobic gas channel and discussed an alternative proton transfer pathway, notably independent of the dithiolate ligand and including a conserved lysine residue in hydrogen-bonding distance to the distal CN ligand. Although certain conclusions from this study did not stand the test of time, the closing statement reads virtually prophetic:

“Perhaps the most striking feature of the novel diiron active-site center of the [FeFe]- hydrogenase is its almost complete independence from protein coordination. It is possible that such a center may have been imported from the inorganic world as an already functional unit.”

Fig. 3
figure 3

Early XRD crystal structures of the H-cluster of [FeFe]-hydrogenases CpI and DdHydAB in the oxidized (A, B), CO-inhibited (C), and reduced state (D). The identity of the dithiolate ligand, PDT, was latter corrected to ADT. From A–C, the PDB IDs are 1FEH, 1HFE, and 1C4A (unfortunately, there is no PDB entry for reduced DdHydAB). Reproduced with permission from the AAAS [54], Cell Press [58], and the ACS [59, 60]

In the following years, the group of John Peters published on the mechanism of CO inhibition of [FeFe]-hydrogenase. Studying the photochemistry of CO-inhibited CpI in cristallo [59] established the belief that external CO binds at the apical-binding site of Fed (Fig. 3C). Meanwhile, the group of Juan Fontecilla-Camps investigated the geometry of the H-cluster of DdHydAB crystals grown in the presence of chemical reductant and H2 [60]. Their XRD model of the reduced H-cluster comprised an open apical coordination site and a ‘semi-bridging’ CO ligand at Fed (Fig. 3D). The latter is in variance to the oxidized H-cluster that clearly showed a Fe–Fe bridging CO ligand, µCO [54, 58]. The crystallographic findings were in apparent agreement with room temperature infrared spectro-electrochemistry titrations on [FeFe]-hydrogenase solution that showed the loss of µCO upon conversion of the oxidized state Hox into one of the reduced states, Hred [60]. Accordingly, Fontecilla-Camps et al. assigned the ‘semi-bridging’ H-cluster geometry to Hred, a decision that would dominate the interpretation of [FeFe]-hydrogenase catalysis for nearly 20 years.

The rise of the photosynthetic hydrogenase and artificial maturation

Across the first decade of the twenty-first century, the enzyme CrHydA1 from green algae Chlamydomonas reinhardtii became the new standard [FeFe]-hydrogenase. Happe and Stripp chose the catchphrase “photosynthetic hydrogenases”, because it accepts electrons from ferredoxin at the reducing end of the photosynthetic electron transport chain [61] CrHydA1 is attractive for many spectroscopic approaches (Sect. “The Spectroscopic Characterization of [FeFe]-hydrogenase”) since it lacks the accessory iron-sulfur clusters of CpI and DdHydAB. Following protocols for homologous and heterologous production in Clostridium acetobutylicum or Shewanella oneidenis [62,63,64], eventually transgenic E. coli cells (carrying the genes of [FeFe]-hydrogenase maturases HypEFG) were established as most efficient expression host for CrHydA1 [65, 66]. In 2010, Peters et al. exploited the E. coli protocol to produce and crystallize apo-CrHydA1, i.e., enzyme without the diiron site of the H-cluster [67]. Despite various efforts, the structure of functional holo-CrHydA1 is yet to obtain; it has been argued that the 45 amino acid insertion forms a flexible loop that might be important for the interaction with in vivo redox partner ferredoxin [68]—but ultimately impedes crystallization.

Understanding the enzymatic details of H-cluster biosynthesis was basis for the production of [FeFe]-hydrogenase in E. coli, an organism that naturally relies on [NiFe]-hydrogenase. After the successful modification of E. coli with [FeFe]-hydrogenase maturases HydEFG and the establishment of efficient expression, synthesis, and purification protocols [65,66,67], the next step was to establish the artificial maturation of [FeFe]-hydrogenase [69,70,71]. Fontecave et al. reported that HydF can be loaded with synthetic precursors of the diiron site [69]. When HydF and apo-CrHydA1 was mixed, functional [FeFe]-hydrogenase was produced, which allowed generating ‘cofactor variants’, e.g., using unnatural diiron site precursors (Fig. 4). Less than two months later, Happe et al. demonstrated the direct activation of apo-CrHydA1 and apo-CpI, avoiding HydF altogether [70]. This further simplified the production of native and artificial [FeFe]-hydrogenase. The artificial maturation of DdHydAB was reported as well [71]; however, the majority of cofactor variants has been produced in CrHydA1 and CpI, including isotopomers [72,73,74], variations of the ligands [75], the iron ions [76], and the sulfur atoms [77,78,79]. Numerous CpI cofactor variants have been crystallized [78,79,80].

Fig. 4
figure 4

Synthetic diiron complexes can be used to activate the [FeFe]-hydrogenase maturation enzyme HydF (A) or the [FeFe]-hydrogenase apo-protein directly (B). In the initial studies, a variation of the dithiolate headgroup (X) was demonstrated. Reproduced from Ref. [69] with permission from Nature Springer

Deactivation and inhibition in the presence of O2

Most [FeFe]-hydrogenases are irreversibly deactivated by trace amounts of O2. Although DdHydAB can be isolated under aerobic conditions [58], once activated the H-cluster falls apart when reacted with O2. This is in marked difference to [NiFe]-hydrogenase that are reversibly inhibited by O2 or oxidize O2 to H2O maintaining hydrogen turnover under aerobic conditions [81, 82]. The deactivation reaction of [FeFe]-hydrogenase is less well understood. Crystal structures of CpI obtained at different time points after exposure to O2 point toward oxidative damage at the H-cluster, possibly due a formation of reactive oxygen species (ROS), as suggested in earlier work [83,84,85]. Catalytic conversion of O2 into ROS is supported by the observation that CO inhibits both the reaction with H2 and O2 [20] hinting at a similar binding site, i.e., the open coordination site of Fed. ‘Super-oxidized’ intermediates of the reaction with O2 are typically short-lived, but proton transfer variants like C169A of CrHydA1 allowed stabilizing respective H-cluster species [86]. This hints at H2O2 as a key ROS in the deactivation process. The structure of the H-cluster in the O2-inhibited state Hinact of DdHydAB and CbA5H was debatted until crystallographic analyses revealed that a cysteine residue binds to Fed of the H-cluster, protecting the cofactor from O2 deactivation like a ‘safety cap’ [87]. Interestingly, sulfide has a similar effect and binds to Fed under oxidizing conditions both aerobic and anaerobic (Fig. 5), which allows reversible conversion of the Hox and Hinact states [88,89,90].

Fig. 5
figure 5

XRD crystal structure of DdHydAB (PDB ID 6SG2) and CbA5H (PDB ID 6TTL) in the Hinact state. A In DdHydAB, the apical-binding site of Fed indicates electron density subsequently identified as a sulfur atom by XAS spectroscopy. Reproduced from Ref. [91] with permission from Wiley–VCH. B In CbA5H, adjacent cysteine C367 binds to the apical binding site of Fed in the presence of O2. Reproduced from Ref. [87] with permission from Springer Nature

Future directions

Despite the pioneering work on CO-inhibited and H2-reduced [FeFe]-hydrogenase no other redox state has been explored by protein crystallography in the last 20 years. Certain updates are necessary. For example, to conclude the discussion regarding the identity of the diiron site-reduced redox states Hred and Hsred (Sect. “The Spectroscopic Characterization of [FeFe]-hydrogenase”), it will be required to obtain atomic resolution XRD data on spectroscopically verified protein crystals, i.e., using in cristallo spectroscopy [92, 93]. The effect of temperature will have to be considered as well [94]; to this end, King and Peters et al. published the first XRD structure of oxidized CpI using the X-ray free electron laser (XFEL) at Linac Coherent Light Source, which allows recording radiation damage-free diffraction patterns [33]. The authors measured CpI at cryogenic temperature [33] but in principle, protein structure determination at ambient-temperature is possible using XFEL radiation. Low temperature cannot be avoided in cryogenic electron microscopy (Cryo-EM) experiments and the technique may not be suitable for atomistic details and the comparatively small standard [FeFe]-hydrogenases DdHydAB, CpI, and CrHydA1. However, [FeFe]-hydrogenases are diverse, and there are multimeric enzymes that are poorly understood [19]. Just recently, the Cryo-EM structures of HydABC from Thermotoga maritimaAcetobacterium woodii, and Thermoanaerobacter kivui have been employed to understand the mechanism of electron bifurcation [95, 96]. To conclude this section, the potential of ambient-temperature neutron scattering in the determination of hydrogen atoms is noted. This technique may help identifying site-selective protonation or hydride formation events. Initial work on [NiFe]-hydrogenase has been published [97] but no [FeFe]-hydrogenase structure is reported yet.

figure c

The spectroscopic characterization of [FeFe]-hydrogenase

In the second section, we saw how the structural characterization of CpI and DdHydAB by cryogenic XRD lay the foundation for a principle understanding of [FeFe]-hydrogenase function. To rationalize the details of electron transfer, proton transfer, and the catalytic mechanism, additional methods must be taken into consideration. Such methods, electron paramagnetic resonance spectroscopy (Sect. “Electron Paramagnetic Resonance”) and vibrational spectroscopies (Sect. “Vibrational Spectroscopy”), are complimentary techniques which allow for studying stable intermediates along the catalytic cycle of [FeFe]-hydrogenase.

Electron paramagnetic resonance

Electron paramagnetic resonance (EPR) is a spectroscopic technique to investigate metal centers or organic molecules with unpaired electrons. An advantage of EPR spectroscopy is that the experiment specifically reports on the unpaired electron at the active site. In contrast, nuclear magnetic resonance (NMR) spectroscopy probes all the nuclei in a protein and may have a ‘blind sphere’ around a paramagnetic center (i.e., a metalloprotein active site) whereas the NMR signal is unrecognizably broadened. However, for EPR spectroscopy, there are no protein size limitations and, through the interactions of the electron with the environment (first and second coordination sphere, coupled nuclei, etc.), a unique view of the electronic structure of paramagnetic states along an enzymatic catalytic cycle can be obtained [98]. Both conventional continuous wave (CW) and ‘advanced’ pulse EPR spectroscopy are central in metalloprotein research [99] and have been extensively used in the analysis of [FeFe]- and [NiFe]-hydrogenase [100].

Insights from continuous wave and advanced EPR

The active site cofactor and the accessory iron-sulfur clusters of [FeFe]-hydrogenase give rise to characteristic EPR signals that can be stabilized as intermediates along the catalytic cycle. While the F-clusters are paramagnetic in the reduced state, the H-cluster show EPR activity in the oxidized state [101,102,103], which allows for a clear separation of signals between the active site and electron transfer path.

Investigating the F-clusters to understand electron transfer and magnetic coupling, Peters et al. recorded the EPR spectra of apo-CpI and cofactor variant CpIPDT under increasingly reducing conditions [104]. The authors determined the individual second derivative spectra and midpoint redox potentials (Em) of all metal centers (Fig. 6A). The low potential of the histidine-ligated, distal cluster FS4C (Em ≤ −450 mV) will limit the physiological H2 evolution activity of CpI to higher concentrations of ferredoxin; however, the small differences between FS4B, FS4A, and the H-cluster (ΔEm = 10 mV) suggest unimpeded electron transfer in both directions and suggest a large degree of magnetic coupling [104]. Such effects were argued to influence proton transfer, protonation events at the H-cluster, and catalytic turnover in general [105].

Fig. 6
figure 6

A Extracted second derivative EPR spectra of the H-cluster state Hox in cofactor variant CpIPDT and the F-clusters FS4A–C ([4Fe-4S]) and FS2 ([2Fe-2S]). B Fdx/CpI docking model annotated with the midpoint redox potentials of all hydrogenase metal centers. Reproduced from Ref. [104] with permission from the ACS

Once the EPR spectra of the F-clusters are known, the H-cluster can be analyzed specifically. Early work on aerobically isolated DdHydAB showed that the diamagnetic, O2-inhibited state Hinact is activated under reducing conditions. At about −200 mV vs. SHE one electron jumps at the [4Fe]H cluster, forming the Htrans state that converges irreversibly into the Hox state at about −300 mV vs. SHE. This indicates a reduction of the diiron site upon intramolecular electron transfer [106, 107]. Both Htrans and Hox are paramagnetic and give rise to rhombic EPR spectra (Fig. 7). While Hinact and Htrans were initially not observed in other [FeFe]-hydrogenases, several authors recently demonstrated the formation of Hinact over Hox in the presence of sulfide [89, 90]. When Hox is reacted with CO instead, [FeFe]-hydrogenases adopt the Hox-CO state (Fig. 7) that is characterized by an axial EPR spectrum [103] resulting from strong spin exchange coupling between [4Fe]H cluster and diiron site [108]. This is in agreement with the 57Fe hyperfine couplings observed by Mössbauer spectroscopy [109,110,111]. Today, the redox state of the metal ions in the diiron site is commonly denoted as [Fep(II)-Fed(I)] for Hox (with a more equal distribution of charges in Hox-CO) and [Fep(II)-Fed(II)] for Htrans and a recently discovered, cyanide-inhibited H-cluster state [112, 113]. The paramagnetism of the H-cluster in the latter states results from the reduced [4Fe]H cluster, which also explains the EPR spectrum of the so-called hydride state, Hhyd [114,115,116]. Formally, a ‘super-reduced’ H-cluster species, most of the electron density is located at the apical hydride ligand at Fed, stabilized by a [Fep(II)-Fed(II)]-type diiron site and a reduced [4Fe]H cluster. Figure 7 shows that the rhombic EPR spectra of Hhyd and Htrans are strikingly similar but clearly distinct from Hox. Reduction of [4Fe]H and the diiron site ([Fep(I)-Fed(I)]) produces another ‘super-reduced’ state, Hsred, that shows a broad axial EPR signal. Horch and Zebger et al. suggested that fast spin relaxation across the H-cluster in Hsred can explain these spectral properties [117].

Fig. 7
figure 7

Simulation of EPR spectra at 9.5 GHz (X-band) for key paramagnetic H-cluster states. Characteristic g-tensors are annotated as observed in DdHydAB (Htrans) and CrHydA1 (everything else)

In the previous section, we highlighted the use of CW EPR spectroscopy to characterize the stable intermediates of the active site H-cluster and stepwise reduction techniques to follow the electron pathway of the F-clusters. Continuous wave EPR, gives a snapshot of the electron and the total interaction with the coordination environment. This includes broadening effects, strong nuclei coordination, and spin–orbit coupling of the electron to direct coordination (metal ion; rhombicity). However, finer details of the coordination are unable to be determined due to the linewidths of the inhomogenously broadened spectra that is typical of the [FeFe]-hydrogenase active site [98]. Instead, we use pulse EPR techniques to ‘look within’ the broadened spectrum to characterize direct coordination between the free electron and surrounding nuclei which possess a nuclear magnetic moment (57Fe, 14/15N, 13C, 17O, 33S, 31P, 1/2H, etc.).

In these cases double resonance techniques, such as electron nuclear double resonance (ENDOR) where the nuclei are directly excited by an additional incident radio frequency (< 100 MHz) pulse, or multi-pulse techniques using only microwave pulses, such as electron spin echo envelope modulation (ESEEM) and hyperfine sub-level correlation (HYSCORE) spectroscopies are used, where the EPR signal is modulated by the forbidden transition probabilities of the coupled nuclei [118]. These techniques can be performed at multiple frequencies such as, 9.5 and 35 GHz, but also high frequency (> 94 GHz), and provide the framework to study the coordination chemistry of the active site in detail [118]. By increasing the frequency of the EPR experiment, the anisotropy of the g-tensor is increased further resolving the gxx, gyy, and gzz components. This allows for orientation selection to be performed on a frozen solution sample and the g- and hyperfine-tensors can be estimated.

Although the wild-type [2Fe]H cluster contains 14N ligands in the ADT ligand and the two CN ligands, it would be beneficial to be able to choose the isotopes in the active site to further gain information on the electron interaction within the diiron site. This was achieved by the ability to express [FeFe]-hydrogenase in its apo form, synthesize a chemical mimic of the [2Fe]H cluster, and for the chemical mimic to be inserted in apo-HydA1 producing a functioning enzyme [71, 72]. Doing so allowed the characterization of enriched isotopes (2H, 13C, 15N, 57Fe) of the [2Fe]H cluster to be studied by EPR pulse spectroscopy to learn about the electronic structure of the active site. Additionally, non-natural [2Fe]H clusters could be synthesized to trap intermediates that are typically transient or elusive [119] and give insights into proton transfer pathways [36, 80].

In this section, we will highlight the use of pulse EPR to characterize the active site in the Hox and Hox-CO state. Then, we focus on recent work on characterizing the maturation process of the [2Fe]H cluster using isotope labeled amino acids in HydG, HydE, and HydF. This section ends with a brief overview on where information gaps exist and how further techniques can be exploited on to address these shortcomings.

Characterization of the oxidized H-cluster and Maturation.

The data in Tables 1, 2 represent the current knowledge on the electronic properties of the [FeFe]-hydrogenase in the Hox and Hox-CO states. However, this table is not exhaustive and further information on both the magnitude and orientation of the g-, hyperfine-, and quadrupole-tensors are still needed. These tables are further complicated by the g-value and hyperfine shifts seen between different hydrogenase species, but still serves as a collection of the many hyperfine experiments that have been performed on [FeFe]-hydrogenase [73, 120,121,122]. Since EPR is a direct measurement of the electronic structure of the state of the enzyme, the g- and hyperfine-tensor shifts indicate a perturbation of the electronic structure by the 1st and 2nd coordination sphere. It is this local environment that plays an important role in fine-tuning the overall reactivity of the enzyme [30].

Table 1 EPR parameters determined by frozen solution [FeFe]-hydrogenase in the Hox state through a series of advanced EPR experiments.
Table 2 EPR parameters determined by frozen solution [FeFe]-hydrogenase in the Hox-CO stable catalytic intermediate through a series of advanced EPR experiments.

For the [2Fe]H cluster specifically, differences in the g-values and hyperfine-values using 57Fe between species are hypothesized to be variable coupling to the [4Fe]H cluster. Full DFT models using all six iron ions (as opposed to truncated models of the [2Fe]H cluster in ref. [123]) have put the effective electron spin at on the Fed while DFT calculations of Hox-CO place the majority of the electron spin at the Fep. From experimental data, shown in Table 2 and noted in Ref. [124], although most of the electron spin is located on Fep, hyperfine data of the other ligands suggests substantial distribution over Fed and the CO but not over the CN ligands. This paints an interesting picture of the interaction of the exogenous ligand COexo, the [2Fe]H cluster, and the drastic electronic re-configuration when CO is released forming the active Hox state.

Recent experiments using time-resolved CW EPR by Berggren et al. allowed monitoring the adoption of the [2Fe]H cluster into an apo-HydA protein [128]. These data suggest an incorporation of the diiron site to the apo-HydA1 via electron transfer from [4Fe]H2+ to [4Fe]H+ and changes in the redox state of the Fed ion forming a CO-inhibited [4Fe]H+-[Fep(II)Fed(I)] configuration. This species is the Hred´-CO intermediate [129] that rapidly oxidizes into an [4Fe]H2+–[Fep(II)Fed(I)] configuration, characteristic for the Hox-CO state. This rate-limiting step increases the energy required to remove the [2Fe]H cluster. The final step is the slow release of COexo to form the catalytically active Hox state.

Fundamental work by Britt et al. has put together the biosynthesis model of the [2Fe]H cluster (Fig. 8), using an array of biophysical techniques [130,131,132,133]. Included in these techniques was the isotope labeling of key amino acids in the biosynthetic pathway. Here, the radical S-Adenosyl-L-methionine (rSAM) enzyme, HydG, was found to cleave tyrosine to generate the CO and CN ligands as the precursor to the [2Fe]H cluster. Extensive CW and pulse EPR was performed by Britt et al. for characterizing this rSAM mechanism [134]. By removing the dangler iron attached to the rSAM [4Fe-4S] active site, it was possible to pause HydG reactivity and study the tyrosine substrate by separately labeling two hydrogen pairs with 2H, and a carbon pair with 13C. By comparing the EPR experimental data with DFT, the coordination chemistry and mechanism of HydG was proposed (Fig. 8). Prior to the understanding of HydG, the rSAM enzyme, HydE, could be studied by providing a synthetic substrate to HydE and watching the maturation with EPR [135] and the individual steps were found, again, with isotope labeling and following maturation through double resonance techniques [130].

Fig. 8
figure 8

Proposed sequential model for the roles of HydG, HydE, and HydF in the biosynthesis of the H-cluster. This model has been developed by David Britt and others using an array of biophysical techniques (EPR, FTIR, XRD, EXAFS). This figure is obtained from Ref. [131] under the creative commons license from RSC

In the studies of HydE, freeze-quenched methods with a time resolution of 10 s were employed to prepare EPR samples using a synthesized HydG product (complex-B) named ‘syn-B’. Letting the intermediate complex evolve over approximately 10 min reveals an EPR active complex that is shown to retain the two CO and one CN ligands from HydG biosynthesis. Finally, this EPR active intermediate forms an antiferromagnetically coupled (S = 0, EPR silent) [Fe(I)2S2(CO)4(CN)2] dimer intermediate. Future work is required to provide the biosynthesis of the final [2Fe]H cluster with the required ADT ligand.

Through these finding, this work has also led the way in biosynthesis models of [FeFe]-hydrogenase with only HydF maturase [136]. These findings are important for future work in selective isotope labeling of the [2Fe]H cluster and teasing out further mechanistic studies of HydF and HydE. Although the docking of HydE to HydF is known, the product transfer and final maturation is still unknown and will take a combination of FTIR and EPR studies to find the last steps in the biosynthesis.

Ultimately, as this work matures the understanding of the maturation will further improve synthesis protocols of [2Fe]H cluster for bio-inspired mimics. In a recent review, Broderick et al. summarize the biosynthetic pathway of the [FeFe]-hydrogenase, including the semisynthetic approaches by Britt et al. using advanced EPR methods [137].

Future directions

[FeFe]-hydrogenases are under further investigation using spectroscopic techniques to understand the mechanisms of O2 tolerance. Kasanmascheff et al. reported an interesting radical (R•ox) in the [FeFe]-hydrogenase CbA5H, which is known to be O2-tolerant [87, 138]. When kept under aerobic conditions, the R•ox signal significantly increases while CbA5H remains in the Hinact state. Further studies using 57Fe labeling produced hyperfine couplings in the 25–35 MHz range, suggesting a similar configuration to the Hox-CO state. The origin of R•ox needs further experimental and site-directed mutagenesis studies. For instance, in Hinact of DdHydAB, a sulfur has been shown to be coordinated to Fed [91] and a radical formed from a nearby amino acid could provide protection against reactive oxygen species [139].

Structural studies of the [FeFe]-hydrogenase maturation process are few, but Happe et al. have shown that incorporation of the [2Fe]H cluster is guided by arginine R449 and lysine L358 of CpI [140]. The molecular backbone movements could be studied by dipolar-dipolar EPR spectroscopy, e.g., between the H-cluster and a nitroxide label or between multiple nitroxide labels that have been incorporated by site-directed mutagenesis [141]. The structure of CrHydA1 apo-protein is controversial as the ‘open’ fold identified by Peters et al. significantly differs from holo-enzyme [67], while Happe et al. reported a structure that is barely any different from functional CrHydA1 [80]. Structural in situ data on the process of maturation will help solving the remaining questions.

To date, quantum chemical calculations have difficulties in predicting the principal values of the g-tensor and are only used to qualitatively assign spin densities when simulating interactions with surrounding nuclei [123, 125]. As DFT and atomic orbital quantum chemical computing techniques become more accurate for open shell systems, such as metalloenzymes, experimental data of full g- and hyperfine-tensors will be required to benchmark these simulations [142, 143]. However, the techniques described above rely on orientation selection of the frozen solution. To this end, one can either increase the operating frequency of the EPR experiment (separating the gxx, gyy, and gzz components) or perform single-crystal EPR experiments. Single-crystal EPR experiments are the ultimate method in determining the full magnetic interactions (magnitude and orientation), reflecting the electronic structure of the unpaired electron at an active site of an enzyme, and relating the magnetic principal axes to the proposed protein structure. Unlike [NiFe]-hydrogenase, which produced crystal dimensions on the order of 1 × 0.5 × 0.5 mm3 (250 nl), the application of single-crystal EPR to [FeFe]-hydrogenase was severely limited by the typically small crystal sizes with volumes in the nanoliter to sub-nanoliter range (0.3 × 0.1 × 0.1 mm3; 3 nl). In general, this limitation, due to the size and general flexibility of proteins, makes growing large crystals problematic [144]. Technical developments for EPR have allowed, for the first time, measurement of the [2Fe]H subcluster in the Hox state within a 3 nl protein crystal to be measured and the g-tensor proposed [92]. Further experiments to measure the hyperfine couplings are required to better evaluate DFT EPR calculations. Of course, the whole toolset of isotope doping can be performed if single crystals form.

figure d

Vibrational spectroscopy

The term ‘vibrational spectroscopy’ comprises two main techniques: Fourier-transform infrared (FTIR) spectroscopy and Raman spectroscopy. FTIR spectroscopy probes the frequency of molecules that undergo a distance-dependent change in dipole moment, whereas bonds with a vibrational change in polarizability are investigated by Raman spectroscopy most conveniently. The intense absorbance of liquid water and the amide bonds of the protein backbone (vC=O, vN–H) render FTIR spectroscopy a key technique in protein research [145]; however, probing metal centers directly, Raman spectroscopy is more powerful [146]. Lastly, nuclear resonance vibrational spectroscopy (NRVS) is an X-ray technique that exploits the recoil fraction of Mössbauer absorption, e.g., in 57Fe-labeled samples of hydrogenase [147,148,149], resulting in sidebands that report on vibrations that are coupled to the mental centers.

Characterization of different H-cluster intermediates

The active site cofactor of [FeFe]-hydrogenase carries CN and CO ligands, which was first rationalized as a common feature among hydrogenase by Albracht et al. in 1996 [53]. The stretching frequency of these groups is observed in a regime devoid of (strong) water or protein bands, and the absorption coefficient of both CN and CO is very high. Therefore, FTIR spectroscopy is a valuable tool investigating the H-cluster directly [150]. Following the publication of the XRD structures of DdHydAB and CpI, FTIR spectroscopy was exploited to characterize Hox, Hred, and Hox-CO [151,152,153]. The lack of a low-frequency µCO band in Hred (whose spectrum qualitatively resembles Hsred, compare Fig. 9) appeared to match the ‘semi-bridging’ CO ligand in the reduced XRD structure of DdHydAB [60].

Fig. 9
figure 9

Fits for the FTIR spectra of the same H-cluster states as in Fig. 7, additionally including the diamagnetic states Hred´, Hred, and Htrans. Characteristic frequencies are annotated as observed in DdHydAB (Htrans/Hinact, marked with an asterisk) and CrHydA1 (everything else)

Things became more complicated when CrHydA1 was analyzed by FTIR spectro-electrochemistry in 2009 [154]. The observation of a µCO band under reducing conditions was interpreted to hint at an isomer of Hred carrying a bridging CO ligand, seemingly explaining how the enzyme stabilizes the ‘rotated geometry’ of the diiron site [155,156,157]. Such ligand was absent in Hsred, first observed as a minor species in DdHydAB [107] and thoroughly characterized in CrHydA1 in 2012 [158]. Investigating CrHydA1 by resonance Raman spectroscopy and FTIR spectroscopy, various authors reported that the µCO band of Hred must be assigned to another one-electron reduced state, Hred´ [34, 159]. This intermediate, reduced at the [4Fe]H cluster, has been observed in H-cluster variant PDT earlier, where the central amine of the ADT ligand is replaced by a propane head group that efficiently prevents protonation and reduction of the diiron site [119]. Figure 10A highlights how small energetic differences (< 10 cm−1) between Hox/Hred´ and Hred/Hsred reflect the redox state of the [4Fe]H cluster while redox changes at the [2Fe]H site result in larger shifts (> 50 cm−1). The conversion from Hox into Hred includes an additional change in geometry: upon reduction, the Fe–Fe distance increases by 0.15 Å [160], the µCO ligand is liberated and ends up in a terminal position. Systematically comparing different models of the Hred or Hsred states, the IR pattern in Fig. 9 represents an H-cluster intermediate with a bridging hydride (µH) most likely [161]. Stripp and Haumann have argued that such intermediates are not part of the catalytic cycle of hydrogen turnover but might play a role in ‘regulatory’ [FeFe]-hydrogenases [25], which often adopt Hred as a quasi-resting state [162, 163]. Interestingly, FTIR spectroscopy has shown that the CO ligand can be arrested in Fe–Fe bridging position upon reduction at cryogenic temperatures [94], giving rise to redox states HredH+ and HsredH+, presumably protonated at the ADT ligand [164, 165]. Due to the lack of direct experimental evidence for a µH ligand in Hred/Hsred or +NH2 in HredH+/HsredH+, the controversy whether these states are catalytically relevant maintains [25,26,27]. We will return to a brief discussion of the catalytic cycle at the end of this section.

Fig. 10
figure 10

A FTIR difference spectra of the electrochemical reduction of CrHydA1 at pH 9 (‘Hred´–Hox’) and pH 6 (‘Hsred–Hred’). The small band shifts result from redox changes at the [4Fe]H cluster (only the strong Fed–CO bands are annotated). Note that the H-cluster lacks a µCO signal in the diiron site-reduced states Hred and Hsred. The band at 1972 cm−1 (*) remains unassigned. B FTIR difference spectra of the reaction of cofactor variant CrHydA1ODT with H2 (H2O) or D2 (D2O) in the CO regime of the H-cluster. Negative and positive bands are assigned to Hox and Hhyd, respectively. While none of the terminal CO ligands are affected, the downshift of the µCO band from 1862 cm−1 (blue) to 1856 cm−1 (orange) upon deuteration proves the presence of a terminal hydride in Hhyd

While the EPR spectrum of Hhyd (Fig. 7) is in good agreement with an Htrans-like electronic structure ([4Fe]H+–[Fep(II)–Fed(II)]), vibrational spectroscopy was employed to identify the terminal hydride ligand (tH) directly. Figure 10B shows an FTIR difference spectrum in the H-cluster’s CO regime that results from the reaction of CrHydA1 cofactor variant ODT with H2 (in the presence of H2O) or D2 (in the presence of D2O). In this semi-artificial enzyme, the natural NH head group is exchanged against oxygen [69], which leads to an inhibition of proton transfer and accumulation of Hhyd when reduced with hydrogen gas [166]. Upon reaction with D2 instead of H2, a specific downshift of the µCO band at 1862 cm−1 is observed (Fig. 10B) that has been explained by vibrational energy transfer between tH and µCO [115]. The ‘deuteride’ ligand (tD) shows no such ‘trans coupling’, therefore, the µCO vibration shifts to slightly lower energies. Investigating 57Fe-labeled CrHydA1ODT by NRVS the low-frequency Fe-H bending modes have been identified and demonstrated to shift by ~ 100 cm−1 in deuterated samples [149]. This is direct proof for the presence of a terminal hydride ligand in Hhyd, a finding that was independently confirmed by NMR spectroscopy on CrHydA1 cysteine variant C169A [167].

These studies demonstrated an enrichment of Hhyd in different variants and established its FTIR fingerprint as a marker for amino acid residues relevant in catalytic proton transfer [55]. However, Stripp et al. were able to demonstrate the accumulation of Hhyd in wild-type [FeFe]-hydrogenase as well, reacting the enzymes with H2 at low pH [166]. Under these conditions, the oxidized starting state differs from Hox by a small shift to higher frequencies, which inspired us to propose an oxidized state with a protonated cysteine at the [4Fe]H cluster, HoxH [34]. The pH dependence of Em in the Hox/Hred´ conversion and the pH-dependent equilibrium between Hred and Hred´ under reducing conditions facilitated a model that requires proton transfer in order to stabilize the reduced [4Fe]H+ cluster [168,169,170]. Given the direct interconversion of HoxH and Hhyd, the same logic applies to the hydride state [166].

The catalytic cycle

Similar to the question whether the reduced diiron site is protonated at the ADT ligand (HredH+ and HsredH+) or in Fe–Fe bridging position (Hred and Hsred), direct evidence for proton transfer (PT) toward the [4Fe]H cluster in the formation of Hred´ is missing. In fact, Birrell et al. were not able to reproduce the pH dependence of the Hox/Hred´ transition [32], suggesting a simple electron transfer (ET) step rather than proton-coupled electron transfer (PCET). The different interpretations of the catalytic mechanism of [FeFe]-hydrogenase can be illustrated in a simplified cycle comprising of either five steps or three steps (Fig. 11).

Fig. 11
figure 11

Catalytic mechanism of [FeFe]-hydrogenase including the five main H-cluster states. The cubane represents the [4Fe]H cluster in the oxidized (+ 2) or reduced state (+ 1). The rectangular box depicts the diiron site in the mixed-valence (I/II), reduced (I/I), and ‘super oxidized’ state (II/II). {H} and (H) refer to protonation of the [4Fe]H cluster and diiron site, respectively. See text for details on the ‘five-step cycle’ (ABCDE) and the ‘three-step cycle’ (A*B*E)

In H2 evolution direction, the ‘five-step model’ starts with external reduction of [4Fe]H cluster (A), followed by intramolecular ET from [4Fe]H to the diiron site and protonation of the ADT ligand (B). This step is rationalized by the assumption that a reduction of the diiron site does not change the µCO geometry of the H-cluster [164], ruling out the formation of a bridging hydride. A chemically reasonable explanation for the observed pH dependence [165] would be protonation of the ADT ligand, hence we are using the HredH+ nomenclature here. External re-reduction of [4Fe]H produces HsredH+ (C) before PCET from the ADT ligand to the distal iron ion triggers the formation of Hhyd (D). The later step, hypothetical for a long time, has been demonstrated by cryogenic FTIR in photochemical experiments in 2020 [117]. With a second proton coming in, the hydride ligand combines to H2, which restores the Hox state (E). The three-step model starts with PCET in the formation of Hred´ (A*). Note that the proton {H} is not part of the hydrogen turnover reaction but binds at the [4Fe]H cluster to impede intramolecular ET to the diiron site. In a second step of PCET, Hred´ is converted into Hhyd directly (B*) before H2 is released upon protonation of the hydride ligand as described above (E), either recovering Hox (when {H} is released from the [4Fe]H cluster) or HoxH. The later state is observed under turnover conditions at low pH exclusively [170].

Future directions

The strong and distinct IR absorption of the H-cluster’s CO and CN ligands has facilitated time-resolved ‘flash photolysis’ experiments using quantum cascade laser (QCL) radiation [171,172,173]. Within hundreds of µs, Dyer et al. investigated the kinetic competence of various reduced H-cluster states concluding that, “the rate limiting step of catalysis is not the interconversion of any of the known intermediate states” [172]. Understanding hydrogen turnover with [FeFe]-hydrogenase hence means understanding the electronic properties of the H-cluster rather than the redox kinetics. To this end, ultrafast two-dimensional infrared (2DIR) spectroscopy may help [174]. Using this technique, the vibrational excitation of single ligands and energy transfer among ligands can be probed without the need for isotope editing [175], as demonstrated by Horch et al. at the example of [NiFe]-hydrogenase [176,177,178]. Similar experiments have not been published on [FeFe]-hydrogenase yet; however, an application can be expected in the future.

Moreover, 2DIR spectroscopy is a technique used to probe the coupling between cofactor and protein environment, identifying structural or electrostatic changes in the second coordination sphere. Such data are included in difference spectra ‘beyond’ the energy regime of the active site H-cluster. In [NiFe]- and [FeFe]-hydrogenase, in situ FTIR difference spectroscopy helped understanding proton transfer between active site and solvent [35], highlighting the role of SH- and COOH-containing amino acids, and water molecules [56, 179]. Generating difference spectra is severely complicated when light is excluded as a trigger. While certainly a helpful tool, visible light irradiation affects the H-cluster in various ways [117, 175] and must be handled with care. Gas titrations present a unique way to trigger the reactivity of hydrogenases and generate ATR FTIR difference spectra [150]. Such experiment may elucidate the process of oxidative H-cluster inactivation, CO inhibition, or H2 binding.

While biophysical experiments are mostly performed in condensed phase, structural data with atomistic resolution is obtained from XRD on protein crystals. Bridging this gap, FTIR microspectroscopy and FTIR spectro-electrochemistry has been used to probe [NiFe]-hydrogenase crystals [180, 181] and [FeFe]-hydrogenase crystals [93]. In the future, we expect that FTIR spectroscopy and EPR spectroscopy (92) will be used on hydrogenase crystals routinely.

figure e