Open Access Article
Satoru Tsushima
*ab and
Koichiro Takao
*b
aInstitute of Resource Ecology, Helmholtz-Zentrum Dresden-Rossendorf, Bautzner Landstraße 400, 01328 Dresden, Germany. E-mail: s.tsushima@hzdr.de
bLaboratory for Zero-carbon Energy, Institute of Integrated Research, Institute of Science Tokyo, 2-12-1 N1-32, O-okayama, Meguro-ku, 152-8550 Tokyo, Japan. E-mail: ktakao@zc.iir.isct.ac.jp
First published on 2nd January 2026
This Perspective article reviews the current knowledge regarding the interaction of actinide elements with biomolecules, particularly amino acids, peptides, and proteins. We assess the significance of these interactions, especially in connection to nuclear waste disposal and the potential role of these interactions in the origin of life. Actinides have been observed to form stable complexes with carboxylate groups, resulting in oligomerization and affecting their environmental mobility. The text discusses the complex coordination chemistry of actinides, including the prevalence of hexanuclear An4+ clusters, and the implications of these findings for actinide transport and bioavailability as well as remaining challenges especially for mechanistic and thermodynamic aspects of this chemistry. Recent discoveries of lanthanide- and actinide-dependent enzymes, including methanol dehydrogenase and lanmodulin, suggest a potential for these elements to have been actively involved in early metabolic processes, rather than solely acting as environmental stressors. Despite the absence of direct evidence connecting natural reactors to the process of abiogenesis, a comprehensive understanding of actinide–biomolecule interactions is imperative for the evaluation of the nuclear geyser model and the resolution of the long-term challenges posed by the management of radioactive waste.
During this period, the isotopic proportion of fissile 235U was as high as 20%, significantly higher than the current 0.72%. Under such circumstances, even low-grade uranium ore could have served as a functional nuclear reactor, given the presence of water as a neutron moderator. It should also be considered that the Moon was closer to the Earth in the Hadean Era, resulting in enormous tidal variations.
Natural nuclear reactors could have facilitated the circulation of material and energy, enabling chemical reactions leading to the synthesis of complex organic compounds including amino acids, peptides, and nucleobases. This process may have generated the earliest forms of metabolism, with uranium acting not just as fissile material but also as oligomeric catalyst and coordinating ion for prebiotic metalloenzymes.5 Furthermore, increasing evidence for proteins and bacteria dependent on lanthanides or actinides6–11 supports the notion that these elements played an essential role in early metabolism. Paleontological studies have unearthed the earliest eukaryotic fossil from the Paleoproterozoic Francevillian Group in Gabon, a region where natural nuclear reactor once existed.12
Taking into account these observations, and given the rich redox chemistry exhibited by actinides,13 along with their remarkable coordinating ability,14 their interaction with amino acids and subsequent oligomerization likely played a key role in prebiotic evolution and homochirality. Previous studies on the interaction between actinide and amino acid or with primitive peptide such as glutathione have not discussed the potential importance of these interactions in prebiotic systems.
Despite that the nuclear geyser model still remains hypothetical, and linking actinide with the origin of life being controversial, it is also important to assess whether current biochemical knowledge of actinide-amino acid interaction aligns with the nuclear geyser model to test its validity. Fig. 1 depicts the circulation of materials on Hadean to Eoarchean Earth with the focus on uranium (+ minor actinides) and amino acids. This model proposes that natural nuclear reactors contributed to amino acid formation, which then facilitated the emergence of actinide carboxylates and related oligomers, transported in the geosphere. This process may have led to the formation of primitive proteins and enzymes potentially utilizing actinide ions.
This Perspective aims to provide a systematic review of past studies on the interactions of actinides with amino acids and proteins, to consider the broader implications of this rapidly developing field of research. Ultimately, we would like to help verify the nuclear geyser model from a chemical as well as biophysical perspectives rather than from an astronomical standpoint, and to suggest a research direction for that purpose.
Surprisingly, the number of reported crystal structures of UO22+ complexes with AAs is rather limited. A search of the Cambridge Structural Database revealed only two documented instances of crystal structures of UO22+ complexes with AA, both of which are Gly and its deprotonated form (Gly−) through O^O bidentate coordination.27,28 Other examples include its N,N,N-trimethylated derivative (i.e., betaine),29 and iminodiacetate.30 An N^O bidentate coordination, which would be expected to form a stable 5-membered chelate ring according to conventional coordination chemistry, has not been observed in a crystal structure of UO22+. However, this binding mode is plausible, as evidenced by the NMR spectrum of an aqueous UO22+–Gly–F− system,26 and has been demonstrated with related ligands such as iminodiacetate,31,32 nitrilotriacetate,33 EDTA4−,33 and His-based Schiff bases.34 Another O^O coordination mode of Gly and Ala is utilized to form infinite 1D coordination polymers of UO22+ (Fig. 2(A) and (B)), where syn–syn bridging of these AAs occurs between neighbouring UO22+ units,23,35,36 thereby inhibiting further growth of the UO22+–μ3-O(H) 2D sheets of meta-schoepite, (UO2)2O(OH)6·5H2O, (Fig. 2(C))37 that initially form upon UO22+ hydrolysis.
![]() | ||
| Fig. 2 Infinite 1D coordination polymers of UO22+ composed of uranyl pentagonal bipyramids that are chelated by Gly (A) and Ala (B) in a syn–syn bridging (top and side views) reported by Forbes and coauthors35 and infinite 2D sheet of meta-schoepite, (UO2)2O(OH)6·5H2O (C), along a- and b-axes reported by Weller et al.37 | ||
In the presence of carboxylate (RCOO−), it is necessary to consider the competition between the hydrolytic polynucleation of An4+ and coordination interactions of RCOO− with An4+, that may occur in a concurrent manner. Although a variety of coordination manners of RCOO− are theoretically possible, the syn–syn bridging by μ-RCOO− is exclusively observed in the cases of An4+. In the early era of actinide coordination chemistry, isolated compounds from An4+–carboxylate systems were erroneously assigned to simple compound such as [An(RCOO)4]n, where the metal ions were supposed to be interconnected by μ-RCOO−, to form infinite 1D coordination polymers.15 Even in the aftermath of the hydrolysis of An4+, the polymeric structure were considered to remain largely intact, with partial replacement of RCOO− by OH−. Today, it is hard to accept such an infinite 1D coordination polymer to occur as a soluble species in mother liquors. In fact, already in 1920s, the presence of Th(IV) oligomeric species has been suggested through the study on the Th4+–HCOO− system to propose the compound [Th3(OH)5(HCOO)6]·A·nH2O (A = ClO4, NO3, HCOO, SCN, ClO3) first by Weinland and Stark in 1926
39 and later by Reihlen and Debus in 1929.40
Eight decades later, we have successfully identified the precise molecular and crystal structures of polynuclear complexes of Th4+ and U4+ with HCOO− by means of single crystal X-ray diffraction.41 These An4+ compounds exhibit common molecular structures consisting of discrete hexanuclear complexes, [An6(μ3-O)4(μ3-OH)4(μ-HCOO)12(H2O)6] (An = Th, U; Fig. 3(A)), where μ3-O2−/OH− are included as tripodal connectors for three An4+ ions to form an An4+ hexamer (Fig. 3(B)). The octahedral core of the {An4+}6 structure is further decorated by μ-HCOO−, which serves to bind each pair of neighbouring An4+ ions. Note that the compositions of the compounds we identified are nearly double as those suggested in 1920s. This results in the following formula: [Th6(OH)10(HCOO)12]·A·nH2O = [Th6(μ3-O)4(μ3-OH)4(HCOO)12]·2HA·2(n+1)H2O, where some variations in co-crystallized salts/acids can be found. Furthermore, extended X-ray absorption spectroscopy (EXAFS) and UV-vis spectroscopy successfully confirmed the formation of these [An6(μ3-O)4(μ3-OH)4(μ-HCOO)12(H2O)6] complexes in aqueous solutions (Fig. 3(C) and (D)), thereby establishing a clear connection between solution chemistry and solid-state structures. Although the formation of such an An6O8 core unit has been previously observed in diphenylphosphate and triflate systems,42,43 in which the deprotonated phosphate diesters or triflate anions are located at the edges of the {An4+}6 octahedron in a manner analogous to μ-HCOO− of Fig. 3(A), research activities on polynuclear complexes of hydrolysed An4+ with μ-RCOO− and related organic molecules have been revitalized.
![]() | ||
| Fig. 3 Structural chemistry of An4+–RCOO− hexanuclear complexes reported so far. Whole molecular structure of [Th6(μ3-O)4(μ3-OH)4(μ-HCOO)12(H2O)6] (A, where H atoms were omitted for clarity) and its An6O8 core (B) together with Fourier transforms of k3-weighted LIII-edge EXAFS spectra of An4+ (An = Th (C, 50 mM),41 U (D, 15 mM),41 Np (E, 25 mM),46 Pu (F, 3.9 mM)51) in aqueous solutions containing 1 M RCOOH (R = H, CH3) under specific pH conditions. Panel (C), (D), (E), and (F) were reproduced from ref. 41, 46 and 51, respectively, with permission from Wiley-VCH GmbH (Copyright 2009) and American Chemical Society (Copyright 2012, 2022). | ||
As a matter of fact, this class of chemistry was further expanded to the heavier tetravalent actinides, Np4+ and Pu4+. In connection with our findings of the hydrolysed hexamers of Th4+ and U4+ decorated by μ-HCOO− described above,41 and the following works with PhCOO− (ref. 44) and CH3COO−/ClCH2COO−,45 we have explored solution coordination chemistry of Np4+ under the presence of RCOO− (R = H, CH3) in aqueous systems by means of UV-vis and EXAFS.46 Predominant formation of [Np6(μ3-O)4(μ3-OH)4(μ-RCOO)12] was commonly observed in both R = H, CH3 systems of 1.00 M RCOOH at pH ≳ 2 as pronounced by representative intermetallic interactions in Fourier transforms of the EXAFS spectra (Fig. 3(E)) at R + Δ = 3.80–3.81 Å and 5.39–5.40 Å for adjacent and diagonal pairs of Np atoms, respectively. The presence of the terminal H2O observed in the X-ray structures with the lighter An4+ (Fig. 3(A)) could not be unambiguously confirmed in the case of Np4+ complexes due to uncertainty in the EXAFS analysis. The coordinating water molecules may not be present, considering the smaller ionic radius of Np4+ (0.98 Å, CN = 8) compared with Th4+ (1.05 Å for CN = 8, 1.09 Å for CN = 9) and U4+ (1.00 Å for CN = 8, 1.05 Å for CN = 9).47 However, this issue remains controversial, as both the presence and the absence of an H2O molecule in hexanuclear complexes of the smaller Pu4+ have been reported in the literatures (vide infra).48,49 The EXAFS analysis of aqueous solutions revealed a distinction between the distances of Np–μ3-O2− (2.22–2.23 Å) and Np–μ3-OH− (2.42–2.43 Å). This observation is noteworthy, as such distinctions are often ambiguous in crystal structures due to the inherent disorder of μ3-O2−/OH− as exemplified by [Th6(μ3-O)4(μ3-OH)4(RCOO)12(H2O)6] (R = H, CH3, CH2Cl)45 deposited under the absence of NaClO4 in the mother liquors. An analogous Np4+ hexamer decorated by PhCOO− was also observed in an EtOH/aqueous 1
:
1 mixture, while this solid phase was characterized only by powder XRD.50 Subsequent to the X-ray crystallographic studies on Pu4+ hexamers with Gly and 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetate (HnDOTAn−4), which we will discuss later,48,49 the CEA group undertook an examination of solution coordination chemistry in an aqueous Pu4+–CH3COO− system.51 In a manner analogous to the Np4+ case,46 the Pu6O8 cluster decorated by the syn–syn bridging CH3COO− such as [Pu6(μ3-O)4(μ3-OH)4(μ-CH3COO)12(H2O)6] was found to be formed predominantly under the presence of 1 M CH3COOH at pH 4.5 (Fig. 3(F)), where H2O molecule capping each vertex of the {Pu4+}6 octahedral core was also included in the EXAFS fit with a fixed coordination number for each shell. In a manner analogous to Np4+, PhCOO− also allowed to yield a Pu4+ hexanuclear cluster, [Pu6(μ3-O)4(μ3-OH)4(PhCOO)12(H2O)4], wherein only eight PhCOO− are involved in the syn–syn bridging between neighbouring Pu4+ atoms, and the remaining four PhCOO− are either monodentate or bidentate-chelating to bind to a single Pu4+.50
While the majority of reported works remain focused on Th4+ and U4+, in contrast to the limited numbers of known examples of Np4+ and Pu4+, several review articles are currently available that offer comprehensive summaries on the coordination chemistry of polynuclear An4+ complexes with RCOO− and polycarboxylates.38,52–56 These two ligand systems, RCOO− and polycarboxylates, are closely related to each other, making it difficult to distinguish them distinctly. However, the former (RCOO−) places greater emphasis on the fundamental aspects of this structural chemistry class, and it is also relevant to understanding the reaction mechanisms of AnO2 colloid formation, especially for PuO2 colloidal nanoparticles. The main concern of the latter (polycarboxylates) is the utilization of An6O8 core unit as building blocks of metal–organic frameworks (MOFs) mimicking non-radioactive tetravalent transition metal ions, most typically Zr4+, to afford UiO-6X (X = 6, 7) type MOF structures.
In contrast to RCOO−, the coordination chemistry of An4+ with AAs is even more limited despite their higher biological relevance compared to RCOO−. In the majority of instances, this coordination chemistry is exclusively confined to the systems with Gly. The [An6(μ3-O)4(μ3-OH)4(μ-Gly−)x(μ-Gly)12−x(H2O)6]12−x was obtained commonly for An = Th (x = 6)57,58 and U (x = 0).59 In these cases, the syn–syn bridging manner of the carboxylate moiety of Gly and/or its deprotonated form (Gly−) followed that observed for RCOO− (Fig. 4(A)). Using EXAFS, we observed predominant formation of [Np6(μ3-O)4(μ3-OH)4(μ-AA)12]y+ (AA = Gly, Ala, Cys; 0 ≤ y ≤ 12) in aqueous solutions under acidic conditions even below pH 1 (Fig. 4(B)).46 It is important to note that no significant effects of sidechains (R) of Ala (R = –CH3) and Cys (R = –CH2SH) have appeared in the actual Np4+ hexamer formation. The initial report on the Pu4+ version of the hexanuclear coordination clusters in this context is X-ray structure determination of [Pu6(μ3-O)4(μ3-OH)4(μ-Gly)12(H2O)6]12+ co-crystallized with Li+, Cl−, and crystalline water molecules,48 where μ3-O2−/OH− are not distinguishable due to their disorder. Each vertex of the {Pu4+}6 core is capped by a H2O molecule (Pu–Ow = 2.688 Å) to form a nine-coordination geometry around each Pu4+. This is notable because the ionic radius of Pu4+ is smaller than that of Np4+. Due to the zwitterionic nature of AAs in the charge-neutral state, the syn–syn bridging mode of their carboxylate group is offered in both the original (H3N+–CHR–COO−) and the deprotonated (H2N–CHR–COO−) forms. This phenomenon underlies the formation of [An6(μ3-O)4(μ3-OH)4(μ-AA−)x(μ-AA)12−x]12−x (AA−: deprotonated AA) even in acidic conditions, as evidenced in the Np4+ case.46 However, the hydrolysis tendencies of An4+ of interest should also be taken into account.
![]() | ||
| Fig. 4 Structural chemistry of M4+–AA systems reported so far, where H atoms were omitted for clarity unless specified. (A) [Th6(μ3-O)4(μ3-OH)4(μ-Gly−)6(μ-Gly)6(H2O)6]6+.57 (B) Fourier transforms of k3-weighted Np LIII-edge EXAFS spectra of [Np6(μ3-O)4(μ3-OH)4(μ-AA)12]y+ in aqueous solutions (AA = Gly, Ala, Cys; 0 ≤ y ≤ 12).46 (C) [Zr6(μ3-O)4(μ3-OH)4(μ-bet)8(κ-bet)4(H2O)4]12+.62 (D) [Pu6(μ3-O)4(μ3-OH)4(HDOTA)4(H2O)6].49,63 (E) [Th2(H2DOTA)2(H2O)10]4−.63 (F) [Th(μ-Gly)3(H2O)2]n4n+.57 H atoms in panels (A) and (C)–(E) and terminal CH3 groups in panel (C) were omitted for clarity. Panel (B) and (C) were reproduced from ref. 46 and 62 with permission from American Chemical Society (Copyright 2012) and Elsevier (Copyright 2012), respectively. | ||
The scarcity of knowledge regarding AA systems is not confined solely to An4+; it also pertains to tetravalent Group 4 metal ions in the periodic table. This tendency is particularly evident in the case of Zr4+, which has undergone extensive development in this direction, though the only known hexanuclear complexes with AA are those with Gly60,61 and its N,N,N-trimethylated variant (i.e., betaine).62 In the former system, the assignment of the positions of H+ on Gly remains ambiguous, as evidenced by the different formulas [Zr6(μ3-OH)8(μ-Gly)4(μ-Gly−)4(H2O)8]12+ and [Zr6(μ3-O)4(μ3-OH)4(μ-Gly)8(H2O)8]12+ in the articles from the same group. The reduced number of syn–syn bridging carboxylates compared with the An4+–Gly systems described above would be ascribed to steric demand arising from the smaller ionic radius of Zr4+ (rZr = 0.84 Å, rAn = 0.96–1.05 Å at CN = 8).47 In the latter case of betaine, the N-terminal of Gly was quaternarized by trimethylation to resolve the ambiguity problem of H+, where the zwitterionic form of betaine, (H3C)3N+CH2COO−, can be maintained during complexation. As a result, preparation and structural characterization of [Zr6(μ3-O)4(μ3-OH)4(μ-bet)8(κ-bet)4(H2O)4]12+ (bet: betaine) in conjunction with its Hf4+ analogue proved to be successful, as demonstrated in Fig. 4(C). In this study, half of the water molecules present in the preceding Gly systems were substituted with monodentate betaine, κ-bet. We have also attempted to synthesize Zr4+ hexamers with other ordinary AAs such as Ala, Val, Cys, and Leu to simulate the An4+–AAs coordination chemistry, but no crystalline deposits could be obtained.62 The factors that disturb the crystallization of Zr4+ hexamers with these AAs could not be fully identified. Nevertheless, there is no rationale for prohibiting the formation of the M6O8 complexes with AAs other than Gly for any M4+, including An4+, in aqueous solutions. As previously mentioned, the predominant formation of Np4+ hexamers with Ala and Cys was already confirmed by EXAFS.46
In relation to AAs, 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (H4DOTA) was employed to stabilize the hydrolyzed hexanuclear complexes of An4+, [An6(μ3-O)4(μ3-OH)4(HDOTA)4(H2O)8] (An = U, Np, Pu, Fig. 4(D)),49,63 where its triply deprotonated form, HDOTA3−, is involved in these An4+ hexamers. Only one out of three deprotonated carboxylates in each HDOTA3− is involved in the syn–syn bridging between the neighbouring An4+ ions, while the other two are bound to An4+ in a monodentate manner. Upon initial observation, it appears that the sole protonated carboxylic group in the HDOTA3− ligand is not committed to any chemical bond. However, a closer look into the structure reveals that it is hydrogen-bonded to μ3-OH−, coordinating H2O, and crystalline H2O. These interactions contribute ultimately to the stabilization of hexanuclear clusters which has reduced number of syn–syn bridging carboxylates compared with other RCOO− and AA systems described above. In contrast, neither intra- nor intermolecular interactions of any non-coordinating N atoms in the cyclen moiety of HDOTA3− occur in the crystal structures of [An6(μ3-O)4(μ3-OH)4(HDOTA)4(H2O)8]. Despite the potential of Th4+ to form a hexanuclear complex with carboxylates as exemplified above, an aqueous mixture of Th4+ and H4DOTA resulted in the formation of a dimeric complex via H2DOTA2− crosslinkers between Th4+ ions, [Th2(H2DOTA)2(H2O)10]4− = [Th(H2DOTA)(H2O)5]24− (Fig. 4(E)), where no olation/oxolation of Th4+ were observed presumably due to its least hydrolytic nature among the An4+ series.64 The similar trend can be observed in the weakest intensity of the Th⋯Th interaction at R + Δ = 3.70 Å in the EXAFS spectrum of the Th4+–HCOO− system (Fig. 3(C)) compared with those of other An4+ (Fig. 3(D)–(F)), indicating the least stability of [Th6(μ3-O)4(μ3-OH)4(μ-RCOO)12(H2O)6] among the An4+ series. A ClO4− salt of [Th(μ-Gly)3(H2O)3]n4n+ coordination polymer of Fig. 4(F) without hydrolysis of Th4+ was also yielded in a relatively acidic system at pH 1.0 as an initial condition, although it is generally difficult to identify the final pH of the mother liquor after slow evaporation of the solvent for crystal deposition.57
To date, the previous studies on actinide hexamers have exclusively focused on structural chemistry, because molecular structures of such polynuclear species are the most distinctive feature of this class of coordination compounds. EXAFS is a valuable tool for the analysis of the structures of chemical species that are present in solution. However, through this method, only the averaged radial distributions of coordinating and related atoms around a central metal may be obtained, whereas it is hard to get any information about angular functions of atomic coordinates of interest, unless specific multiple scattering is available in a system of interest. SCXRD, in contrast, is the most powerful technique for precise structure determination, and therefore, is widely employed. However, as illustrated by our preliminary trials for Zr4+–AA systems,62 crystallization of a coordination compound of interest often appears to be rather challenging. Furthermore, even if crystallization is successful, the subsequent structural analysis of such an intricated compound is not always straightforward. Mainly due to crystallinity issues and remarkable disorder in atomic coordinates, the structure refinement often needs to be compromised by treating some or all non-H atoms solely with isotropic temperature factors and/or by excluding H atoms assignments. It is noteworthy that, despite the fact that a typical hexanuclear structure of [An6(μ3-O)4(μ3-OH)4(μ-RCOO)12] is basically charge-neutral, there has been no reported crystal structure composed solely of actinide hexamer units. Instead, the reported structures invariably involve salts/acids (cations: N2H5+, Na+, Li+, H+; anions: ClO4−, NO3−, Cl−, HCOO−, SCN−, ClO3−, etc.) and/or solvent molecules. While these additional species do not directly interact with An4+, they are nevertheless likely to play important roles to hold the [An6(μ3-O)4(μ3-OH)4(μ-RCOO)12] molecule through non-covalent interactions, most typically hydrogen bonding, to construct its whole crystal structure for successful deposition of a compound of interest. Accordingly, the selection or design of a co-crystallizing salt, acid, and/or solvent should be made with the objective of facilitating the crystallization of An4+ polynuclear species and to further exploring the structural chemistry of hydrolytically-polymerized discrete An4+ coordination clusters decorated by carboxylate-based ligands.
In this context, the observed difficulty in crystallization in the AA systems can be attributed to the variable total electric charge of the M6O8 clusters, arising from degree of freedom of protonation/deprotonation at the terminal amino groups. Our quaternarization approach for the terminal N atom of AAs to immobilize the positive charge on it would be straightforward to resolve this issue even in An4+–AA systems.62 Indeed, N,N,N-trimethylated AA derivatives have already been prepared as bis(trifluoromethylsulfonyl)imide (Tf2N−) salts,65 while the original aim of that work was development of a new class of AA-based ionic liquids. Although deprotonation from such AA-derivatives to liberate them from Tf2N− counter anion is still required for utilization of such zwitterionic betaine-type AAs, exploration of the structural chemistry of An4+ hexamers with AAs could be anticipated in such a manner.
To the best of our knowledge, the sidechains of Ala (R = –CH3) and Cys (R = –CH2SH) play only a bystander role during the formation of Np4+ hexamers in aqueous solutions (Fig. 4(B)).46 Due to the limited exploration of An4+–AA systems, the impact of the sidechains of AA on the coordination chemistry of An4+ remains to be elucidated. This is particularly evident for acidic AAs, such as Glu (R = –C2H4COOH) and Asp (R = –CH2COOH), that comprise carboxylates within their sidechains. In the case of peptides and proteins, in which the carboxylic groups of AAs on the backbone are no longer available for coordination after their condensation with the –NH2 of the neighbouring AA to form peptide bonds, the presence of carboxylic groups on the sidechains increases its significance. As a matter of fact, the EXAFS study by Daronnat et al. reported that Pu4+ forms a Pu6O8 cluster after binding to a wild-type calmodulin peptide, CaMWT (DKDGDGYITTKE).66 This finding was confirmed by Pu⋯Pu interactions at R + Δ = 3.8 Å and 5.2–5.3 Å, which are representative signatures of the An4+ hexamers (see Fig. 3(C–F) and 4(B)). Subsequent molecular dynamics simulations suggested that Pu4+ can be solvated by H2O molecule(s) even after being captured by the coordination site of CaMWT. This process enables the initiation of the hydrolysis of Pu4+ to form the Pu6O8 structure. In contrast, this phenomenon was not observed when another calmodulin peptide with a mutation, CaME (DKDGDGYIEAAE), was employed. In this case, no H2O was found to enter the coordination sphere of Pu4+. Despite the lack of comprehensive picture depicting the final accommodation of the Pu6O8 core within CaMWT, this work clearly represents importance of sidechains of AAs in the An4+ coordination chemistry. Herein, interconnections between the fundamental coordination chemistry of small molecular structures of An4+ with organic substances and extended biologically-relevant systems such as peptides and proteins have been established.
An An4+ hexanuclear complex commonly contain an An6O8 core unit, where all An4+–O2−, An4+–OH−, and An⋯An distances follow those in the Fm
m fluorite-like AnO2 crystalline phases (U–O: 2.368 Å, U⋯U: 3.866 Å in UO2; Th–O: 2.424 Å, Th⋯Th: 3.958 Å in ThO2).38 Therefore, a minimal AnO2 crystalline unit can be postulated to be isolated within a hexanuclear complex, wherein all twelve edges of the AnO2 unit cell are occupied by RCOO− to inhibit further hydrolytic interactions of incoming An4+ to μ3-O2−/–OH−. The decoration of the surface of the An6O8 core with RCOO− efficiently terminates the crystal growth into bulk AnO2, and eventually prohibits its deposition. Indeed, the presence of RCOO− enables the preparation of aqueous solution samples with An4+ concentrations as high as ≥10−2 M, as exemplified in Fig. 3 and 4. This concentration level is significantly higher than that of the original aqueous solubility of An4+ (e.g., ≲10−8 M at pH > 4 for An = Pu), which is expected from their hydrolytic nature.67 This finding has critical implications for the concept and strategy for geological disposal of nuclear waste. In relation to this subject, the CEA group has investigated the growth of the PuO2 colloidal nanoparticles from the perspective of H/D kinetic isotope effects (KIE). Formation of the Pu6O8 cluster occurs during the early stage of the reaction, followed by an O–H bond cleavage (probably at μ3-OH− of the hexamer) as a rate-determining step.68
Now, a fundamental question emerges from these observations; does An4+ undergo further hydrolytic oligomerization from the An6O8 unit to ultimately form AnO2 nanoparticles? If so, how? This mechanistic matter is critically responsible for the solubilization of An4+, which is generally believed to be immobilized as sparingly soluble AnO2 phase. Furthermore, it is important in evaluating actinide transport in the geosphere following the disposal of nuclear waste in deep underground repositories. While there have been DFT calculations that have simulated tri-, tetra-, hexa-, and 16-nuclear hydrolyzed Pu4+ clusters decorated by CH3COO−, a plausible oligomerization process is yet to be concluded through experiments.51 Fig. 5 summarizes An4+ oligomer structures that have been characterized to date. The Pu16O23,69 Pu22O32,69 and Pu38O56
69–71 clusters deposited from aqueous mother liquors were identified by X-ray crystallography, where Cl− was employed instead of carboxylates to afford higher nuclearities of Pu4+ in these oligomers. However, no clear evidence of oligomeric nanostructures other than the Pu6O8 clusters in the reaction process was obtained at this time, implying either the absence or minor contribution of the larger oligomers during the formation of the PuO2 nanoparticles.72 In the case of U4+, Falaise et al. obtained a hydrolysed 38mer after solvothermal treatment under controlled hydrolytic conditions in THF.73 This U38O56 cluster bears a striking resemblance to the chloro-decorated Pu4+ 38mer described above, while a mixed ligand system comprising PhCOO− and Cl− has been strategically employed to regulate the growth of UO2. Additionally, Mazzanti's group detected more diverse U4+ oligomers with 6, 13, 16, 24, and 38-nuclearities in the PhCOO−–Cl− mixed ligand systems under the presence of a limited amount of H2O in non-aqueous solvents such as CH3CN and pyridine.74 The 10, 12, and 22mers of U4+ were also suggested to occur in the reaction mixture. In general, the detection of some oligomers with specific numbers of metals should be regarded as merely a “snapshot” and their presence should not be overinterpreted. On the other hand, Pu4+ has been observed to exhibit a greater tendency for colloidal formation and migration,75 as evidenced by the 1.3 km Pu transport observed at the Nevada nuclear test site.76 Therefore, it is highly probable that U4+ and Pu4+ undergo distinct polymerization processes. It is worthwhile that, in most cases, the An6O8 core unit of the smallest unit of the fluorite-like crystalline AnO2 are included in the reported An4+ hydrolysed oligomers with the higher nuclearities as highlighted in Fig. 5. A comprehensive understanding of the tendency of olation and oxolation of An4+ remains elusive, particularly in the context of variation among different actinides.77 Further investigation in this area is necessary to gain a more comprehensive understanding of these phenomena.
![]() | ||
| Fig. 5 Previously identified An4+ (An = U and/or Pu) oxo/hydro oligomers relevant to further polymerization and nanoparticle formation. Numbers in bold indicate the number of actinides within each oligomer. Only species relevant to aqueous chemistry are depicted. The inset at the top right shows the structure of {An}38 core, with a single hexamer unit highlighted in yellow. The figure is based on the works by Nyman,59 Hixon,69 Soderholm,70 Loiseau,73 Wilson,71 Virot,72 and Mazzanti.74 | ||
Besides the mechanistic features described above, it is also mandatory to compile chemical thermodynamic data of An4+ hydrolysed oligomers, particularly those of hexamers, to quantitatively and universally comprehend the phenomena occurring within the aqueous system of interest. This encompasses the behaviour of An4+ in relation to additional concerns, such as adsorption and migration behaviour of An4+ in natural environments related to the nuclear waste disposal. The stability constants of the An4+ oligomers exemplified above represent the most fundamental data necessary to predict the predominance of each species under specific conditions. In particular, the formation of an n-nuclear An4+ cluster is promoted in proportion to the n-th power of its concentration, more correctly “activity”, in accordance with the mass action law. Therefore, simulating the solution chemistry of An4+ at micro- or nano-molar levels under the presence of carboxylate-based ligands is of crucial interest. However, to the best of our knowledge, virtually no research work has been done in this area, in contrast to extensive body of work on structural chemistry. We have previously established the stability constants of [Np6(μ3-O)4(μ3-OH)4(μ-RCOO)12] (R = H, CH3) in conjunction with the related monomeric species, [Np(RCOO)(OH)2]+, which functions as a precursor for the Np4+ hexamer as outlined below:
| 6 Np4+ + 12 RCOO− ⇄ [Np6(μ3-O)4(μ3-OH)4(μ-RCOO)12] + 12 H+ | (1) |
| Np4+ + RCOO− ⇄ [Np(RCOO)(OH)2]+ + 2 H+ | (2) |
β6,12,−12 = 42.7 ± 1.2 and log
β1,1,−2 = 2.51 ± 0.05; R = CH3, log
β6,12,−12 = 52.0 ± 0.7 and log
β1,1,−2 = 3.86 ± 0.03, where log
βn,l,m is the logarithmic gross stability constant of a Np4+–RCOO−–H+ ternary complex with stoichiometries of these components denoted by n, l, and m, respectively. Based on these thermodynamic data, the speciation diagram for each system was successfully plotted as illustrated in Fig. 6(A) and (B),46 where predominant formation of the Np4+ hexamers is predicted to start even under acidic conditions below pH ∼ 2. In Fig. 6(C), pH dependencies of Np4+ solubility under the presence of HCOO− or CH3COO− under different concentration levels (10 mM, 1 M) were calculated from the above stability constants plus hydrolysis data of Np4+ available in the literature,64 and overlaid on that of Np4+ hydrous oxide comprehended by Neck and Kim.67 Note that the aqueous solubility of Np4+ is significantly enhanced especially in the acidic region after loading RCOO−, while these conditions have been arbitrarily selected for simulation. No thermodynamic data are available for any An4+–AA systems even today despite its higher biological relevance and stronger complexation tendency with metal ions including An4+. Further research in the field of thermodynamics is necessary to ascertain the significance of actinide oligomers in environmental contexts.
![]() | ||
| Fig. 6 Speciation diagrams of Np4+–RCOO− systems (R = H (A), CH3 (B))46 as a function of pH together with logarithmic solubility of Np4+ in aqueous solutions (C) under absence (black)67 and presence of RCOO− (R = H (blue), CH3 (red); [RCOO−] = 10 mM (dashed), 1 M (solid))46 at different pH. Panel (A) and (B) were reproduced from ref. 46 with permission from American Chemical Society (Copyright 2020). Panel (C) was adapted from ref. 67 with permission from De Gruyter Brill (Copyright 2001). | ||
Glutathione (GSH) is one of the simplest peptides consisting of just three amino acids (Glu, Cys, Gly) and known to be ubiquitous and found in all cells of the body. Kretzschmar et al.79 conducted a study on the interaction between GSH and uranium(VI). The primary inquiry concerns the manner in which U(VI) binds with GSH. It has been determined that this binding occurs through the involvement of a carboxylic group, rather than a thiol group. This finding is consistent with the earlier study by Fahmy et al.,80 who employed isothermal titration calorimetry (ITC) to investigate the interaction between the two substances. Furthermore, Kretzschmar et al. observed the reduction of U(VI) to U(IV) in the presence of GSH via an intermolecular mechanism in an aqueous medium in vitro. This process induced the formation of nanocrystalline, mixed-valence uranium oxide particles. Kretzschmar et al. also studied the interaction between U(VI) and glutathione disulfide (the disulfide complex of two glutathione molecules, GSSG) to observe the formation of the insoluble, low-mobility U(VI)–GSSG complex.81 Fahmy et al.80 studied the detoxification of uranium by GSH, using anaerobically grown Lactococcus lactis as a model organism. Their data suggest that the primary detoxification mechanism involves the intracellular sequestration of carboxyl-coordinated U(VI) into an insoluble complex with GSH, which is presumably related to the formation of U(VI)–GSSG or U(IV).
The interaction between short tripeptide like GSH and U is significant in the context of redox reaction, rather than due to its binding through one or two functional groups. Usually a longer peptide (>10 amino acids) is required to observe more specific binding patterns. The structural characteristics of calcium-binding proteins, particularly the EF-hand motif, have served as a model for the development of lanthanide-binding peptides, because of the similarity in the size of ionic radius of Ca2+ and Ln3+. The binding of Ln3+ by calmodulin–peptide has been thoroughly studied,82 Pu4+ interactions with two variants of calmodulin–peptide have been also studied,66 and recently such study has been extended to lanmodulin peptides83 and even those with the reversed sequences84 (Fig. 7). So-called lanthanide-binding tags (LBT) with high and selective recognition of lanthanides are also known for more than 20 years,85–87 which was later extended to Am3+-binding research.88 Jeanson and coauthors designed An(III) and An(IV) binding peptides from a “scratch” and made systematic investigation, though their strategy was primarily focused on utilizing acidic amino acids exclusively.89,90
![]() | ||
| Fig. 7 The optimized structures of Eu3+-bound reversed lanmodulin EF Hand 1 peptide obtained by DFT calculations with dispersion corrections and LC-ECP (six 4f electrons incorporated into the frozen core) applied on Eu3+. Green sticks, orange ball, blue sticks depict the peptide, Eu3+, three coordinating water molecules, respectively. Gray sticks show the previously-calculated structure84 using MD simulations based on 12-6-4 LJ potential on Eu3+. Two structures (DFT and MD) are superimposed to highlight their agreement. | ||
Uranium (and thorium) is perhaps one of the most intensively studied elements among lanthanides and actinides for its binding with peptides,91–95 due to their affordability and minimal radioactivity, which facilitates their handling in laboratory settings. However, these elements may not be the most intriguing for study in terms of recovery and separation because they are the most abundant actinide elements in nature. Conversely, their abundance in nature has led to significant research interest in their detection. The potential health implications of human exposure to uranium encompass both radiological and chemical toxicity.96 The average U concentration in drinking water varies significantly among different regions and countries,97,98 and the presence of U in aquifers has the potential to pose health concerns mainly because of its chemotoxicity. The development of its in situ measuring technique is awaited. The development of peptide-based sensors for U/Th detection has emerged as a prominent research area in recent years.99–101 Phosphorylation of peptide is frequently employed to enhance their binding affinity to actinides.102–105 However, this process transforms the peptide into a mere scaffold, prompting the question of whether the use of noble materials like peptides is truly essential when their sole function is to serve as the backbone that hold the phosphoryl groups hands for the capture of metal ions. In an alternative approach, His-decorated phenanthroline diimide was utilized for the recovery of Am3+.106 In this instance, the incorporation of amino acids was intended to enhance solubility and they are not involved in metal-binding. Some prominent chelators, including DOTA (dodecane tetraacetic acid), EDTA (ethylenediaminetetraacetic acid), and NTA (nitrilotriacetic acid) are classified as aminopolycarboxylic acids. These chelators share Gly as a parental complex. The complexation of these chelators with trivalent actinides is a subject of intensive research in the direction of separation of lanthanides and minor actinides.107
Radiopharmaceuticals are also complexes that contain a peptide segment in their structure, as well as a separate metal-capturing chelator component. In the case of radiopharmaceuticals, however, the peptide portion plays a distinct role in binding with a receptor at its extracellular domain, thereby inducing a conformational change in the receptor and activating its intracellular domain. For the metal-capturing component, a strong chelator such as DOTA is typically selected. In recent years, a novel research domain has emerged that utilizes these chelators for radiotherapy. 225Ac-based radiopharmaceuticals have gained increased attention for its application in α-particle therapy, a treatment modality that has been shown to be more effective in the destruction of tumour cells when compared to β-emitting radionuclides.108 In this particular context, as well, phosphorylation has been identified as a promising strategy to enhance the binding affinity of Ac3+.109,110 It should be noted that the coordination chemistry of Ac3+ is still rather poorly understood,111 and studies on its daughter products, such as Fr and At, are even more scarce. It is frequently assumed that Ac3+ persists in binding to the chelator subsequent to its decay to its daughters. However, this is exceedingly improbable, given that Fr+ is the largest cation in the periodic table. It is imperative that further investigation be conducted into the coordination chemistry of these elements (Ac, Fr, At), which have received the least attention in the periodic table so far.
Now, we will shift the discussion to an entirely different topic; the implementation of computer chemistry in this research domain. While the application of computer chemistry, such as density functional theory (DFT), has been extensively implemented in the field of general actinide chemistry, its application in the context of actinide interaction with peptide or protein has been extremely scarce. However, this emerging field has undergone significant development in the past decade, a topic which will be discussed in the remainder of this section. Due to smaller size of peptides in comparison to proteins, they offer greater potential for computer simulation, and the application of computational chemistry in this field is expected to become more widespread in the future. Quantum chemical calculations (e.g. DFT) and classical molecular dynamics (MD) simulations can be employed to study peptide structures and their metal bindings. Each method addresses different aspects and has different strengths. Quantum chemistry is an ab initio method that provides accurate electronic and geometric structures of peptides. In contrast, classical MD simulation is based on empirical parameters and is more suited for simulating dynamic behaviour and conformational changes of peptides. Peptides are known for their structural flexibility, which complicates the prediction of their structures, including those in metal-bound states. Peptide structure optimization frequently lacks any initial structures to begin with. In such cases, the use of DFT can prove challenging in terms of identifying the global energy minimum with reasonable computational cost. The employment of a more expeditious and dynamic approach, such as classical MD simulations, is often a more suitable alternative. There was great advent in this field in the last decade largely due to the parameterization of M3+ and M4+ ions using the 12-6-4 Lennard-Jones-type nonbonded model by the group of Kenneth Merz.112–114 Their parametrization included exotic ions such as Th4+, U4+, and Pu4+. The structures of Eu3+-bound peptide obtained by using their 12-6-4-type potential is comparable to the DFT-optimized structure, which prove their validity (Fig. 7). Hay and coworkers extended this potential (parametrized for water coordination) to the protein system to better reproduce the binding energy for M3+-binding peptides.115 Not all MD programs are compatible with 12-6-4-type potentials, and the widely-used GROMACS software cannot handle this type of potentials, which is a significant setback to the research in this direction. In an another effort, the group of Stefan Grimme has recently extended robust atomistic generic force field GFN-FF to lanthanide and actinide biomolecule system and claimed that the structures they obtained are almost comparable to those using DFT.116 Another interesting and important development by Grimme and his coworkers is the development of composite methods which he calls “Swiss army knife” and utilizes relatively small basis sets and special corrections to achieve high accuracies at a fraction of the computational cost of a calculation approaching the basis set limit. In the realm of computational chemistry, intermediate methods that straddle the divide between quantum chemistry (e.g. DFT) and classical mechanics include the QM/MM approach and the FMO (fragment molecular orbital) method (Fig. 8). These methods possess two key advantages. Firstly, they are significantly less expensive than purely quantum chemistry-based methods. Secondly, they are more accurate in energetics than purely classical mechanics-based approaches. We have recently successfully applied FMO methods in various lanthanide/actinide-containing biological system with the help of group of Yuji Mochizuki.25,117,118
A number of attempts have been made to recover hexavalent uranium (UO22+) using protein-based materials.119–122 However, such technology remains ambitious for practical application considering the relatively low market price of uranium (merely ∼10 times that of copper).123 In practice, there are much more economic approaches for the recovery of uranium from seawater,124–127 though their economic viability still remains suboptimal, necessitating ongoing technological development. Nevertheless, in scenarios where the primary objective is the recovery of Ln/An from wastewater to avert environmental contamination, cost considerations become a secondary issue, and protein-based materials eventually emerge as a viable alternative.128,129 It also remains a significant challenge to separate Ln and An or amongst themselves, especially in the nuclear industry, for which protein-based materials could be utilized.130,131 In the context of metal ion recovery and isolation from a matrix or from wastewater for the purpose of bioremediation, the utilization of abundant materials such as surface layer (S-layer) protein is advantageous due to the heterogeneity of these materials, which does not pose a significant problem.128 On the other hand, in scenarios where the objective is mutual separation of metal ions, the homogeneity of the material becomes a central concern, necessitating the purification of the material. In general, peptide is more appropriate for such purpose rather than protein. This is because only few amino acid residues are anyway committed to metal-binding. Consequently, in the absence of a significant breakthrough in the research, such as the confirmation of bulk, biologically mass-producible proteins that are effective in the separation of actinides and/or lanthanides, further research in this area appears to be challenging. In this particular context, alcohol dehydrogenase (ADH) enzymes have been observed to exhibit the capacity to utilize lanthanides and actinides, and even demonstrate a degree of discrimination among them. This issue will be addressed subsequently in this section.
With regard to the interaction between actinide and protein in human-related scenarios, Creff et al.132 provided an excellent comprehensive review on this topic, which summarizes previous studies conducted since the Manhattan project. Creff et al. describe that primary tasks of previous researches can be classified into three categories; first the identifications of the proteins that are relevant to human-related scenarios, second the investigations on the structures and the affinities of the An-binding proteins, and third to study the impact of uptake of An-bound proteins. Serum proteins have been extensively employed to study its affinity to actinides, as they play crucial role in facilitating metal transportation within the human body, suggesting a potential for involvement in actinide transport processes. This includes studies of actinide interactions with albumin,133 transferrin,134–136 ferritin,137,138 and hemoglobin.139 The collective findings of these studies indicated high affinity between actinide and proteins, with the potential binding site of actinide being either specific or non-specific (Fig. 9). In vivo studies employing rats after implantation with depleted uranium fragments have demonstrated U accumulation in their kidneys and bones. However, the study also exhibited a minor increase in serum U levels, attributable to rapid urinary excretion.140 Therefore it is important to note that while certain proteins possess a high capacity to incorporate actinides, this does not inherently signify that those proteins function as actinide accumulators within the human body. On the other hand, it has been hypothesized that the process of bone turnover involving Th(IV) and Pu(IV) is associated with the presence of hyperphosphorylated proteins, such as osteopontin.141 Phosphorylated proteins are generally identified as strong actinide binders as in the case of phosvitin142,143 and bovine milk protein.144 This is also one of the reasons why phosphorylation is frequently used as a strategy to increase the affinity of peptides towards lanthanides and actinides, as we discussed in the preceding section. Another protein that has been extensively studied in relation to lanthanide and actinide binding is calmodulin, a calcium-binding messenger protein. It has been demonstrated that calmodulin can accommodate up to four Ln3+ (or An3+) ions.117 Cotruvo et al. have discovered calmodulin-like protein which has high selectivity for Ln3+ (and An3+) and named it as lanmodulin.7,145 However, the actinide–calmodulin interaction appears rather irrelevant in connection to human health issues, and its high affinity is discussed and utilized in a separate setting. Overall, there are still many open questions in this field of research. For instance, the genotoxicity of uranium has been demonstrated not to be necessarily associated with U–protein interaction. Rather, U genotoxicity has been suggested to be more associated with DNA strand breaks and chromosome aberrations.96 The issue of actinide toxicity in the human body is not a simple problem that can be attributed to the interaction with several particular proteins. Further comprehensive research is needed to elucidate the mechanisms by which this toxicity occurs. At the same time, it is also imperative to have parallel studies on decorporation agents of actinides to prepare for accidental exposure of humans.146–148
![]() | ||
| Fig. 9 The structures of plutonium(IV)-bound proteins from molecular dynamics (MD) simulations136,137 based on 12-6-4-type Lennard-Jones potentials on Pu4+. The left-hand side of the figure displays Pu4+-bound ferritin, a single L chain of horse ferritin (PDB reference 1AEW) with six Pu4+ ions present. The right-hand sides of the figure displays Pu4+-bound transferrin, a transferrin (PDB reference 1FCK) incorporating two Pu4+ ions. In both figures, protein structures are depicted as green ribbon and Pu4+ ions as red balls. Waters and other ions (Na+, CO32−, etc.) are omitted for clarity. Note that in the case of ferritin, Pu4+ ions are bound to the carboxylic groups of random acidic residues (Asp, Glu) in a non-specific manner, whereas in the case of transferrin two Pu4+ ions are stably accommodated in the iron-binding sites of transferrin with a closed conformation. | ||
The third category of research constitutes an emerging new field of science, which is concerned with the proteins (and enzymes) present in the nature that are capable of accommodating lanthanide or actinide ions. Trivalent lanthanides were long believed to lack any essential biological function. However, in 2011, Kawai and coauthors revealed the catalytic role of trivalent lanthanide ions in the enzyme methanol dehydrogenase (MDH).6,149,150 Later, the growth of the Methylacidiphilum fumariolicum strain SolV, isolated from volcanic mud water in Italy, was found to be strictly dependent on the concentration of Ln3+ ions.151 It was even confirmed that XoxF-type MDH can discriminate amongst different lanthanides.152 In XoxF-type MDH, lanthanide binding to the cofactor pyrroloquinoline quinone (PQQ) was found to transition from a chelate to an unidentate conformation as the lanthanides transition from lighter to heavier elements. This observation was made through the use of molecular dynamics (MD) simulations.118 The enzymatic activity exhibited by XoxF-type MDH, which is exclusive to lighter lanthanides, was identified as the underlying cause of this phenomenon (Fig. 10). These discoveries expanded the significance of Ln3+ ions in biochemistry and in bacterial metabolism. Recent findings have even confirmed that active microbial methanol oxidation in the ocean by MDH is almost entirely Ln-dependent despite the fact that Ln exist only in trace concentrations,153 and that Ln–MDHs are far more broadly distributed than the Ca–MDHs. In the field of biology, it is common to observe examples where the life of organism is dependent on trace elements, and these elements become toxic at higher concentrations. The prevalence of Ln-dependent MDH in natural environments suggests the potential for a novel and hitherto unrecognized role for lanthanide (and potentially actinide as well) in a complexity that is both novel and hitherto unappreciated. In this context, the discovery of actinide-dependent MDH was not unexpected.10 Consequently, the hypothesis that actinides played a certain role in the evolution of primitive metabolic processes (as we discussed in Introduction) may not be entirely dismissed. Despite that actinide-dependent MDH has been confirmed exclusively for heavy actinides Am(III) and Cm(III) in a laboratory setting, given the presence of a high H2 concentration in the Hadean earth's atmosphere (some sources refer to this as large as ∼0.1 bar), the possibility of Pu(III) being present, along with their association with MDH, remains a possibility. Indeed, it is hypothesized that Pu in Oklo (Gabon) from natural reactor in reaction zone 16 existed as Pu(III).154 The isolation and identification of protein lanmodulin7 revealed its capacity to utilize light to heavy actinides, including Ac(III),155 Am(III) and Cm(III).145,156–158 A recent study revealed that a specific type of lanmodulin from Hansschlegelia quercus exhibits sensitivity to the size of Ln3+.159 This finding is analogous to the discovery previously made for Ln-dependent MDH,152 though in the case of MDH the effect of the size of Ln3+ ions appears more critically in their structures by switching the coordination mode of cofactor pyrroloquinoline quinone (PQQ) (Fig. 10). In comparison to its Ca2+-binding counterpart calmodulin, lanmodulin exhibits a significantly shorter sequence, containing only half as many residues in its EF hand loop (12 or 13 for lanmodulin, whereas 24 or 25 for calmodulin). The results of evolutionary studies are awaited to ascertain whether these two proteins are genetically related and whether they have undergone an increase or decrease in size and complexity through processes such as gene duplication and fusion. Such study may provide preliminary evidence that shorter Ln/An-bound proteins may have existed earlier in evolution than their Ca2+-bound counterparts, and may lend further credit to the hypothesis that primitive proteins and enzymes have utilized Ln and/or An. A recently discovered lanthanide-binding protein, Lanpepsy, has been found to possess three to four binding sites for Ln3+ ions.160 This finding provides further evidence for the prevalence of lanthanide-dependent life.
![]() | ||
| Fig. 10 The structures of La3+- and Yb3+-bound methanol dehydrogenase (MDH, XoxF-type) from molecular dynamics simulations.118 On the left- and right-hand sides are La3+- and Yb3+-bound MDH, respectively. The top views show global conformation of the metalloenzyme whereas the bottom views show the central metal ion and its vicinity including cofactor pyrroloquinoline quinone (PQQ). Only in the La3+-bound MDH, PQQ coordinates to metal ion in chelate binding mode thereby allowing nucleophilic attack of the methanol oxygen on PQQ. | ||
The emerging field of actinide-dependent proteins, as exemplified by lanthanide-dependent methanol dehydrogenase and lanmodulin, may supports the notion that lanthanides and actinides were not merely passive participants in early metabolic processes, but could have been actively incorporated into the catalytic role in early life. The discovery of these enzymes challenges the long-held assumption that life is exclusively reliant on a limited set of metal cofactors and opens the possibility that lanthanide (or actinide)-based metalloenzymes may have played a crucial role in the earliest stages of evolution. Further phylogenetic analyses are necessary to confirm that lanthanide (or actinide)-dependent enzymes possess in fact ancestral features and that the calcium-dependent versions evolved later, as discussed elsewhere.161 The observation that some of these proteins exhibit selectivity for different lanthanides and even actinides, and that this selectivity is linked to their ionic radii and coordination preferences, underscores the potential for fine-tuning of metabolic pathways through subtle variations in metal composition. Additionally, there is little knowledge regarding the interactions between tetravalent actinide and proteins, despite the significance of this oxidation state in geologic implications (as exemplified by the formation of An4+ polymers). Addressing this knowledge gap is imperative.
Significant challenges remain in fully evaluating the validity of the nuclear geyser model. While we have presented a compelling case for the chemical feasibility of actinide-driven abiogenesis, direct evidence linking natural nuclear reactors to the origin of life remains elusive. Furthermore, the inherent complexity of prebiotic systems necessitates a multidisciplinary approach that integrates geochemistry, radiochemistry, biochemistry, and computational modelling. Ultimately, resolving the question of whether actinides played a significant role in the origin of life will require a sustained and collaborative effort involving researchers from diverse disciplines. Such investigations will not only provide a deeper understanding of the fundamental processes that gave rise to life on earth, but also a more informed approach to the high-level radioactive waste disposal in the geosphere.
| This journal is © The Royal Society of Chemistry 2026 |