Anton V.
Domnin
a,
Yaroslav V.
Solovev
b,
Denis S.
Syrko
c,
Andrey V.
Golovin
bde,
Robert A.
Evarestov
a and
Mikhail V.
Polynski
*f
aQuantum Chemistry Department, Institute of Chemistry, St Petersburg State University, Universitetsky Prospect 26, Saint Petersburg 198504, Russia
bM.M. Shemyakin and Yu.A. Ovchinnikov Institute of Bioorganic Chemistry of the Russian Academy of Sciences, Miklukho-Maklaya 16/10, Moscow 117997, Russia
cD. Mendeleev University of Chemical Technology of Russia, Miusskaya Ploshchad, 9, building 1, Moscow, 125047, Russia
dFaculty of Bioengineering and Bioinformatics, Lomonosov Moscow State University, Leninskie Gory 1, building 73, Moscow 119234, Russia
eDepartment of Computational Biology, Sirius University of Science and Technology, 354340 Sirius, Russia
fDepartment of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, Singapore 117585, Singapore. E-mail: polynskimikhail@gmail.com; mvp@nus.edu.sg
First published on 17th October 2025
Artificial metallopeptides hold immense potential to combine enzymatic activity with the versatility of organometallic catalysts. However, computational de novo design is largely limited to theozyme models that may neglect second-sphere atomic structure, overlook hydrogen-bonding networks, and ignore metal-induced conformational selection. We overcome these limitations for the case of helical metallopeptides and metal-containing helical motifs by proposing a DFT-based bottom-up methodology applied to the design of Pd-binding (Met-X)n sequences (X = Ala, Val, Ile). Line group symmetry theory is employed to accelerate the calculations by leveraging helical monoperiodicity for computational efficiency. The methodology (a) reproduces the geometric parameters of α-poly-Ala with near-experimental accuracy; (b) to the best of our knowledge, provides the first evidence that the α → π-transition may manifest as a first-order phase transition; (c) identifies (Met-Ala)n π-helices as preferred matrices for canonical Pd(II) Suzuki coupling intermediates. In contrast, Pd incorporation in the α-helical matrix poses significant challenges, as shown by relaxed potential energy scans. From the periodic π-helix, we extract a cluster containing over 250 atoms and model it in aqueous solution at the ωB97X-V/def2-TZVP-gCP//B97-3c level to obtain reliable energetics for the free energy profile of the key oxidative addition step. The profile featured a low activation barrier and exergonic product formation, with reaction energy falling within the optimal window and barriers lower than those reported for bis-phosphine Pd(0) complexes. This methodology offers an efficient strategy for the de novo design of helical peptides and motifs and environmentally benign bioinorganic catalysts, from sequence to the reactivity of the metal center.
However, computational de novo design of ArMs should address challenges arising from both proteomics and organometallic chemistry, many of which remain unresolved. On the proteomics side, uncertain protonation states near the metal site, exchange between low-energy helical conformations that reshape second-sphere contacts, and long-range electrostatics from the periodic hydrogen-bond network all influence structure and reactivity in ways that truncated cluster models systematically miss. On the metal side, coordination-number changes, as well as hemilabile ligands and their possible dissociation, leading to aggregation that competes with catalysis, complicate modeling.30,31 Accurately modeling bond formation and bond cleavage is sometimes beyond the practical domain of conventional force field methods,32 or requires re-parameterization,33 while many machine learning interatomic potentials either lack the corresponding data in their training sets or require fine-tuning,34,35 which motivates the use of quantum chemical methods. Yet even modern semiempirical and tight-binding approaches frequently fail to capture the energetics of relevant chemical reactions.36,37 Consequently, for each system, one must jointly choose an appropriate DFT method and the relevant biomolecular environment. Cluster models of the catalytic center, which are overly restrictive size-wise, can yield misleading results.38 At the same time, models that are too detailed become computationally prohibitive. These constraints motivate frameworks that retain DFT accuracy and generalizability while representing the full biomolecular environment.
In this work, we aim to preserve quantum-chemical accuracy while maintaining computational tractability and without constraining the active-site model to the theozyme. Conventional molecular models are constrained by the theozyme size, whereas a line group theory-based approach can emulate infinite monoperiodic (metallo)polymers and nanostructures with helical symmetry. This methodology has yielded accurate predictions of structures and properties for helical polymers and nanotubes.39,40 It applies to any system with monomers aligned along a helical axis without crystalline periodicity, irrespective of composition.41,42 Section 4 discusses the application of line group theory to monoperiodic systems. By treating the model systems as helical metallopolymers, we model regular peptides and metallopeptides at DFT-level accuracy, circumvent size limitations, and explore extended systems that serve as proxies for ArM scaffolds.
This conceptual model, which treats metal–peptide systems as metal centers embedded in a polymer matrix having a screw axis, is directly relevant to experimentally realized Pd-containing peptide-based and other polymer-based materials. For example, helically chiral Pd-bearing polymer matrices active in asymmetric Suzuki–Miyaura transformations have been reported.43 Site-specific protein–protein conjugates can be formed by modifying a cysteine residue into the S-aryl-Pd-X derivative, which could covalently bind another cysteine residue on the external protein.44 Cooperative catalysis between a lipase active site and a Pd center has been shown to accelerate alkyl–alkyl cross-coupling of 1-bromohexane with B-hexyl-9-BBN in a stable single-atom enzyme–metal complex.45 In addition, DFT calculations predicted that several proteinogenic amino acids can tightly bind both Pd(0) and Pd(II) centers that catalyze Suzuki cross-coupling.46 Taken together, these precedents motivate a quantum chemical, line group theory-based framework for the design of Pd–peptide systems.
Here, we propose a bottom-up methodology for the design of helical metallopeptides and helical metal-containing motifs based on DFT calculations within the line group formalism. First, we demonstrate that a symmetry-based line group formalism can reproduce the geometries and relative energies of canonical helical secondary structures of proteins with near-experimental accuracy. We then design helical peptide matrices that contain methionine-based Pd binding sites and evaluate these Pd–peptide systems at the DFT level. We evaluate structural rearrangements and stability across the Suzuki coupling cycle and assess reactivity in the key oxidative addition step. The framework enables the engineering of chimeric, metal-containing peptide matrices as models of helical (metal-binding motifs of) ArMs, with desired structures, properties, and compositions; it also yields DFT-quality potential-energy surfaces suitable for downstream applications, including training machine-learning interatomic potentials.
![]() | ||
| Fig. 1 Bottom-up design methodology: from the selection of metal-binding residues and spacers to the evaluation of the metallopeptide reactivity. | ||
Since the second step in Fig. 1 involves the design of a peptide sequence that incorporates and stabilizes non-native Pd(0)/Pd(II) centers and primes them for catalysis, we need to optimize intermetal distances between adjacent sites while preventing Pd(0) dissociation, which otherwise promotes rapid aggregation.50 To this end, we insert uncharged spacer residues, alanine (Ala), valine (Val), or isoleucine (Ile), between Met donors to tune steric bulk and backbone hydrophobicity. This arrangement serves three purposes: (i) it spatially separates individual metal centers; (ii) it stabilizes the overall helical architecture; and (iii) it increases the hydrophobicity of the matrix, creating hydrophobic regions that can non-covalently bind organic substrates near the catalytic sites, thereby priming the scaffold for activity. In addition, although these hydrophobic residues do not bind Pd(0)/Pd(II) directly (see below), their side chains provide steric shielding reminiscent of common phosphine and N-heterocyclic carbene ligands.51,52 At the same time, sterically unprotected Gly residues should be avoided due to the undesirable interaction of Pd(II)53–56 with the polypeptide chain.
Step 2 of the methodology (Fig. 1) is explained in detail in section 2.2. Using line-group symmetry, we assemble helical polypeptides from the selected amino acid residues and, within the same formalism, scan electronic energies over a broad range of backbone torsion angles. This enables a systematic mapping of the geometric parameters and relative energies of the α- and π-helical conformations and provides an estimate of the barrier for the α-to-π-helical transition. Section 4 also includes the results of the DFT method selection and shows close agreement between computed and experimental geometrical parameters for the polyalanine α-helix.
Introducing a single spacer residue between Met residues yielded the most favorable helical conformations. In these constructs, the S–S distances in the adjacent Met residues that define each metal-binding site approached those previously reported for bis-Met Pd(0) complexes,46 while the polypeptide backbone remained shielded by bulky nonpolar side chains. Accordingly, we focused subsequent analyses on (Met-X)n polypeptides, where X denotes Ala, Val, or Ile.
Step 3 in Fig. 1 introduces the metallocenters and evaluates their reactivity within the peptide matrix. We track structural rearrangements of the designed Pd–peptide helices across the canonical steps of the Suzuki cross-coupling cycle, and reactivity in the key oxidative addition step. Step 3 completes the computational workflow and is presented in section 2.3.
Summarizing this subsection, we note that the design of helical (metallo)peptides was carried out in light of the MALEK principle, which states Met, Ala, Leu, Glu, and Lys as the strongest helix-forming residues in aqueous media.57 At the same time, for Val and Ile, formation of helical structures in water is unfavorable,58 although they do occur in helices within membrane segments, which are hydrophobic environments.59 Therefore, spacer preselection should be made with the target medium in mind. Therefore, in section 2.2, we focus on (Met-X)n polypeptides, where X denotes Ala, Val, or Ile, to model the intrinsic conformational properties of these polypeptides in the helical state using DFT. At the same time, when introducing Pd and evaluating activity, we primarily use water as the medium (characteristic of artificial metalloenzymes and “green” catalysis settings) and include non-polar THF for comparison. This approach allows us to describe the helical framework using DFT and the SMD implicit-solvent model, which treats bulk electrostatics and includes an implicit H-bond contribution via solvent acidity/basicity descriptors (see section 2.3).
We modeled the geometries of (Met-X)n in α-helical and π-helical conformations. Fig. 2 summarizes structural and energetic differences between (Met-X)n helices in the α- and π-conformations and the parameters that govern the α-to-π-transition. For (Met-Ala)n, the transition from the more densely packed α-helix to the less dense π helix increases the inner radius from 2.31 to 2.84 Å. Concomitantly, the number of hydrogen bonds per helical pitch rises from 3.60 to 4.45. The relationship between h and Qaa shows that, as Qaa increases during the transition, h first increases to about 5.6 Å, then abruptly decreases to about 5.0 Å near Qaa ≈ 4.1. After the formation of new hydrogen bonds characteristic of the π-helix, the pitch increases again. Notably, the h vs. Qaa trends are only weakly sequence-dependent across X = Ala, Val, and Ile.
We now analyze the general energetics of the α-to-π-helix transition. A comparison of the top and bottom panels in Fig. 2a shows that the positions of the energy minima do not coincide with the minima in helical pitch. For the α-helix, the energy minimum occurs at h approximately 5.35 Å rather than at the smallest pitch near 5.1 Å. For the π-helix, the energy minimum is at h approximately equal to 5.1 Å, instead of being at the lowest observed pitch of about 5.0 Å. Across all α- and π-helix pairs, the relative energies, ΔE, differ by less than 2.5 kJ mol−1.
While all Met-containing polypeptides considered here undergo the α-to-π-helix transition at Qaa = 4.18, the corresponding activation barriers span from 20 kJ mol−1 for poly(Met-Ile)n to 24 kJ mol−1 for poly(Met-Ala)n (Fig. 2a). This 4 kJ mol−1 spread reflects an unexpected steric acceleration that becomes more pronounced as the helix untwists along the α to π pathway. The transition follows an axial elongation pathway that redistributes the H-bond network; the screw motion intrinsic to the α-to-π conversion entails a concerted sliding of the backbone that positions N–H donors in the helical H-bond network between two C
O acceptors. At the PBE/def2-TZVP level, the associated imaginary frequencies are 27.1i, 22.1i, and 19.6i cm−1 for poly(Met-Ala)n, poly(Met-Val)n, and poly(Met-Ile)n, respectively, based on analysis of the corresponding local H-bond clusters. The supporting ZIP archive includes the coordinates of the PES saddle points corresponding to the N–H donor slides in the local H-bond clusters, with the corresponding animated imaginary modes, together with visualizations of the collective screw-expansion motion.
The computed energy profile displays two basins of comparable depth, while the helical pitch changes abruptly from about 5.5 Å in the α basin to about 5.0 Å in the π basin (Fig. 2a, bottom panel). This discontinuity is accompanied by a notable reorganization of backbone hydrogen bonds (Fig. 2b) and a measurable widening of the helical core. Structural changes of this magnitude cannot arise from infinitesimal fluctuations about a single minimum. In Landau's phenomenological theory,61 second-order transitions feature a continuous order parameter, whereas first-order transitions exhibit discontinuous jumps in the order parameter, coexisting phases with distinct symmetries, and an energy barrier between them. In this particular case, α-to-π interconversion demonstrates all these characteristics, providing a rationale to see it as a first-order phase transition.
In our formalism, the chains are infinite, so the thermodynamic limit is effectively reached. In this regime, the system could exhibit a latent enthalpy and may show metastability and hysteresis on molecular dynamics timescales, consistent with viewing the α- and π-helices as alternative phases of the same polymer. A more complete group-theoretic analysis that constructs an explicit path in order-parameter space and mathematically devises the relevant symmetry-breaking invariant is suggested in the future, as well as a more complete consideration of lateral packing present in real crystalline systems, for a rigorous assignment of the α-to-π-transformation as a first-order phase transition.
| Polymer/parameter | Poly(Met-Ala) | Poly(Met-Val) | Poly(Met-Ile) | |||
|---|---|---|---|---|---|---|
| Pd-free | With Pd | Pd-free | With Pd | Pd-free | With Pd | |
| Pd atoms are shown as cerulean spheres. Peptide residues are shown as sticks (Met in grey, X in green). S atoms in Met residues are orange. The helical backbone structure is superimposed. | ||||||
| α-Helical conformation | ||||||
| Q aa | 3.61 | 3.63 | 3.61 | 3.61 | 3.64 | 3.63 |
| h, Å | 5.37 | 5.39 | 5.40 | 5.41 | 5.39 | 5.42 |
| f, Å | 1.49 | 1.49 | 1.49 | 1.50 | 1.48 | 1.50 |
| π-Helical conformation | ||||||
| Q aa | 4.43 | 4.50 | 4.42 | 4.50 | 4.41 | 4.52 |
| h, Å | 5.11 | 5.18 | 5.13 | 5.22 | 5.11 | 5.23 |
| f, Å | 1.15 | 1.15 | 1.16 | 1.16 | 1.16 | 1.16 |
Although all three poly(Met-X)n Pd-free polymers adopt very similar helical structures, poly(Met-Ala)n is easier to prepare under standard peptide synthesis protocols,62,63 exhibits slightly higher structural stability across the α-to-π-transformation (hence slightly more controllable dynamics), and contains fewer atoms, which improves computational tractability. At the same time, selecting sterically unprotected Gly is inadvisable (see section 2.1). Importantly, in the reactivity evaluations below, we target water media. As noted in section 2.1, Met and Ala have a strong propensity to form helical structures in water, whereas helices containing Val and Ile are more typical of hydrophobic (membrane-like) environments. We therefore selected the [Pd(Met-Ala)2]n polymer for further investigation.
We next introduced Pd(II) centers that mimic key intermediates in the Suzuki coupling of PhBr with Ph- and vinyl-boronic esters into the peptide matrices. For both α- and π-helical forms, we optimized the geometries and performed relaxed potential energy surface (PES) scans to probe structural rearrangements along the catalytic cycle. Results for the π- and α-helical structures are summarized in Table 2. The π helix is structurally conservative across all intermediates considered, with Qaa, h, and f changing only marginally. In contrast, the α-helical [Pd(Met-Ala)2]n chain shows a larger response, especially in the case of [PhPdBr(Met-Ala)2]n, where both Qaa and h decrease relative to the other intermediates, consistent with local tightening within each helical pitch and the buildup of structural strain.
| Parameter | [Pd(Met-Ala)2]n | [PhPdBr(Met-Ala)2]n | [PhPd(OH)(Met-Ala)2]n | [Ph2Pd(Met-Ala)2]n | [PhVinPd(Met-Ala)2]n |
|---|---|---|---|---|---|
| π-Helical conformation | |||||
| Q aa | 4.50 | 4.46 | 4.48 | 4.46 | 4.47 |
| h, Å | 5.18 | 5.17 | 5.18 | 5.17 | 5.18 |
| f, Å | 1.15 | 1.16 | 1.16 | 1.16 | 1.16 |
| α-Helical conformation | |||||
| Q aa | 3.63 | 3.59 | 3.65 | 3.66 | 3.62 |
| h, Å | 5.39 | 5.31 | 5.40 | 5.42 | 5.39 |
| f, Å | 1.49 | 1.48 | 1.48 | 1.48 | 1.49 |
Across all Pd(II)–peptide polymers studied, the α-helical forms are significantly less favorable than their π-helical counterparts (Table 2 and Fig. S1). In the case of [Ph2Pd(Met-Ala)2]n, the relative energy of the α-helical chain, ΔE, is +88.5 kJ mol−1 per monomer. The α-helical conformations of [PhPdBr(Met-Ala)2]n and [PhPd(OH)(Met-Ala)2]n are +55.1 and +67.8 kJ mol−1 per monomer, relative to the corresponding π-helical conformations. We therefore conclude that the π-helical form is universally preferred for incorporating canonical Pd(II) intermediates of the Suzuki coupling, whereas the α-helical alternative is strongly energetically disfavored.
To estimate the potential of the [Pd(Met-Ala)2]n polymer in C–C coupling reactions, we analyzed the key oxidative addition step. Along a relaxed PES scan in the case of the periodic, less favorable α-helical matrix, we found that dissociation of one Met ligand from the [Pd(Met)2] center (state 1 in Fig. 3a) was required to form the Pd(0)–aryl halide π-complex, [Pd(Met)(η2-PhBr)] (1′). Formation of 1′ is strongly endothermic, ΔE = +109 kJ mol−1, relative to 1. The barrier from 1′ to the oxidative addition transition state TS1 is small, ΔE‡ = 17 kJ mol−1; however, referenced to 1, it amounts to 126 kJ mol−1 because forming 1′ is highly endothermic. The subsequent formation of the α-helical product, [PhPdBr(Met-Ala)2]n (intermediate 2), is exothermic relative to 1′ (ΔE = −84.7 kJ mol−1) while being endothermic relative to 1 (24.3 kJ mol−1).
In summary, whereas the Pd-free polypeptides display two nearly isoenergetic minima separated by barriers below approximately 24.5 kJ mol−1 (Fig. 2), Pd(II) incorporation constrains the helix in the π-conformation (Table 2). The corresponding oxidative addition requires energetically unfavorable Met dissociation, also being endothermic, likely due to strain. These considerations motivate our preferential focus on the more favorable π-helical scaffold below.
To obtain the energetic parameters of the oxidative addition in the π-helical matrix within model water and THF solvents at a higher level of theory, we proceeded as follows. A chain containing four [PhPdBr(Met-Ala)2]n units was extracted from the periodic structure, with the terminal atoms capped with Me substituents. One of the inner Pd centers was selected as the reactive center (Fig. 3b). Geometry optimizations of the pre- and post-oxidative addition intermediates, as well as the transition state, were performed at the PBE/def2-SVP (H: def2-TZVP) level. All atoms were constrained except those of the reactive center and the aliphatic side chains of the corresponding two Met residues. The resulting Hessian in the TS system had the target imaginary mode corresponding to C–Br bond cleavage, with frequency 106.15i cm−1. TS coordinates for both the extended helix and the cluster models, along with the mode visualization, are provided in the supporting ZIP archive. Then, we performed single-point energy refinements at the ωB97X-V/def2-TZVP-gCP with SMD solvation (water and THF) and combined these with thermochemical corrections computed at the PBE/def2-SVP (H: def2-TZVP) level to estimate Gibbs free energies of activation and of elementary steps. ωB97X-V was selected because it ranks the best among non-double-hybrid functionals on the comprehensive GMTKN55 benchmark (including thermochemistry, barrier heights, and noncovalent interactions) and has been recommended for metalloenzyme reactions as one of the best options.64,65
Fig. 3b summarizes the energetic parameters and optimized geometries for the oxidative addition of PhBr to the [Pd(Met)2] center in the π-helical matrix. Formation of the pre-reaction intermediate from the metallopeptide and PhBr (state 1) yields 1′ with a bridging Br ligand and lies at ΔG = 0.0 and +5.2 kJ mol−1 relative to 1 in water and THF, correspondingly. From 1, oxidative addition proceeds over a relatively low barrier of 64.5 and 73.7 kJ mol−1 in water and THF, respectively. The values reported for typical bisligated phosphine Pd(0) complexes vary from 104 to 122 kJ mol−1.66 We took the same bisligated complexes (L = PPh3, PCy3, and SPhos; P(t-Bu)3 did not form the bisligated state) and computed the activation barrier in 1 → TS1 at the ωB97X-V/def2-TZVP-gCP//B97-3c level with SMD solvation. The barriers ranged from 82 to 122 kJ mol−1 in water and from 97 to 136 kJ mol−1 in THF, correspondingly; in all the cases, the PPh3-ligated complex corresponded to the lowest barrier (all values and structures are provided in the supporting ZIP archive). Finally, the formation of the [PhPdBr(Met)2] center is strongly exergonic, with ΔG = −118.7 and −111.4 kJ mol−1 in water and THF, respectively, relative to 1. Taken together, these results indicate that the π-helical matrix can accommodate Pd(0)/Pd(II) centers and facilitate the key oxidative addition step relevant to Suzuki and many other cross-coupling reactions.
This approach not only facilitates accurate non-empirical modeling of infinite (metallo)peptide systems by exploiting helical symmetry (thereby overcoming the computational limitations of large molecular models, when considered at the DFT level) but also provides a robust theoretical foundation for characterizing structural transitions such as the α-to-π helix transformation. This transition—manifesting itself in the reformation of hydrogen-bond networks and backbone rearrangements—exhibited a 20–24 kJ mol−1 energy barrier and an abrupt helical pitch change (from ∼5.5 Å to ∼5.0 Å), suggesting that such a transformation could manifest itself as a first-order phase transition.
Our results demonstrate the efficacy of this approach in designing [Pd-(Met-X)2]n metallopeptides (X = Ala, Val, Ile). We designed the spatial structures and modeled potential energy surfaces near the minima corresponding to the canonical intermediates in the Suzuki cross-coupling reaction, identifying stable helical conformations with Met residue arrangement favoring Pd(0) and Pd(II) binding. Incorporation of Pd(0) and Pd(II) centers revealed that π-helical conformations are energetically more favorable for stabilizing models of the canonical intermediates, especially [PhPdBr(Met-Ala)2]. In contrast, α-helical structures exhibited strain, requiring energetically costly dissociation of one Met residue for PhBr oxidative addition to proceed, highlighting the superior properties of Met-containing π-helices for catalysis involving the [Pd(Met)2] center and indicating potential applicability to other helical metallopeptide systems incorporating “softer” transition metals. These findings highlight the critical role of helical conformation in modulating catalytic activity, offering actionable insights for tailoring metallopeptide sequences.
To obtain the kinetic and thermodynamic parameters of PhBr oxidative addition in an aqueous (or THF) medium using a high-accuracy DFT protocol, we generated a cluster model directly extracted from the DFT-optimized periodic π-helix, avoiding a simplistic theozyme model. Indeed, by construction, the cluster inherits second-sphere packing, backbone strain, and hydrophobic shielding from the parent periodic π-helix. The (free) energy profile showed a low barrier of PhBr oxidative addition, well below barriers typical for bis-phosphine Pd(0) complexes.
Such metallopeptides serve as proxies for artificial metalloenzymes due to their ability to mimic the coordination environments and catalytic functionality of natural enzymes while incorporating non-biological metals like Pd for synthetic transformations. Beyond their utility as enzyme mimics, such metal–peptide systems are promising because of their hybrid nature, blending the structural versatility of polypeptides and inherent biodegradability of the peptide matrix with the reactivity of organometallic complexes. The pursuit of such designed Pd–peptide assemblies reflects the growing interest in experimental studies. Metal–peptide frameworks, such as metal-α-helix systems, demonstrate that porous frameworks with characteristic metal sites can be constructed from short peptides.67 At the same time, the functional potential of non-helical Pd–peptides has been validated in a biological context: a Pd(II)-stapled β-sheet miniprotein has been shown to catalyze bioorthogonal depropargylation reactions inside living cells.68 Pd-containing helical structures have also drawn interest beyond biosystems: the helically chiral polymer ligand PQXphos afforded axially chiral biaryl esters in asymmetric Suzuki coupling with high enantioselectivities, and its helical chirality can be switched by solvent to access either enantiomer.43 Collectively, these works motivate exploration of new structural frameworks, such as the Pd-Met-based helices proposed here. By providing a predictive framework for designing metallopeptides and metallo-helical polymers with tailored reactivity, our work facilitates the engineering of bioinspired catalysts and addresses the pressing need for environmentally benign yet efficient and selective catalytic systems.
The canonical structures of α and π helices are described by a screw axis. This corresponds to a cyclic group with generator Z = (CQ|f), where Q is the order of the screw axis and f denotes the magnitude of the shift along the screw axis (with the condition that Q ≥ 1). Technically, Q may either be a rational or an irrational number. As the CRYSTAL17 software used for quantum chemical modeling requires translational periodicity, we will consider only the case in which Q is rational. Only symmetry groups with rational Q include a translation group as a subgroup; they can be expressed within crystallographic notation (q, p, t) and polymer notation (q, r, f).
A rational value of Q can be expressed as the quotient of two co-prime positive integers: Q = q/r. In this case:
| (CQ|f)q = (CQq|q·f) = (E|t), | (1a) |
| qf = t, | (1b) |
![]() | (2) |
This implies that before reaching a translationally equivalent position, the helix completes r full turns, while the monomer repeats q times. Consequently, Q = q/r is the number of monomers per turn, r is the number of turns, q is the total number of monomers in the structure. Thus, the height (pitch) of a single turn is expressed as:
![]() | (3) |
These parameters under consideration are presented schematically in Fig. 4. Finally, in our case, a monomer in the chain consists of a single amino acid. Therefore, Q corresponds to the number of amino acid residues per turn (Qaa).
![]() | ||
| Fig. 4 Schematic representation of a helical structure, along with its characteristic structural parameters described in eqn (1b), (2), and (3). Monomers are represented by spheres; in (a), the blue spheres indicate monomers belonging to the first turn of the helix; (b) shows rotation angle φ between monomers connected via helical transformation, with spheres indicating distinct positions. | ||
Given the values of q and r, the parameter p can be determined using the following formula:
rp = lq + 1 or rp = 1 mod q, | (4) |
Helix transitions between α and π conformations were studied by sequentially optimizing structures across a range of helical axis orders (Q). For each structure, the symmetry-irreducible unit of the preceding monomer and a shift parameter (f) were used to generate the next structure. Geometry optimization was then performed while fixing Q. The helical axis order Q was converted to Qaa (amino acid residues per turn) using standard helical parameter equations (see SI for details). Structures were optimized for Q∈(1.9, 2.1), which corresponds to φ∈(172°, 189°) and Qaa∈(3.8, 4.2).
Additional calculations on cluster models were performed using ORCA 5.0.3.81 Geometry optimizations were conducted using the PBE functional, with the def2-TZVP basis set for hydrogen atoms and the def2-SVP basis set for all other elements.82 The def2/J auxiliary basis set83 was employed to enable the RI approximation.84 A TightSCF convergence criterion and the DefGrid2 integration grid were used. The same settings were applied for vibrational frequency calculations to obtain thermochemical corrections.
Spin-restricted single-point Hessian calculations to locate saddle points for H-bond sliding were performed at the PBE/def2-TZVP level, using the def2/J auxiliary basis set for the RI approximation. A VeryTightSCF convergence criterion and the DefGrid3 integration grid were employed to ensure high accuracy.
Single-point energy refinements were performed with the ωB97X-V functional,85 using the resolution-of-identity with chain-of-spheres exchange approximation,86 paired with the def2-TZVP basis set and the def2/J auxiliary basis. A TightSCF convergence criterion was applied with a DefGrid2 integration grid to balance accuracy and computational cost.
For solvation effects, single-point energy calculations were conducted with the SMD model using the B3LYP functional,87–89 LANL2DZ for Pd and Br,90,91 and 6-31G* for other elements,92,93 following the recommended protocol of B3LYP/6-31G* in the original publication94 as closely as possible. TightSCF with VerySlowConv and DefGrid2 was enforced to ensure numerical stability. Thermochemical corrections were computed in ORCA using the default settings. A standard-state correction of 7.93 kJ mol−1 was added to the thermochemical corrections to convert from the 1 atm ideal gas reference to the 1 M solution.
| This journal is © the Partner Organisations 2026 |