Laura
Burchill
,
Arashdeep
Kaur
,
Artur
Nastasovici
,
Mihwa
Lee
* and
Spencer J.
Williams
*
School of Chemistry, Bio21 Molecular Science and Biotechnology Institute, University of Melbourne, Parkville, Victoria 3010, Australia. E-mail: mihwa.lee@unimelb.edu.au; sjwill@unimelb.edu.au
First published on 9th September 2024
2,3-Dihydroxypropanesulfonate (DHPS) and sulfolactate (SL) are environmentally important organosulfur compounds that play key roles as metabolic currencies in the sulfur cycle. Despite their prevalence, the pathways governing DHPS and SL production remain poorly understood. Here, we study DHPS-3-dehydrogenase from Cupriavidus pinatubonensis (CpHpsN), a bacterium capable of utilizing DHPS as a sole carbon source. Kinetic analysis of CpHpsN reveals a strict preference for R-DHPS, catalyzing its 4-electron oxidation to R-SL, with high specificity for NAD+ over NADP+. The 3D structure of CpHpsN in complex with Zn2+, NADH and R-SL, elucidated through X-ray crystallography, reveals a fold akin to bacterial and plant histidinol dehydrogenases with similar coordination geometry around the octahedral Zn2+ centre and involving the sulfonate group as a ligand. A key residue, His126, distinguishes DHPS dehydrogenases from histidinol dehydrogenases, by structural recognition of the sulfonate substrate of the former. Site-directed mutagenesis pinpoints Glu318, His319, and Asp352 as active-site residues important for the catalytic activity of CpHpsN. Taxonomic and pathway distribution analysis reveals the prevalence of HpsN homologues within different pathways of DHPS catabolism and across bacterial classes including Alpha-, Beta-, Gamma-, and Deltaproteobacteria and Desulfobacteria, emphasizing its importance in the biogeochemical sulfur cycle.
Bacterial sulfoglycolysis of SQ forms S-SL and S-DHPS (Fig. 1).11–15 The individual enantiomers of DHPS are produced by various marine diatoms and coccolithophores, with some coccolithophores producing both molecular antipodes.7 A pathway for the interconversion of DHPS enantiomers in Cupriavidus pinatubonensis has been proposed involving a two-component system of HpsO and HpsP, which are NAD(P)+-dependent DHPS-2-dehydrogenases.16 Oxidation of DHPS by NAD+-dependent DHPS-3-dehydrogenase HpsN gives SL.16 A related pathway has been proposed in Roseobacter pomeroyi.7 The enantiomers of SL can be interconverted by the NAD(P)+-dependent SL-2-dehydrogenases SlcC and ComC, via sulfopyruvate.17 SL is a substrate for various metabolic pathways, as shown in Fig. 2. One pathway involves the Fe2+-dependent SL lyase SuyAB, which catalyzes the elimination of sulfite from SL to give pyruvate.16–18 A second pathway involves the oxidation of SL to sulfopyruvate (catalyzed by SlcC or ComC), which allows for decarboxylation, catalyzed by ComDE, to give sulfoacetaldehyde. Sulfoacetaldehyde is subsequently converted to acetyl phosphate and sulfite through the catalytic action of thiamine diphosphate (ThDP)-dependent sulfoacetaldehyde acetyltransferase Xsc.19 A third pathway entails the reductive amination of sulfopyruvate with glutarate, affording L-cysteate. CuyB is a racemase that interconverts L-cysteate with D-cysteate, with the latter being the preferred substrate for the pyridoxal 5′-phosphate (PLP)-dependent CuyA, leading to the formation of pyruvate, ammonia and sulfite.19,20
All of the above sulfolysis pathways feature DHPS dehydrogenase HpsN. This enzyme was originally identified in Cupriavidus pinatubonensis JMP134, a bacterium that can grow on DHPS as sole carbon source.16 HpsN, when purified and studied was a homodimer that converted racemic DHPS to SL using NAD+ as a cofactor. HpsN was predicted to act specifically on R-DHPS to generate R-SL,16 with recent experimental evidence confirming this prediction.7 In C. pinatubonensis, the gene responsible for HpsN lies within the hpsRNOUPsuyAB gene cluster. This cluster also encodes transcriptional regulator HpsR, DHPS-2-oxidoreductases HpsO and HpsP, a major facilitator superfamily uptake system HpsU, and SL lyase SuyAB. Notably, HpsN is sequence- and structurally-related7 to histidinol dehydrogenase (HisD), a Zn2+ and NAD+ dependent enzyme that oxidizes L-histidinol to histidine. While HisD has been extensively investigated,21–23 very little is currently known about the molecular mechanisms underlying catalysis by DHPS-3-dehydrogenase HpsN.
In this study, we present kinetic and structural analyses of DHPS-3-dehydrogenase HpsN from C. pinatubonensis. We conducted Michaelis–Menten kinetics on individual enantiomers of DHPS, revealing a marked preference for R-DHPS, and high specificity for NAD+versus NADP+. We show that HpsN catalyzes oxidation of R-DHPS to afford SL, and that SLA is also a substrate for the enzyme. Notably, HpsN is highly selective for sulfonated substrates and does not exhibit any activity towards L-histidinol or the structural analogue glycerol-1-phosphate. Furthermore, we report the 3D structure of HpsN, determined by X-ray crystallography, in complex with NADH and R-SL. The structure defines the coordination environment about the Zn2+ centre, pinpoints a key residue involved in recognizing the sulfonate group, and identifies possible catalytic residues, which we investigate by site-directed mutagenesis. Lastly, we explore the taxonomic distribution of DHPS-3-dehydrogenases within DHPS degradation pathways, shedding light on the ecological distribution of DHPS metabolic pathways.
We initially assessed whether CpHpsN could catalyze the oxidation of the two enantiomers of DHPS. Reaction mixtures of CpHpsN, NAD+ and R- or S-DHPS were analyzed by liquid chromatography-mass spectrometry (LCMS) with a triple quadrupole (QqQ) mass spectrometer in product ion mode. The SL peak formed from R-DHPS was much larger than that formed from S-DHPS, indicating a clear preference of CpHpsN for R-DHPS (Fig. 3a). As HpsN catalyzes a four-electron oxidation, we tested whether the proposed intermediate, sulfolactaldehyde (SLA) could serve as a substrate for CpHpsN. LCMS analysis demonstrated formation of a substantial peak for SL, suggesting that this is an even better substrate for this enzyme (Fig. 3a).
While it is well-established that histidinol dehydrogenases are Zn2+-dependent metalloenzymes, the metal dependency of HpsN has not been comprehensively studied. In the presence of 1 mM EDTA, the activity of Zn2+-loaded CpHpsN was reduced 250000-fold. We then dialysed the EDTA-treated CpHpsN, to obtain demetallated CpHpsN, and added various divalent metals to study the reconstitution of activity using [NAD+] = 0.3 mM and [R-DHPS] = 8.0 mM. As mentioned above, demetallated HpsN lost essentially all activity, establishing it as a metalloenzyme (Fig. S2†). All divalent metals led to at least a partial recovery of activity. Maximum activity was observed with Co2+, followed closely by Zn2+ > Mn2+ > Mg2+ ≈ Ca2+ ≈ Ba2+ ≈ Cu2+ > Ni2+. Based on the close relationship with Zn2+-dependent histidinol dehydrogenases, the high intracellular concentration of Zn2+, and the rarity of cobalt-metalloenzymes, we continued to study the Zn2+ form of CpHpsN.
Entry | Variable substrate | Constant substrate | Concentration (mM) | k appcat (s−1) | K appM (mM) | (kcat/KM)app (mM−1 s−1) |
---|---|---|---|---|---|---|
a ND, no activity detected. b Saturation not achieved. | ||||||
1 | R-DHPS | NAD+ | 0.30 | 0.97 ± 0.03 | 1.3 ± 0.15 | 0.75 ± 0.20 |
2 | S-DHPS | NAD+ | 0.30 | — | — | 7.7 × 10−7b |
3 | NAD+ | R-DHPS | 8.00 | 1.60 ± 0.11 | 0.47 ± 0.10 | 3.4 ± 1.10 |
4 | R-SLA | NAD+ | 0.30 | 1.58 ± 0.05 | 0.36 ± 0.06 | 8.6 ± 0.83 |
5 | NAD+ | S-DHPS | 8.00 | — | — | 0.048b |
6 | R-DHPS | NADP+ | 0.30 | ND | ND | ND |
7 | Glycerol phosphate | NAD+ | 0.30 | ND | ND | ND |
8 | L-Histidinol | NAD+ | 0.30 | ND | ND | ND |
Our HPLC analysis indicated that SLA is a more favorable substrate than R-DHPS. Since our synthetic SLA is racemic, we initially examined its consumption. The incubation of a solution of racemic SLA (0.3 mM) with CpHpsN and excess NAD+ (8 mM) gave a progress curve that suggested complete reaction after 2 h (Fig. S4†). Addition of more CpHpsN did not result in further reaction. Using the extinction coefficient for NAD+, we calculate that 48 ± 1% of the SLA was consumed. This finding implies that CpHpsN is stereospecific for R-SLA, and thus for kinetic analysis we adjusted the concentration for only this stereoisomer (i.e. [SLA]/2). At 0.3 mM NAD+, the pseudo first-order parameters for R-SLA were: kappcat = 1.58 s−1, KappM = 0.36 mM and (kcat/KM)app = 8.6 mM−1 s−1. Therefore, R-SLA is approximately 12-fold more efficient as a substrate for CpHpsN than R-DHPS in terms of (kcat/KM)app value, mainly caused by a 3.6-fold lower KappM value. This should be considered a lower estimate of the greater efficiency of R-SLA, as the enantiomer S-SLA may act as a competitive inhibitor.
Next, we investigated if CpHpsN has activity on other non-sulfonated substrates. Glycerol-1-phosphate, which is structurally related to DHPS, is produced through the reduction of dihydroxyacetone phosphate or glycerol phosphorylation. No enzymatic activity was detected when glycerol phosphate was incubated with CpHpsN and NAD+. Similarly, no activity was observed when L-histidinol was incubated with CpHpsN and NAD+. Based on structural analogy with R-SL, we examined whether the 2-amino substituted analogue L-cysteic acid (R-cysteic acid) was an inhibitor of CpHpsN. At [R-DHPS] = 1.0 mM and [NAD+] = 0.3 mM, L-cysteate inhibited CpHpsN with IC50 = 2.4 mM (Fig. S5†).
CpHpsN forms a domain-swapped tight dimer with 28% of surface area buried at the dimer interface, with the interface featuring extensive hydrophobic and hydrogen-bond interactions (Fig. 4a). Each protomer consists of four distinct domains; domains 1 and 2 are globular and contain the active sites while domain 3 is engaged in dimerization (Fig. 4b and S6†). Domain 4 lies at the C-terminus and contributes to the active site and is also involved in dimerization through a metal binding site. Coordination of Zn2+ involves three residues from one monomer and one residue (His411*; * denotes the other monomer in the dimer) from the other monomer (Fig. 5 and S7†), showing that the enzyme forms an obligate dimer. Zn2+ adopts an octahedral coordination geometry in CpHpsN. While most Zn metalloenzymes feature tetrahedral Zn coordination, with direct participation of Zn2+ in catalysis by activating a water molecule, it has been shown that within various plant and bacterial histidinol dehydrogenases, Zn2+ adopts an octahedral geometry.23,24
![]() | ||
Fig. 5 Comparison of Zn-coordinating residues and proposed catalytic residues between CpHpsN and MtHisD. (a)–(c) Zn centres in CpHpsN·Zn2+ (a), CpHpsN·Zn2+·NADH·R-SL (b), and MtHisD·Zn2+·NAD+·L-histidine (PDB entry 5VLD) (c) with schematic coordination geometry shown below. Carbon atoms in the residues from chain A are colored in light pink (CpHpsN) and light cyan (MtHisD) while those from chain B are in light grey. Carbon atoms in the product (R-SL and histidine) are shown in green and Zn2+ shown in grey spheres. (d) and (e) Weblogo diagrams (CpHpsN (d) and MtHisD (e)) showing unique sequence motifs identified herein, and unique zinc- and ligand-binding histidine residue (red *) for HpsN. |
Comparison of the 3D structures of the CpHpsN·Zn2+ and CpHpsN·Zn2+·NADH·R-SL complexes reveals that binding of NADH and R-SL causes a significant movement and closure of domains 1 and 2 (Fig. 4c and S8a†). The closure of the two domains results in a decrease in the distance from the tip (Arg153) of domain 1 to the tip (Thr284) of domain 2 from 21.5 Å to 13.5 Å (Fig. S9†). Analysis of the interfaces using the Protein Interfaces, Surfaces and Assemblies interactive tool25 indicates that the CpHpsN·Zn2+·NADH·R-SL complex forms a more compact dimer versus CpHpsN·Zn2+, with the increased buried surface area from 9620 to 13250 Å2 and decrease in ΔGdiss from 84.1 to 98.7 kcal mol−1. The root-mean square deviation of the CpHpsN·Zn2+ and CpHpsN·Zn2+·NADH·R-SL structures is 2.0 Å over 407 common Cα positions. A further difference in the two structures is a change in positions of His126 and Asp352, as well as in the flexible loops in which these residues are located (Fig. 4c). This will be discussed in more detail below.
A structure-based search using Foldseek26 with the ‘open’ CpHpsN·Zn2+ structure as query identified E. coli histidinol dehydrogenase (EcHisD, PDB 1KAR),23 in complex with Zn2+ and histamine (a substrate/product analogue), as the closest structural homologue (sequence identity 29.7%, E-value = 2.6 × 10−32) (Fig. S8b†). However, using the ‘closed’ CpHpsN·Zn2+·NADH·R-SL structure as query, the search yielded as the top ranked target Medicago truncatula histidinol dehydrogenase (MtHisD, PDB 5VLD),24 in complex with Zn2+, NAD+ and L-histidine (sequence identity of 33.6%; E-value = 4.5 × 10−38) (Fig. S8c†). Thus, when cofactor and product are bound CpHpsN adopts the same ‘closed’ conformation as MtHisD, while when crystals were grown without ligands, the resulting crystal structure adopts the same ‘open’ conformation as EcHisD.
In contrast, the crystal structure of CpHpsN·Zn2+·NAD+, obtained by soaking crystals of CpHpsN·Zn2+ in the open conformation with NAD+, represents a nonproductive complex where the electron density for the nicotinamide group is disordered in the crystal structure and therefore, modelled with zero occupancy (Fig. S9a†). The distance between the projected position of the nicotinamide and the Zn center is beyond the expected distance for a hydride transfer from the substrate. Therefore, the closure of the domains 1 and 2 are essential for productive complex formation.
In addition to CpHpsN·Zn2+·NADH·R-SL complex, we also obtained crystals of the CpHpsN·Zn2+·NADH·L-cysteate complex that diffracted to 1.75 Å using a similar co-crystallization approach (Fig. S10c†). The overall structure and the active site coordination of the cysteate complex is essentially identical to that of the CpHpsN·Zn2+·NADH·R-SL structure, demonstrating that this inhibitor functions through mimicry of R-SL.
A notable difference between the HpsN and HisD structures is the absence of an equivalent histidine residue to His126 of CpHpsN in HisD. This histidine residue is conserved within HpsN homologues, but not within HisD homologues, where it is predominantly Leu. Therefore, the conformational changes about the zinc center are not observed in HisD structures. Multiple sequence alignments reveals an HpsN-specific motif that distinguishes homologues of HisD: G-S-A-P-L, from homologues of HpsN: G-R-Y-A/S–H. The final His in the HpsN-specific motif (His126 in CpHpsN) is the axial Zn2+-coordinating residue observed in the CpHpsN complex lacking R-SL that instead binds the sulfonate in the CpHpsN·Zn2+·NADH·R-SL structure (Fig. 5d and e). Possibly, this residue may assist in maintaining a high affinity for Zn2+ in the absence of substrate/product, and for discrimination of R-SL versusL-histidinol.
To probe the roles of Glu318, His319 and Asp352, we conducted site-directed mutagenesis to convert each residue independently to Ala. The E318A and H319A mutants of CpHpsN suffered 580-fold and 240-fold reductions in (kcat/KM)app values, respectively, which was mainly due to a reduction in kappcat values (Fig. S11† and Table 2). No activity was detected for the D352A mutant of CpHpsN, most likely due to the inability to form a productive zinc-coordination complex with substrate. The 3D structure of the CpHpsN D352A mutant in complex with Zn2+ and NAD+ was solved and refined to 2.23 Å resolution and is essentially identical to the CpHpsN·Zn2+·NADH structure in the ‘open’ conformation (Fig. S12a–c†). In contrast, the structure of CpHpsN H319A in complex with R-DHPS and NADH (refined to 2.01 Å resolution) adopts the closed conformation and is similar to the CpHpsN·Zn2+·NADH·R-SL structure in that the sulfonate and secondary hydroxyl groups participate in the octahedral zinc coordination (Fig. S12d and e†). The 3D structure of the ‘closed’ CpHpsN H319A·Zn2+·NADH·R-DHPS complex displays a productive geometry with the primary hydroxyl carbon of DHPS positioned 3.3 Å away from C4 of the nicotinamide group. In this complex, the Zn and R-DHPS sites are not fully occupied, and the final coordinates are modelled with the occupancy of 0.7, whereas the cofactor NADH has a full occupancy. The partial occupancy of zinc presumably arises due to incomplete reconstitution during purification. However, the concentration of R-DHPS used during crystallization was in excess (>100×) and is not a limiting factor. This observation therefore suggests that zinc is essential for ligand binding in the active site. Collectively, these data provide evidence for a role for Asp352 in zinc coordination to form a catalytically productive complex, and for Glu318 and His319 in the catalytic mechanism of CpHpsN.
Mutation | Variable substrate | Constant substrate | Concentration (mM) | k appcat (s−1) | K appM (mM) | (kcat/KM)app (mM−1 s−1) |
---|---|---|---|---|---|---|
a ND, no activity detected. | ||||||
E318A | R-DHPS | NAD+ | 0.30 | 5.1 × 10−3 ± 0.01 | 2.8 ± 0.3 | 1.8 × 10−3 ± 0.03 |
H319A | R-DHPS | NAD+ | 0.30 | 3.7 × 10−3 ± 0.03 | 1.2 ± 0.3 | 3.1 × 10−3 ± 0.10 |
D352A | R-DHPS | NAD+ | 0.30 | ND | ND | ND |
The 272 HpsN sequences were used to generate a SSN at varying alignment scores (AS) (Fig. S13†). An SSN with AS = 170 was chosen as it generated an SSN with a single cluster, but which naturally segregated into interconnected sub-clusters. These sub-clusters exhibited high intra-subcluster connectivity and low inter-subcluster connectivity, and their fine structure aligned with taxonomy at the class level (Fig. 6a). HpsN sequences were distributed across a range of bacterial classes including Alphaproteobacteria, Betaproteobacteria, Gammaproteobacteria, Deltaproteobacteria and Desulfobacteria.
In Fig. 6 we provide examples of gene clusters from organisms that encode the three different pathways. Cupriavidus pinatubonensis JMP134 and Dinoroseobacter shibae DFL12 represent the SuyAB pathway, and contain genes encoding HpsNOP for epimerization of DHPS and oxidation to SL; as well as SuyAB for cleavage of C–S bond of SL to give pyruvate. Jannaschia sp. (strain CCS1), Ruegeria sp. ANG-S4, and Granulosicoccus antarticus are representatives of the Xsc pathway and include genes encoding HpsNOP for epimerization of DHPS and oxidation to SL; SlcD for oxidation of SL to sulfopyruvate; ComDE for decarboxylation of sulfopyruvate to sulfoacetaldehyde; and Xsc for cleavage of the carbon sulfur bond to give acetyl phosphate and sulfite. Ruegeria pomeroyi DSS-3 and Roseovarius nubinhibens are representatives of the CuyA pathway and feature genes encoding HpsNOP for epimerization of DHPS and oxidation to SL; SlcD for oxidation of SL to sulfopyruvate; and CuyA, which catalyzes deamination, sulfite elimination and formation of pyruvate. The identity of the genes encoding aminotransferase (CoA) that converts sulfopyruvate to cysteate, and the cysteate racemase (CuyB) in R. pomeroyi, are unknown.19,20
To further study the neighboring genes, we employed the EFI-GNT tools to examine open reading frames within a ±10 range. The genes were used to construct an SSN of neighbors (SSNN). The SSNN revealed isofunctional clusters for the three sulfolyases: Xsc (n = 12), SuyB (n = 58; including four members with a fused SuyA-SuyB protein), and CuyA (n = 34). However, the total number of members of these three clusters (n = 104) was much smaller than the 272 HpsN sequences used in the original SSN. It is worth noting that sulfolyases are not always co-located with HpsNOP, leading us to manually conduct a BLASTp search for each pathway protein in the organisms that lacked adjacent genes encoding sulfolyase enzymes. This search revealed that 125 bacteria contained SuyB, 112 contain CuyA, and 66 contain both Xsc and ComDE.
Nodes in the HpsN SSN were colored based on the presence of the hpsN gene in a bacterium containing putative pathways for the breakdown of SL through different sulfolyases: the SuyAB pathway (93/272; containing hpsNOP, suyB); the CuyA pathway (51/272; containing hpsNOP, cuyA); and the Xsc pathway (37/272; containing hpsNOP, comDE, xsc) (Fig. 6c–e). To categorize whether individual organisms contained multiple pathways we employed a Venn diagram (Fig. 6b). This enumerates organisms containing both the SuyAB and CuyA pathways (n = 32), and organisms containing both the Xsc and CuyA pathways (n = 29). The SuyAB pathway was found in all the Proteobacteria and Thermodesulfobacteria candidates, the CuyA pathway was found in some of the Alphaproteobacteria, Gammaproteobacteria and Thermodesulfobacteria; while the Xsc pathway was only found in the Alphaproteobacteria.
Histidinol dehydrogenases from E. coli and other organisms are Zn2+ dependent enzymes. Therefore, CpHpsN reconstituted with Zn2+ was used for all kinetic and structural studies. Treatment of Zn2+-loaded CpHpsN with EDTA resulted in a loss of activity, confirming that CpHpsN is a metallo-enzyme. A metal screen of a range of divalent metals identified good activity for a wide range of transition and main group dications, with Zn2+ among the most active. The 3D structure of CpHpsN reveals a dimer with similar fold and quaternary structure to E. coli and M. truncatula histidinol dehydrogenases. All enzymes contain an octahedral metal binding site, formed by amino acid residues from both protomers within the dimer, and binds substrate in similar orientations about the zinc centre, but with the sulfonate group of R-SL taking the place of the imidazole group of L-histidine.
Comparison of complexes of CpHpsN·Zn2+ with CpHpsN·Zn2+·NADH·R-SL reveals conformational changes in two flexible loops that allow remodeling of the coordination environment about the zinc center. In the absence of R-SL, the axial site is occupied by a conserved histidine residue (His126) in one loop, which is displaced in the presence of R-SL, while a conserved aspartate (Asp352) in another loop is repositioned and occupies an axial site. This remodeling of the zinc coordination environment appears to be unique to DHPS 3-dehydrogenases, as histidinol dehydrogenases lack a residue equivalent to His126, and identical coordination environments being observed with, and without, L-histidinol or L-histidine bound.
Our data collectively reveal a characteristic sequence motif that distinguishes DHPS and histidinol dehydrogenases (Fig. 5d and e). For CpHpsN this motif comprises residues 122–126, with the final His126 being conserved among HpsN homologues and binding at the axial site of Zn2+ in the absence of substrate and cofactor. This residue relinquishes its role in zinc coordination to Asp352 upon substrate and cofactor binding, and instead participates in substrate recognition and coordination through a hydrogen bond with the sulfonate group of R-DHPS. In the case of MtHisD, the equivalent motif comprises residues 170–174 and lacks the terminal His residue found in HpsN homologues. Kinetic analysis of the individual CpHpsN Glu318Ala, His319Ala, Asp352Ala variants showed large decreases in catalytic activity, consistent with these residues playing a role in zinc coordination and the catalytic mechanism of CpHpsN. By analogy with the two-phase mechanism proposed for EcHisD,23,24 we propose that catalysis by CpHpsN involves initial binding of R-DHPS to one axial and one equatorial sites of Zn (sites #1 and #3) (Fig. 5b). We propose that, like EcHisD, His319 acts as a general base to deprotonate the primary hydroxyl of R-DHPS and promote hydride transfer to NAD+ to generate R-SLA (Fig. 7). Next, Glu318 acts as general base, promoting nucleophilic addition of water to R-SLA, while protonated His319 acts as general acid, generating R-SLA hydrate. In the second phase, His319 acts as general base to deprotonate the R-SLA hydrate and promote a second hydride transfer to a second molecule of NAD+, forming R-SL.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc05114a |
This journal is © The Royal Society of Chemistry 2024 |