Next Article in Journal
Codium fragile Extract Ameliorates Respiratory Function by Controlling Allergic Inflammation in Ovalbumin-Induced Bronchial Disorders in Mice
Previous Article in Journal
Bioactivity Assessment and Untargeted Metabolomics of the Mediterranean Sea Pen Pennatula phosphorea
Previous Article in Special Issue
The Isolation, Structural Characterization, and Biosynthetic Pathway of Unguisin from the Marine-Derived Fungus Aspergillus candidus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

The Enigma of Sponge-Derived Terpenoid Isothiocyanate–Thiocyanate Pairs: A Biosynthetic Proposal

by
Tadeusz F. Molinski
1,2
1
Department of Chemistry and Biochemistry, MC3568, University of California, 9500 Gilman Drive, La Jolla, San Diego, CA 92093, USA
2
Skaggs School of Pharmacy and Pharmaceutical Sciences, University of California, 9500 Gilman Drive, La Jolla, San Diego, CA 92093, USA
Mar. Drugs 2025, 23(5), 220; https://doi.org/10.3390/md23050220
Submission received: 18 February 2025 / Revised: 9 May 2025 / Accepted: 10 May 2025 / Published: 21 May 2025
(This article belongs to the Special Issue Biosynthesis of Biologically Active Marine Natural Products 2025)

Abstract

:
The co-occurrence of rare terpenoid thiocyanates (R-SCN), structurally similar to their more common isothiocyanate isomers (R-NCS), poses an enigma: how does the accepted path, terpenyl cation R+ → R-NC → R-NCS, accommodate R-SCN? The mystery can now be rationalized by the consideration of three biosynthetic motifs: terpenoid carbocation (R+) capture by cyanoformate, NC-COOH (itself in equilibrium with NC and CO2); co-localized rhodanese (a dual-function enzyme) that can both convert fugitive inorganic NC to thiocyanate ion, NCS, and alkyl isonitriles to alkyl isothiocyanate (R-NC → R-NCS) and adventitious capture of the NCS by R+. The former two scenarios explain the preponderance of isothiocyanates, R-NCS, as products of a linear reaction path—the α-addition of S0 to R-NC—and the third scenario explains minor, less stable thiocyanates, R-SCN, as products of the adventitious capture of liberated NCS by the penultimate R+ precursor. DFT calculations support this proposal and eliminate other possibilities, e.g., the isomerization of R-NCS to R-SCN.

Graphical Abstract

1. Introduction

Exotic terpenoid isonitriles (TIs, Figure 1), which occur exclusively within the domain of certain genera of marine sponges (Porifera), e.g., Acanthella, Adocia, Axinella, Axynissa, Cymbastella, among others, and the sea slugs (nudibranchs) that prey upon them have been known since the early 1970s. Marine isonitriles and their derivatives have been extensively reviewed [1,2,3]. Examples of cyclized terpenoid isonitriles and their accompanying α-adducts (1ac, 2al, 3) are depicted in Figure 2.
Only recently have certain members (e.g., 7,20-diisocyanoadociane (3) [4] and kahilinols A (1a) [5] and B (1b) [6]) been identified and investigated as potential antimalarial therapeutics. The natural product 1b exhibits potent activity against the chloroquine-resistant plasmodium Dd2 (IC50 = 4.6 nM). MED6-189 (1c), a synthetic analog developed from 1b, shows promising in vitro antiplasmodial activity (Dd2 IC50 = 47 ± 7 nM) and strong efficacy in plasmodium-infected mice, with no apparent toxicity or hemolytic liabilities [7].
TIs are often accompanied by the corresponding products of α-addition to the triple -NC bond: amides, carbonimidic dichlorides (CIDs), and, most commonly, alkyl isothiocyanates (ITCs) (Figure 1) [8]. Sponge-derived isonitrile biosynthesis can be rationalized by the capture of late-stage terpenoid carbocations arising from activated precursors (GPP, FPP, and GGPP) that promote C-C cyclizations, Wagner–Meerwein rearrangements, and a final capture by a biosynthetic ‘NC vector’ [9] to generate isonitriles, R-NC. Isothiocyanates, R-NCS, e.g., kalihinol M (1d [10]) (Figure 2), and 2i, j, k, follow as the products of sulfur transfer reactions from low-valent S donors (e.g., thiosulfate, etc.) to R-NC. The terpenoid carbon backbones of TIs and related derivatives, including isothiocyanates (ITC, Figure 1), occur in limited plant species, but TIs are synthesized only by sponges. For example, the common plant sesquiterpene (C15) skeletons of amorphane, drimane, and bisabolane are all represented in the sponge TI, ITC, and TC family, but sponges and their predators (nudibranchs) are the exclusive provenance of the carbon skeletons axinyssane [11,12], pupukeanane [13,14], neopupukeanane [15] and the diterpenes (C20) adociane [4], amphilectane [16,17], and isoneoamphilectane [18], among others.
The startling discovery of co-localized thiocyanates, TCs, (R-SCN), e.g., the cis-fused bicyclic 4-thiocyanato-9-cadinene 2a [19] (more accurately, an amorphene, having the same carbon skeleton as the amorphene isonitrile 2l [20]), and regioisomeric 2- and 4-thiocyanatoneopupukeananes (2c,d) [21,22] poses an enigma: ‘how can a linear biosynthetic pathway of R-NC → R-NCS accommodate R-SCN?’ In plants, the biosynthesis of allyl thiocyanate and its isomeric allyl isothiocyanate, and related pungent volatile natural products of mustard oil, horseradish, garlic, and many other Cruciferate, has long been understood [23], e.g., decomposition of the glucosinolate, sinigrin, derived from the amino acid, Met, but the origin of sponge-derived thiocyanates is less clear. Other unusual features present themselves. For example, the incipient precursor of 2c and 2d resembles the storied ‘non-classical’ norbornyl cation [24] where the positive charge is delocalized between the C2 and C4 positions which may explain the outcomes of SN: exo TC regioisomers isolated in almost equal proportions [21,22] with the same skeleton as 9-isocyanoneopupukeanane 2n [15]. Rounding out the thiocyanates are (–)-cavernothiocyanate (2m) [25] and 2g, h [26], which share the pupukeanane skeleton with isothiocyanates (R-NCS, e.g., 2i [27], j [14].
Now this enigma is addressed by the consideration of the putative interplay of three integrated biosynthetic sequelae: the capture of the terpenoid carbocation R+ by cyanoformate, NC-COOH (CF, see below [9] (itself in equilibrium with NC and CO2 and derived from Gly), reactions promoted by co-localized rhodanese, a dual-function enzyme that can both convert inorganic NC to thiocyanate ion, NCS but also alkyl isonitriles to alkyl isothiocyanate (R-NC → R–NCS), and the promiscuous interception of terpenoid carbocations by co-liberated inorganic NCS.

2. Results and Discussion

2.1. Chemistry and Bonding in TCs and ITCs

While hundreds of marine-derived TIs, ITCs, and FAs have been described [1], only seven TCs are known (Figure 2). The structures of the TCs have been examined on a case-by-case basis, with their isomeric or similar isothiocyanates (ITCs). The ‘reverse engineering’ of the likely terminal steps in their biosynthesis by probing the thermodynamics of bond breaking and bond forming provides insights into what could be possible and what is improbable. For the purposes of an informed discussion, a brief review of the chemistry of the cumulated-bond functional group, NCS, in TC and ITC molecules is in order. TC and ITC isomers are easily differentiated spectroscopically by 13C NMR and IR. ITC shows a very intense, broad stretching frequency at ~2100 cm−1, while the TC IR band is about the same frequency, but narrower and of medium intensity. The 13C NMR signal of C at the point of attachment reflects the difference in the electronegativity of S and N: ITC shows C-Nα of approximately δ~75, while the C-Sα signal in TC is δ~60 ppm [19].
The isomers t-butylthiocyanate (4) and t-butylisothiocyanate (5) (Figure 3) serve as convenient models for comparisons with their terpenoid natural product counterparts. Bonding in the triatomic grouping NCS differs between TCs and ITCs. The DFT calculations of the optimized geometries and energies of each (Figure 1) show, as expected, that the geometry of N-C-S is linear due to sp hybridization at C (ω = 179.8°), consistent with the X-ray crystal structures of the natural products ITCs [14] and TCs 19]. The bonding of the heteroatom attached to the alkyl group C (notated, in this paper, as the α atom) differs significantly. The corresponding bond angles and bond lengths of 4 and 5 (θ(C-S-C) = 100° and (θ(C-Nα-C) = 176.8°) reflect predominantly sp3 and sp hybridization, respectively, of the α-heteroatom attached to the t-Bu group. The shorter C-Nα bond length in 5 (d2 = 1.44 Å) is evident of a stronger σ bond between the t-Bu group and its α-heteroatom compared to 4 (d1 = 1.88 Å; see below, as expected, for row 2 versus row 3 elements) recapitulates the thermodynamic stability of 5 over 4 [19]. The DFT minimized structure of 1-isothiocyano-1-methylcyclohexane (S1) displays similar bond parameters. From the DFT-calculated energies of 4 and 5, the latter is shown to be more stable (∆E = –14.8 kcal·mol−1; see Supporting Information), and the isomerization of 5 to 4 is not spontaneous.
ITCs are electrophilic: like isonitriles, they undergo addition reactions at the cumulated bond, even with weak nucleophiles. TCs reluctantly undergo SN2 substitution reactions at the attached C when 1° or 2°, but not 3°. Allylic ITCs and TCs both participate in sigmatropic [3] rearrangement (see below). Three scenarios that may account for the generation of TC natural products are considered.

2.2. Dissociation–Reassociation

The isomerization of ITC to TC was examined as a possible linear pathway linking the corresponding TI (R-NC → R–NCS → R-SCN). Necessarily, such an isomerization must follow a unimolecular rate law, R = k[R-NCS], and an SN1 mechanism due to the 2° or 3° alkyl group that is almost invariably the point of heteroatom attachment in TI, ITC, and TC natural products. The leaving group ability of the thiocyano group in R-SCN is at best weak, while in R-NCS, it is considerably poorer due to the stronger R-N bond. Experimentally, the kinetic outcome of SN reactions with NCS favors the TC product [28].
For various reasons, calculating the competitive rates of the SN reaction of a row 2 element, N, with a row 3 element, S, is difficult, but we can easily obtain semi-quantitative estimates from the product distributions of sigmatropic [3] rearrangements of allylic thiocyanate (e.g., 6a) to isothiocyanates (e.g., 6b) and the degenerate rearrangement of isoelectronic allyl azide (7, Figure 4), where C-N/C-S bond-forming and bond-breaking reactions are simultaneous. While the reaction rates of the [3] rearrangement of allylic azides (Winstein rearrangement [29]) are fast (most rapidly isomerize at ambient temperatures [30]), CH2=CH-NCS and CH2=CH-SCN interchange more slowly [28]. For example, the isomerization of the latter to the former occurs upon distillation (bp 150-2 °C) [28]); at Keq, the ITC is favored because of the greater σ bond strength of C-N in the product (vide supra). The explanation of the slower rate reaction of [3]-rearrangement of ITC-TC pairs is poor overlap of frontier orbitals in the transition state (row 2 versus row 3 element). At equilibrium in cyclohexane, allyl thiocyanate is undetectable, but in acetonitrile the mol fraction becomes 9–11%, a result consistent with the higher dipole moment of allyl isothiocyanate and an ionic component (dissociative) to the transition state [29].

2.3. Adventitious Capture of NCS

ITCs are expressed by bacteria from isonitriles by task-specific adapted rhodanese. Rhodanese is distributed widely in the Nature and carries out the important role of scavenging fugitive inorganic cyanide formed during metabolism, e.g., ‘leakage’ from C-1 tetrahydrofolate metabolism or other cyanide-generating reactions [31,32,33]. A key intermediate, cyanoformate (NC-COOH, CF), is formed from 1-aminocyclopropane-1-carboxylate (oxidation product of Gly) in the biosynthesis of the plant hormone ethylene (CH2=CH2) [34]. In Burkholderia gladioli that produces a non-terpenoid isonitrile, Hertwick and coworkers showed that rhodanese RhDE, in addition to donating competency to scavenge cyanide NC into thiocyanate, NCS, catalyzes the substrate-specific sulfur transfer reaction R-NC → R-NCS generating an ITC, sinapigladioside [35], by utilizing thiosulfate as an S donor [36]. Hertwick speculates that the enzyme was recruited from a ‘ubiquitous detoxification enzyme for the formation of a bioactive specialized metabolite’ [36]. The concept of the bacterial recruitment of rhodanese, with evolved substrate specificity to the latter task, may also find support in the biosynthesis of terpenoid ITCs in sponges and the observation that TIs are not uniformly accompanied by their ITC counterparts. While rhodanese is ubiquitous across the Kingdoms of Life, no reports of its specific occurrence in Porifera have appeared yet.

2.4. Alkyl TC from SN Reaction with Kinetically Competent NCS

Pearson and coworkers compiled the relative rates of nucleophilic substitutions of methyl iodide, CH3I, with selected nucleophiles under comparable conditions (25 °C, Table 1) [37]. The second-order rate constant values, k2, for the reaction with NCS and NC are similar (Entries 8 and 9: 5.74 × 10−4 and 6.5 × 10−5 mol−1·s−1, respectively (for comparison, the k2 values for the reactions with thiophenoxide, PhS, and thiosulfate, S2O32–, are 1.07 and 0.114 mol−1·s−1 [37]).
A simplified scenario can be proposed for the terminal step in TC and ITC biosynthesis (Scheme 1). The SN1 capture of NCS at S (the more nucleophilic end) yields TCs (minor axis). ITCs formed in this way could contribute to the pool of ITCs generated by the S° addition to the corresponding TIs. Experimentally, it is conceivable that the fraction of ITCs that arises from minor and major axes could be estimated from isotopic labeling (see below).
Inspection of the structures of secondary metabolites containing the R-SCN group (TCs) precludes an enzyme system for the tailoring of R-NCS to R-SCN, but can support a role of rhodanese as a generator of the ambident nucleophile, NCS. ITCs may also arise from the less-favored capture of NCS at N by SN1 addition, a reaction that would add to the pool of ITCs arising from the major axis (Scheme 1).
In the laboratory, mixtures of TCs and ITCs are formed from the reaction of NCS with electrophiles, e.g., R-Br. In contrast, the replacement of NCS with NC in SN substitution reactions yields only nitriles, R-CN, not isonitriles, RNC, by an exclusive nucleophilic addition at C [9]. At equilibrium, isomerization would favor ITCs over TCs based on the relative bond strengths of C-N compared with C-S (see Supporting Information), but this is kinetically prohibitive.

2.5. A Simple Proposal—Thiocyanates Can Arise Adventitiously

How do these foregoing data affect the interpretation of the provenance and distribution of alkyl isothiocyanate, ITC, versus alkyl thiocyanate, TC? Alkyl TCs are rarer than their isomeric ITCs. Two pathways or axes for the generation of S isonitriloids can be proposed (Scheme 1). Biosynthetically, ITCs arise mostly from the linear pathway (major axis, i) of TIITC by the formal addition of S° to the terminus of R-NC from a suitable donor. The less abundant TCs more likely arise from free thiocyanate, NCS (minor axis, ii), delivered by rhodanese-promoted capture of fugitive cyanide, and its nucleophilic addition to the terpenoid carbocation, R+ (see Supporting Information for expanded Scheme S1). The partition ratio, p (TC:ITC), along the minor axis ii may be measurable by {14C}-NCS labeling in a suitable sponge candidate known to produce both isomeric ITCs and TCs. TCs would be expected to predominate over ITCs (p > 1) in ii, but, overall, i dominates the production of ITCs.
Scheme 1 leads to a unifying biosynthetic proposal for the biogenesis of TC, ITC and TI isonitriloids; for example, the three iso-morphic neopupukeanane natural products, 2c, d, k (Scheme 2), link thiocyanate and isonitrile formation. The ionization of farnesyl pyrophosphate (FPP), followed by stepwise cyclizations, as first proposed by Scheuer [15] and Opatz [1] and respective coauthors, yields stabilized incipient carbocation, i (represented in this paper by non-classical resonance forms [24,38]), which is captured by a C1 nucleophile (NCS or NC-COOH) to yield the isonitriloids 2c, d, k. While the absolute configuration (AC) of neopupukeanane isonitriloids have not been independently assigned, 2c, d, k as depicted here conform to the rational biosynthetic proposal by Scheuer from amorphanyl cation [15] and his original assignment of the AC of the pupukeanane skeleton by CD [13].
One question remains: ‘how is exogenous NCS assimilated by sponges that make TC-ITC pairs?’ HCN is a weak Brønsted acid (pKa = 9.25), but HSCN is strong (pKa = −0.7 [39] and fully ionized at physiological pH values. While exogenous cyanide can intercept the intracellular dynamic equilibrium of cyanoformate (CF, NC-COOH) by passive diffusion in its neutral form, HCN, the pKa of HSCN would seem prohibitively low for a similar process; a ‘thiocyanate ion transporter’ would need to be invoked [40]. Yet, there is precedence. Thiohalobacter spp., with induced capacity to metabolize NCS, have been raised by repeated passage in culture media containing NCS [41].
It is notable that, in the TI-ITC-TC-producing sponges, capture by H2O is rarely or never observed (the absence of corresponding alcohols, R-OH, or alkene elimination products), suggesting a tight coupling of terpenoid–NC capture, possibly in a reaction within an enzyme active site where H2O molecules are excluded. The evaluation of this proposal must await a missing dataset: the first high-resolution X-ray or cryo-EM structure of an isonitriloid terpenoid cyclase.
Allylic thiocyanate, 8c, and isothiocyanate, 8b (Figure 1), hypothetically obtainable from the nucleophilic substitution of farnesyl pyrophosphate, or 8b separately from S-transfer to farnesyl isonitrile (8a) (which, itself, ‘remains elusive’ [1]), are kinetically labile. TCs, in time, would be expected to isomerize by allylic rearrangement to their more substituted tert-NCS counterparts (e.g., 9). Most sponge-derived TCs, however, are ‘fixed’, stabilized by substitution at 2° or 3° alkyl groups. Farnesyl isothiocyanate (8b) [42] and formamide, 8d, are known natural products [43], but the foregoing reasons make it unlikely that the ‘missing’ 8c will ever be isolated from natural sources, as [3] rearrangement irreversibly converts 8c to its isomeric nerolidyl isothiocyanate isomer, 9. The fact that 9 has not been found in Nature suggests that 8c, too, is absent. The syntheses of 8c and its geranyl homolog were found as mixtures of 8c and the [3] rearrangement product, 9 (8c:9~83:17), even after mild vacuum distillation conditions [44]. Although allyl derivatives 6 and 7 exhibit good antibacterial properties, their geranyl and farnesyl analogs (e.g., 8b,c) showed none. It should be noted that the names ‘stylotellane A’ and ‘stylotellane B’ were conferred upon the carbonimidic dichloride 10 and related 11 [11] from Styletta aurantium [40]. Compound 11 was originally isolated by Wratten and Faulkner from Pseudaxinyssa pitys—the first example of this functional group in a natural product [11].
Compounds 8b,d and 10 complete the set of known farnesyl N-derivatives. It appears that 2m is the only allylic TC among these marine natural products, with the SCN group substituted at the ‘tail’ of the first isoprene group in the precursor FPP. The unexpected, unbranched long-chain ITCs, 12, from Pseudoaxinyssa, contain vinyl-substituted α,ω-bisisocyanato groups and appear to have different, indeterminate biosyntheses [45].
One consequence of a strictly unimolecular rate law of reaction of NCS with chiral R+ is diastereomeric mixtures. The product distribution will be dependent upon the usual stereoelectronic factors that govern the stereofacial preferences of SN1 reactions. For example, in the case of epimeric 2g and 2h, isolated from the sponge Axinyssa aculeata [20], the reported ratio is 3:2 [46], but, from local symmetry considerations, there would be little or no stereofacial preference for attack on the incipient carbocation. In Phyllidia varicosa, the nudibranch that depredates A. aculeata and sequesters 2g and 2h, the epimer ratio is altered: in the dorsal mantle it is 1:2, but about equimolar in the digestive organ. Fractionation of secondary metabolites by nudibranchs from their sponge diets, as well as metabolic modification, have both been described before, e.g., trisoxazole macrolides from the Spanish Dancer, Hexbranchus sangineus [46].
When thought of in this way, the capture of carbocations by free NCS is a ‘clock reaction’, a kinetic monitor of the partition between the shunt reactions of cyanoformate, CF (NC-COO Scheme 1 and Scheme 2)—with dissociation to NC and CO2—and the direct nucleophilic SN1 capture by NCS. Garson [12,47] and others [48] have shown through numerous radiolabeling experiments that inorganic [14C]-CN is assimilated into sponge TIs. For example, 7,20-diisocyanoadociane (3) is radiolabeled by [14C]-CN [49]. These observations have recently been interpreted as an interception of an equilibrium cyanide pool formed by the dissociation of CF, the putative ‘NC’ vector, and exogenous uptaken HCN [9]. In the same study, Garson found no radiolabel incorporation into 3 when the sponge was incubated with [2-14C] Gly or a number of other amino acids [50]. It is felt that a lack of detection of radiolabeled 3 after incubation with [2-14C]Gly is not paradoxical, but rather a consequence of below-threshold incorporation and much faster assimilation into sponge tissue protein and other intermediary metabolism.
The prevailing belief to date was that inorganic NC is the precursor of TIs, but, for several reasons, this was shown to be untenable [9]. If it were so, a simple test can be made: the expected labeling of living sponges with [14C]-CN should also induce the formation of limited amounts of R-SCN (from rhodanese-generated NCS-favored SN1 nucleophile capture at S), but this has not been observed to date. A counter explanation would be that fraction of exogenous cyanide that is not incorporated into the structures of sponge TIs is rapidly and efficiently converted to NCS by rhodanese in separate cellular compartments, spatially distal from the active sites of TI biosynthetic enzymes and becomes unavailable for TC or ITC production. Garson and coworkers showed that radiolabeled thiocyanate (K [14C]-NCS) is assimilated by living sponges into TIs, including axisonitrile-3 (notably with a formula lacking S), and two ITCs (axisothiocyanate-3), albeit with far lower specific activities than the same radiolabeled natural products obtained from incubations with Na{14C]-CN [51]. The simultaneous labeling of axisonitrile-3 from the incubation of the sponge Acanthella cavernosa [52] with K [14C]-NCS requires, at the very least, an unspecified interchange of thiocyanate with K [14C]-CN (e.g., it is known mammalian oxyhemoglobin can convert thiocyanate to cyanide [53]). In a separate study, Simpson and Garson found the incorporation of both [14C]-SCN and [14C]-CN into the sesquiterpene thiocyanato group of 2b [21], produced by Amphimedon terpenensis from the Great Barrier Reef [54]; both results are consistent with the adventitious thiocyanate model for TC biosynthesis (Scheme 1). What could be testable in the field is the incorporation of {2-14C}-NC-COOH in sponges producing TIs, ITCs, and TCs or {14C}-NCS under more controlled conditions. The outcomes would be compelling, but the logistics of this proposed experiment are beyond the scope of this work.

3. Materials and Methods

General Experimental Procedures

DFT Calculations. All DFT calculations were performed using Spartan ’20 (Wavefunction, Irvine, CA, USA), using functional ωB97X-D and basis set 6-31G* (H2O or gas phase). The coordinates for the optimized structures of 4, 5, and S1 and bond parameters and energies can be found in the Supporting Information. See Supporting Information for a complete citation for the DFT product.

4. Conclusions

A unified biosynthetic scheme is proposed to explain the naturally occurring TI, ITC, and TC isonitriloids that engage the ‘NC’ group of cyanoformate (NC-COO). While the major axis of the biosynthesis of ITCs involves S° transfer to TIs, isomeric TCs are outliers on the linear path ITITC. It is proposed TCs arise from the adventitious capture of incipient terpenoid carbocation, R+, by inorganic NCS, itself delivered by the rhodanese-promoted capture of fugitive NC by an S° donor.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/md23050220/s1. DFT-calculated structures of 4, 5, and 1-isothiocyanato-1-methylcyclohexane (S1).

Funding

This work was not supported by funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Original data will be made available upon reasonable request.

Acknowledgments

The author thanks S. Ohlinger (Spartan) for assistance with the DFT calculations, and many other individuals over the years for helpful discussions. This paper is dedicated to my colleague and friend Haiyin He, whose discovery of the first terpenoid thiocyanate in 1989 inspired this paper.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Emsermann, J.; Kauhl, U.; Opatz, T. Marine Isonitriles and Their Related Compounds. Mar. Drugs 2016, 14, 16. [Google Scholar] [CrossRef] [PubMed]
  2. Garson, M.J.; Simpson, J.S. Marine isocyanides and related natural products—Structure, biosynthesis and ecology. Nat. Prod. Rep. 2004, 21, 164–179. [Google Scholar] [CrossRef] [PubMed]
  3. Chena, T.-Y.; Chen, J.; Tang, Y.; Zhou, J.; Guo, Y.; Chang, W.-C. Current Understanding Toward Isonitrile Group Biosynthesis and Mechanism. Chin. J. Chem. 2021, 39, 463–472. [Google Scholar] [CrossRef] [PubMed]
  4. Baker, J.T.; Wells, R.J.; Oberhänsli, W.E.; Hawes, G.B. A New Diisocyanide of Novel Ring Structure from a Sponge. J. Am. Chem. Soc. 1976, 98, 4010–4012. [Google Scholar] [CrossRef]
  5. Chang, W.J.; Patra, A.; Roll, D.M.; Scheuer, P.J.; Matsumoto, G.K.; Clardy, J. Kalihinol-A, a highly functionalized diisocyano diterpenoid antibiotic from a sponge. J. Am. Chem. Soc. 1984, 106, 4644–4646. [Google Scholar] [CrossRef]
  6. Patra, A.; Chang, C.W.J.; Scheuer, P.J.; Van Duyne, G.D.; Matsumoto, G.K.; Clardy, J. An Unprecedented Triisocyano Diterpenoid Antibiotic from a Sponge. J. Am. Chem. Soc. 1984, 106, 7981–7983. [Google Scholar] [CrossRef]
  7. Chahine, Z.; Abel, S.; Hollin, T.; Barnes, G.L.; Chung, J.H.; Daub, M.E.; Renard, I.; Choi, J.Y.; Vydyam, P.; Pal, A.; et al. A kalihinol analog disrupts apicoplast function and vesicular trafficking in P. falciparum malaria. Science 2024, 385, 7966–7978. [Google Scholar] [CrossRef]
  8. Cafieri, F.; Fattorusso, E.; Magno, S.; Santacroce, C.; Sica, D. Isolation and Structure of Axisonitrile-1 and Axisothiocyanate-1. Two Unusual Sesquiterpenoids from the Marine Sponge Axinella cannabina. Tetrahedron 1973, 29, 4259–4262. [Google Scholar] [CrossRef]
  9. Molinski, T.F. The Paradox of Antimalarial Terpenoid Isonitrile Biosynthesis Explained. Proposal of Cyanoformate as an NC Delivery Vector. J. Nat. Prod. 2025, 88, 205–210. [Google Scholar] [CrossRef]
  10. Xu, Y.; Li, N.; Jiao, W.-H.; Wang, R.-P.; Peng, Y.; Qi, S.-H.; Song, S.-J.; Chen, W.-S.; Lin, H.-W. Antifouling and cytotoxic constituents from the South China Sea sponge Acanthella cavernosa. Tetrahedron 2012, 68, 2876–2883. [Google Scholar] [CrossRef]
  11. Wratten, S.J.; Faulkner, D.J. Carbonimidic dichlorides from the marine sponge Pseudaxinyssa pitys. J. Am. Chem. Soc. 1977, 99, 7367–7368. [Google Scholar] [CrossRef] [PubMed]
  12. Brust, A.; Garson, M.J. Dereplication of complex natural product mixtures by 2D NMR: Isolation of a new carbonimidic dichloride of biosynthetic interest from the tropical marine sponge Stylotella aurantium. ACGC Chem. Res. Commun. 2004, 17, 33–37. [Google Scholar]
  13. Hagadone, M.R.; Burreson, B.J.; Scheuer, P.J.; Finer, J.S.; Clardy, J. Defense Allomones of the Nudibranch Phyllidia varicosa Lamarck 1801. Helv. Chim. Acta 1979, 62, 2484–2494. [Google Scholar] [CrossRef]
  14. Marcus, A.H.; Molinski, T.F.; Fahy, E.; Faulkner, D.J.; Xu, C.; Clardy, J. 5-Isothiocyanatopupukeanane from a sponge of the genus Axinyssa. J. Org. Chem. 1989, 54, 5184–5186. [Google Scholar] [CrossRef]
  15. Karuso, P.; Poiner, A.; Scheuer, P.J. Isocyanoneopupukeanane, a new tricyclic sesquiterpene from a sponge. J. Org. Chem. 1989, 54, 2095–2097. [Google Scholar] [CrossRef]
  16. Wratten, S.J.; Faulkner, D.J.; Hirotsu, K.; Clardy, J. Diterpenoid Isocyanides from the Marine Sponge Hymeniacidon amphilecta. Tetrahedron Lett. 1978, 19, 4345–4348. [Google Scholar] [CrossRef]
  17. Molinski, T.F.; Faulkner, D.J.; Van Duyne, G.D.; Clardy, J. Three New Diterpene Isonitriles from a Palauan Sponge of the Genus Halichondria. J. Org. Chem. 1987, 52, 3334–3337. [Google Scholar] [CrossRef]
  18. Avile, E.; Rodriguez, A.D.; Vincente, J. Two Rare-Class Tricyclic Diterpenes with Antitubercular Activity from the Caribbean Sponge Svenzea flava. Application of Vibrational Circular Dichroism Spectroscopy for Determining Absolute Configuration. J. Org. Chem. 2013, 78, 11294–11301. [Google Scholar] [CrossRef]
  19. He, H.Y.; Faulkner, D.J.; Shumsky, J.S.; Hong, K.; Clardy, J. A sesquiterpene thiocyanate and three sesquiterpene isothiocyanates from the sponge Trachyopsis aplysinoides. J. Org. Chem. 1989, 54, 2511–2514. [Google Scholar] [CrossRef]
  20. Fusetani, N.; Wolstenholme, H.J.; Matsunaga, S. Co-occurrence of 9-isocyanopupukeanane and its C-9 epimer in the nudibranch Phyllidia bourguini. Tetrahedron Lett. 1990, 31, 5623–5624. [Google Scholar] [CrossRef]
  21. Pham, A.T.; Ichiba, T.; Yoshida, W.Y.; Scheuer, P.J.; Uchida, T.; Tanaka, J.; Higa, T. Two marine sesquiterpene thiocyanates. Tetrahedron Lett. 1991, 32, 4843–4846. [Google Scholar] [CrossRef]
  22. He, H.; Salvá, J.; Caiíalos, R.F.; Faulkner, D.J. Sesquiterpene Thiocyanates and Isothiocyanates from Axinyssa aplysinoides. J. Org. Chem. 1992, 57, 3191–3194. [Google Scholar] [CrossRef]
  23. Blaževića, I.; Montaut, S.; Burčul, F.; Olsend, C.E.; Burowe, M.; Rollin, P.; Agerbirk, N. Glucosinolate structural diversity, identification, chemical synthesis and metabolism in plants. Phytochemistry 2020, 169, 112100. [Google Scholar] [CrossRef]
  24. Winstein, S.; Trifan, D.S. The Structure of the Bicyclo[1,2]2-Heptyl (Norbornyl) Carbonium Ion. J. Am. Chem. Soc. 1949, 71, 2953. [Google Scholar] [CrossRef]
  25. Fusetani, N.; Wolstenholme, H.J.; Shinoda, K.; Asai, N.; Matsunaga, S.; Onuki, H.; Hirota, H. Two sesquiterpene isocyanides and a sesquiterpene thiocyanate from the marine sponge Acanthella cf. cavernosa and the Nudibranch Phyllidia ocellata. Tetrahedron Lett. 1992, 33, 6823–6826. [Google Scholar] [CrossRef]
  26. Yasman, Y.; Edrada, R.A.; Wray, V.; Proksch, P. New 9-Thiocyanatopupukeanane Sesquiterpenes from the Nudibranch Phyllidia varicosa and Its Sponge-Prey Axinyssa aculeata. J. Nat. Prod. 2003, 66, 1512–1514. [Google Scholar] [CrossRef]
  27. Simpson, J.S.; Garson, M.J.; Hooper, J.N.A.; Cline, E.I.; Angerhofer, C.K. Terpene Metabolites from the Tropical Marine Sponge Axinyssa sp. nov. Aust. J. Chem. 1997, 50, 1123–1128. [Google Scholar] [CrossRef]
  28. Emerson, D.W. The Preparation and Isomerization of Allyl Thiocyanate. J. Chem. Ed. 1971, 48, 81–82. [Google Scholar] [CrossRef]
  29. Carlson, A.S.; Topczewski, J.J. Allylic azides: Synthesis, reactivity, and the Winstein rearrangement. Org. Biomol. Chem. 2019, 17, 4406–4429. [Google Scholar] [CrossRef]
  30. Gagneux, A.; Winstein, S.; Young, W.G. Rearrangement of allylic azides. J. Am. Chem. Soc. 1960, 82, 5956–5957. [Google Scholar] [CrossRef]
  31. Knowles, C.J.; Bunch, A.W. Microbial Cyanide Metabolism. Adv. Microb. Physiol. 1986, 27, 73–111. [Google Scholar] [PubMed]
  32. Harris, R.E.; Bunch, A.W.; Knowles, C.J. Microbial cyanide and nitrile metabolism. Sci. Prog. 1987, 71, 293–304. [Google Scholar]
  33. Blumer, C.; Haas, D. Mechanism, regulation, and ecological role of bacterial cyanide biosynthesis. Arch. Microbiol. 2000, 173, 170–177. [Google Scholar] [CrossRef]
  34. Hering, C.; von Langermann, J.; Shulz, A. The Elusive Cyanoformate: An Unusual Cyanide Shuttle. Angew. Chem. Int. Ed. 2014, 53, 8282–8284. [Google Scholar] [CrossRef]
  35. Dose, B.; Niehs, S.P.; Scherlach, K.; Shahda, S.; Flórez, L.; Kaltenpoth, M.; Hertweck, C. Biosynthesis of Sinapigladioside, an Antifungal Isothiocyanate from Burkholderia Symbionts. ChemBioChem 2021, 22, 1920–1924. [Google Scholar] [CrossRef]
  36. Mlotek, M.D.; Bose, B.; Hertweck, C. Bacterial Isothiocyanate Biosynthesis by Rhodanese-Catalyzed Sulfur Transfer onto Isonitriles. ChemBioChem 2024, 25, e202300732. [Google Scholar] [CrossRef]
  37. Pearson, R.G.; Sobel, H.; Songstad, J. Nucleophilic Reactivity Constants toward Methyl Iodide and trans-[Pt(py)2Cl2]. J. Am. Chem. Soc. 1968, 90, 319–326. [Google Scholar] [CrossRef]
  38. Walling, C. An Innocent Bystander Looks at the 2-Norbornyl Cation. Acc. Chem. Res. 1983, 16, 448–454. [Google Scholar] [CrossRef]
  39. Belluco, U.; Cattalini, U.; Basolo, F.; Pearson, R.G.; Turco, A. Nucleophilic Constants and Substrate Discrimination Factors for Substitution Reactions of Platinum (II) Complexes. J. Am. Chem. Soc. 1965, 87, 241–246. [Google Scholar] [CrossRef]
  40. Pedemonte, N.; Caci, E.; Sondo, E.; Caputo, A.; Rhoden, K.; Pfeffer, U.; Di Candia, M.; Bandettini, R.; Ravazzolo, R.; Zegarra-Moran, O.; et al. Thiocyanate transport in resting and IL-4-stimulated human bronchial epithelial cells: Role of pendrin and anion channels. J. Immunol. 2007, 178, 5144–5153. [Google Scholar] [CrossRef]
  41. Oshiki, M.; Fukushima, T.; Kawano, S.; Kasahara, Y.; Nakagawa, J. Thiocyanate Degradation by a Highly Enriched Culture of the Neutrophilic Halophile Thiohalobacter sp. Strain FOKN1 from Activated Sludge and Genomic Insights into Thiocyanate Metabolism. Microbes Environ. 2019, 34, 402–412. [Google Scholar] [CrossRef] [PubMed]
  42. Simpson, S.; Raniga, P.; Garson, M.J. Biosynthesis of dichloroimines in the tropical marine sponge Stylotella aurantium. Tetrahedron Lett. 1997, 38, 7947–7950. [Google Scholar] [CrossRef]
  43. Li, Z.-Y.; Yu, Z.-G.; Guo, Y.-W. New N-Containing Sesquiterpenes from Hainan Marine Sponge Axinyssa sp. Helv. Chim. Acta 2008, 91, 1553–1558. [Google Scholar] [CrossRef]
  44. Pala, G.; Mantegani, A.; Coppi, G. Terpene Compounds Mono- and Sesquiterpene Thiocyanates and Isothiocyanates. J. Med. Chem. 1969, 12, 725–726. [Google Scholar] [CrossRef]
  45. Karuso, P.; Scheuer, P.J. Long-chain α,ω-bisisothiocyanates from a marine sponge. Tetrahedron Lett. 1987, 28, 4633–4636. [Google Scholar] [CrossRef]
  46. Pawlik, J.R.; Kernan, M.K.; Molinski, T.F.; Harper, M.K.; Faulkner, D.J. Defensive Chemicals of the Spanish Dancer Hexabranchus sanguineus, and its Egg Ribbons: Macrolides Derived from a Sponge Diet. J. Exp. Mar. Biol. Ecol. 1988, 119, 99–109. [Google Scholar] [CrossRef]
  47. Garson, M.J.; Simpson, J.S.; Flowers, A.E.; Dumdei, E.J. Cyanide and thiocyanate-derived functionality in marine organisms—Structures, biosynthesis and ecology. In Studies in Natural Products Chemistry; Atta ur, R., Ed.; Elsevier: Amsterdam, The Netherlands, 2000; Volume 21, Part B; pp. 329–372. [Google Scholar]
  48. Cimino, C.; De Rosa, S.; De Stefano, S.; Sodano, G. Marine natural products: New results from Mediterranean invertebrates. Pure Appl. Chem. 1986, 58, 375–386. [Google Scholar] [CrossRef]
  49. Garson, M.J. Biosynthesis of the novel diterpene isonitrile diisocyanoadociane by a marine sponge of the Amphimedon genus: Incorporation studies with sodium [14C] cyanide and sodium [2-14C] acetate. J. Chem. Soc. Chem. Commun. 1986, 35–36. [Google Scholar] [CrossRef]
  50. Fookes, C.J.R.; Garson, M.J.; MacLeod, J.K.; Skelton, B.W.; White, A.H. Biosynthesis of diisocyanoadociane, a novel diterpene from the marine sponge Amphimedon sp. Crystal structure of a monoamide derivative. J. Chem. Soc. Perkin Trans. 1 1988, 1003–1011. [Google Scholar] [CrossRef]
  51. Dumdei, E.J.; Flowers, A.E.; Garson, M.J.; Moore, C.J. The Biosynthesis of Sesquiterpene Isocyanides and Isothiocyanates in the Marine Sponge Acanthella cavernosa (Dendy); Evidence for Dietary Transfer to the Dorid Nudibranch Phyllidiela pustulosa. Comp. Biochem. Physiol. 1997, 118A, 1385–1392. [Google Scholar] [CrossRef]
  52. Simpson, J.S.; Garson, M.J. Advanced Precursors in Marine Biosynthetic Study: The Biosynthesis of Diisocyanoadociane in Amphimedon terpenensis. Tetrahedron Lett. 1999, 40, 3909–3912. [Google Scholar] [CrossRef]
  53. Seto, Y. Oxidative conversion of thiocyanate to cyanide by oxyhemoglobin during acid denaturation. Arch. Biochem. Biophys. 1995, 321, 245–254. [Google Scholar] [CrossRef] [PubMed]
  54. Simpson, J.S.; Garson, M.J. Thiocyanate Biosynthesis in the Tropical Marine Sponge Axinyssa n. sp. Tetrahedron Lett. 1988, 39, 5819–5822. [Google Scholar] [CrossRef]
Figure 1. Lewis structures (closed-shell) of terpene isonitriles (TIs), ‘isonitriloids’–isothiocyanates (ITCs), thiocyanates (TCs), carbonimidic dichlorides (CIDs), and products of reaction with H2O formamides (FAs) and amines (As).
Figure 1. Lewis structures (closed-shell) of terpene isonitriles (TIs), ‘isonitriloids’–isothiocyanates (ITCs), thiocyanates (TCs), carbonimidic dichlorides (CIDs), and products of reaction with H2O formamides (FAs) and amines (As).
Marinedrugs 23 00220 g001
Figure 2. Terpenoid isonitriles (TI); kalihinols A (1a), B (1b), M (1d), and synthetic MED6-189 (1c); 7,20-diisocyanoadociane (3); and related isothiocyanates (ITCs, e.g., 1d, 2ij, 8b and thiocyanates (TCs), 2ad, gk, m. bis-Isothiocyanate 12 is a non-terpenoid.
Figure 2. Terpenoid isonitriles (TI); kalihinols A (1a), B (1b), M (1d), and synthetic MED6-189 (1c); 7,20-diisocyanoadociane (3); and related isothiocyanates (ITCs, e.g., 1d, 2ij, 8b and thiocyanates (TCs), 2ad, gk, m. bis-Isothiocyanate 12 is a non-terpenoid.
Marinedrugs 23 00220 g002
Figure 3. DFT-optimized molecular structures (ωB97X-D 6-31G*), dipole moments, key bond angles, and distances, d, of isomeric TC and ITC. (a) t-butylthiocyanate (4): µ = 4.48 D θthio(C-S-C) = 100.0°, dthio(C–S) = 1.88 Å; and (b) t-butylisothiocyanate (5): µ = 6.56 D, θtiso(C–Nα–C) = 176.8°, diso(C-Nα) = 1.44 Å. The major resonance form (closed-shell Lewis structure) is depicted for each.
Figure 3. DFT-optimized molecular structures (ωB97X-D 6-31G*), dipole moments, key bond angles, and distances, d, of isomeric TC and ITC. (a) t-butylthiocyanate (4): µ = 4.48 D θthio(C-S-C) = 100.0°, dthio(C–S) = 1.88 Å; and (b) t-butylisothiocyanate (5): µ = 6.56 D, θtiso(C–Nα–C) = 176.8°, diso(C-Nα) = 1.44 Å. The major resonance form (closed-shell Lewis structure) is depicted for each.
Marinedrugs 23 00220 g003
Figure 4. Isomerization of allylic isonitriloids, C3H5-NCS and C3H5-N3. (a) Sigmatropic [3] rearrangements: (a) allyl thiocyanate (6a) to allyl isothiocyanate (6b). (b) Degenerate rearrangement of allyl azide (7).
Figure 4. Isomerization of allylic isonitriloids, C3H5-NCS and C3H5-N3. (a) Sigmatropic [3] rearrangements: (a) allyl thiocyanate (6a) to allyl isothiocyanate (6b). (b) Degenerate rearrangement of allyl azide (7).
Marinedrugs 23 00220 g004
Scheme 1. Proposal for the biosynthesis of sponge terpenyl isonitrile (TI), isothiocyanate (RNCS, ITC), and thiocyanate (RSCN, TC) natural products along two axes: major i and minor ii. p is the partition ratio of ITC:TC along ii.
Scheme 1. Proposal for the biosynthesis of sponge terpenyl isonitrile (TI), isothiocyanate (RNCS, ITC), and thiocyanate (RSCN, TC) natural products along two axes: major i and minor ii. p is the partition ratio of ITC:TC along ii.
Marinedrugs 23 00220 sch001
Scheme 2. A unifying proposal for the biosynthesis of 2c, d, k with the intermediate non-classical cation i.
Scheme 2. A unifying proposal for the biosynthesis of 2c, d, k with the intermediate non-classical cation i.
Marinedrugs 23 00220 sch002
Table 1. Second-order rate constant (k2, SN2) of CH3-I with the selected nucleophiles (25 °C) (from Pearson [37]).
Table 1. Second-order rate constant (k2, SN2) of CH3-I with the selected nucleophiles (25 °C) (from Pearson [37]).
EntryNu: or Nu:103·k2/M−1·s−1EntryNu: or Nu:103·k2/M−1·s−1
1MeOH1.3 × 10−77PhO0.073
2NH30.0418NCS0.574
3N30.0789NC0.645
4Br0.079810NCSe9.13
5I3.4211PhS1070
6(CH3)2S0.04512S2O32–114
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Molinski, T.F. The Enigma of Sponge-Derived Terpenoid Isothiocyanate–Thiocyanate Pairs: A Biosynthetic Proposal. Mar. Drugs 2025, 23, 220. https://doi.org/10.3390/md23050220

AMA Style

Molinski TF. The Enigma of Sponge-Derived Terpenoid Isothiocyanate–Thiocyanate Pairs: A Biosynthetic Proposal. Marine Drugs. 2025; 23(5):220. https://doi.org/10.3390/md23050220

Chicago/Turabian Style

Molinski, Tadeusz F. 2025. "The Enigma of Sponge-Derived Terpenoid Isothiocyanate–Thiocyanate Pairs: A Biosynthetic Proposal" Marine Drugs 23, no. 5: 220. https://doi.org/10.3390/md23050220

APA Style

Molinski, T. F. (2025). The Enigma of Sponge-Derived Terpenoid Isothiocyanate–Thiocyanate Pairs: A Biosynthetic Proposal. Marine Drugs, 23(5), 220. https://doi.org/10.3390/md23050220

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop