Abstract
A stereoselective synthesis of anti-1,2-diols has been developed using a multitasking Ru-catalyst in an assisted tandem catalysis protocol. A cyclometalated ruthenium complex catalyzes first a Z-selective cross metathesis of two terminal olefins followed by a stereospecific dihydroxylation. Both steps are catalyzed by Ru, as the Ru-complex is converted to a dihydroxylation catalyst upon addition of NaIO4. A variety of olefins are transformed into valuable highly functionalized and stereodefined molecules. Mechanistic experiments are performed to probe the nature of the oxidation step and catalyst inhibition pathways. These experiments point the way to more broadly applicable tandem catalytic transformations.
Keywords: tandem catalysis, Z-selective, metathesis, dihydroxylation, anti diol
Highly functionalized and stereochemically complex motifs are attractive targets in synthesis due to their diverse molecular interactions in therapeutic and other specialty applications. Efficient synthesis of densely functionalized targets from simple starting materials is thus an important challenge. Assisted tandem catalysis, where coupled catalytic processes are effected by a single catalyst, can significantly increase molecular complexity.[1] Ruthenium metathesis catalysts have frequently been used in tandem reactions since the C–C bond formation step in olefin metathesis can be coupled to a structural elaboration step which introduces additional functionality.[2] In 2006, Blechert2k and Snapper2l demonstrated that cross metathesis using second generation catalysts Ru-1 or Ru-2 followed by Ru-catalyzed dihydroxylation in the presence of NaIO4 as an oxidant led to the corresponding diol (Scheme 1). The dihydroxylation step is highly stereospecific, and thus diol diastereoselectivity is determined by olefin geometry. Since the metathesis occurs under thermodynamic (i.e. substrate) control, primarily E olefins were produced, leading to predominantly syn-diol products. Thus anti-diols,[3] which are important motifs in natural products as well as intermediates in synthesis, are inaccessible by these methods. If a catalyst controlled cross metathesis could be coupled to a dihydroxylation, then anti-diols with predictable and high levels of diastereoselectivity could be accessed. Using this multitasking approach, simple allyl alcohol and allyl amine derivatives could be transformed into valuable densely functionalized products in a catalyst controlled fashion.
Scheme 1.
Tandem metathesis–dihydroxylation. Blechert and Snapper demonstrated that substrate controlled cross metathesis generally leads to syn-diols. We demonstrate that Z-selective catalysts lead to anti-diols in a catalyst controlled fashion via the Z-olefin.
Significant progress has been made in the development of Z-selective olefin metathesis catalysts using Ru,[4–8] Mo[9–11] and W[12] alkylidene complexes.[13] Highly Z-selective cyclometalated Ru complexes (Ru-3 and Ru-4, Figure 1) have been investigated by our group for diverse applications.[14,15] However, these complexes have not been demonstrated to be viable for tandem catalysis, despite the potential to significantly increase the molecular complexity with high stereocontrol in a single-pot sequence.
Figure 1.
Second generation (Ru-1 and Ru-2) and cyclometalated (Ru-3 and Ru-4) ruthenium alkylidene complexes
It is proposed that cyclometalated complexes would be able to catalyze the dihydroxylation of olefins if conditions could be identified to generate a suitably oxidized ruthenium species.2k–2m We anticipated that under acidic aqueous oxidizing conditions, the adamantyl C–Ru bond would be cleaved, generating a species similar to that generated in dihydroxylation with Ru-2.2k Furthermore, we proposed that catalyst controlled Z-selectivity in cross metathesis with Ru-3 or Ru-4 would be translated into high anti-selectivity via a stereospecific pathway. Herein, the successful development of a tandem Z-selective metathesis – dihydroxylation is reported, as well as preliminary mechanistic studies which shed light on catalyst inhibition pathways.
The homodimerization – dihydroxylation of allyl butyrate was examined in order to determine the effect of catalyst and reaction conditions on selectivity (Table 1). The metathesis step was performed under static vacuum conditions, in order to keep the concentration of ethylene in solution low. Shing’s conditions of NaIO4 in 3:3:1 EtOAc:MeCN:H2O[16,17] were used for the dihydroxylation step. Brønsted[18] and Lewis acids[19] have been demonstrated to accelerate dihydroxylation.[20] Second generation complex Ru-2, which is expected to operate under thermodynamic control of olefin geometry, generated the syn diol product with 8:1 selectivity (entry 1). Use of cyclometalated mesityl substituted Ru-3 and diisopropylphenyl substituted Ru-4 generated the desired product 6a in 56% and 68% yield, respectively, with only trace quantities of the syn diol by-product (entries 2 and 3). This anti-selectivity can be attributed to the high Z-selectivity of these catalysts in cross metathesis.[6,21]
Table 1.
Effect of catalyst, additives and conditions on the tandem Z-selective metathesis – dihydroxylation reaction.
| |||||
|---|---|---|---|---|---|
| Entry | Ru | Acid | Changes from standard | Yield 6a (anti)[a] | Yield (syn) |
| 1 | Ru-2 | CeCl3 | none | 5 | 40 |
| 2 | Ru-3 | CeCl3 | none | 56 | 12 |
| 3 | Ru-4 | CeCl3 | none | 68 | 3 |
| 4 | Ru-4 | CeCl3 | No static vacuum | 49 | 4 |
| 5 | Ru-4 | CeCl3 | nBu4NCl (10 mol%) during dihydroxylation | 62 | 5 |
| 6 | Ru-4 | CeCl3 | Ethyl vinyl ether (1 eq) after metathesis step | 47 | 2 |
| 7 | Ru-4 | CeCl3 | 5 mol% Ru-4 | 54 | 7 |
| 8 | Ru-4 | CeCl3 | 30 mol% CeCl3 | 60 | 4 |
| 9 | Ru-4 | None | none | 31 | 1 |
| 10 | Ru-4 | H2SO4 | none | 56 | 5 |
| 11 | Ru-4 | YbCl3 | none | 61 | 3 |
Determined by integration of the crude 1H NMR spectrum using mesitylene as an internal standard.
Achieving high activity and Z-selectivity has been found to depend on the removal of ethylene from solution. Performing the metathesis under static vacuum was critical, as the yield was diminished to 49% when the metathesis step was performed at 1 atm (entry 4).[22] The use of additives, such as nBu4NCl during dihydroxylation, or ethyl vinyl ether after the metathesis reaction, did not improve the yield (entries 5 and 6). Increasing the loading of catalyst Ru-4 (5 mol%) or CeCl3 (30 mol%) resulted in a small decrease in efficiency (54% and 60% yield respectively, entries 7 and 8), while performing the dihydroxylation in the absence of a Lewis acid co-catalyst still resulted in productive dihydroxylation, albeit in only 31% yield (entry 9). Other acids, such as H2SO4 and YbCl3 were also less effective than CeCl3, producing 6a in 56% and 61% yield respectively (entries 10 and 11).
With optimized conditions in hand, we next examined the scope of the Z-selective homodimerization – dihydroxylation. A wide variety of densely functionalized, stereodefined anti-diols could be prepared from comparatively simple starting materials (Table 2). Esters, carbonates, carbamates and amine derivatives were all well tolerated, generating the resulting anti-diol in up to 72% yield. The molecular structure of 6c was determined by X-ray crystallography, supporting the anti-stereochemistry (Figure 2).[23] In addition to probing the overall tandem process, the independent metathesis step was also monitored in each case to ensure high Z-selectivity,[24] since only the homodimerization of allyl acetate has been explored previously with these chelated catalysts.[5]
Table 2.
Tandem Z-selective homodimerization – dihydroxylation of allyl substituted terminal olefins
| ||||
|---|---|---|---|---|
| Entry | R (5) | Product | 6 | Yield (%) |
| 1 | OCOnPr(5a) |
|
6a | 72 |
| 2 | OAc (5b) |
|
6b | 59 |
| 3 | OBz (5c) |
|
6c | 71 |
| 4[a] | OCO2Ph (5d) |
|
6d | 61 |
| 5 | OCONHR (5e) |
|
6e | 63 |
| 6 | OCONHR (5f) |
|
6f | 39 |
| 7 | NHTs (5g) |
|
6g | 70 |
| 8 | NHCBz (5h) |
|
6h | 53 |
Using 1 mol% catalyst in an open vial in the glove box
Figure 2.

POV-ray depiction of the structure of anti-diol 6c determined by X-ray crystallography. Atoms are represented by ellipsoids at the 50% probability level. The crystal was disordered as it contained two conformers – only one has been shown for clarity.
Achieving unsymmetrical substitution patterns via heterocross metathesis – dihydroxylation is an appealing target, particularly if differentially protected products can be obtained. Z-selective heterocross metathesis can be achieved by using an excess of one of the olefin partners.[21,25] Tosyl and Cbz protected allyl amine were used as coupling partners with allyl butyrate or allyl benzoate (Table 2), generating the corresponding substituted amino triols in up to 63% yield. Such orthogonally protected products are valuable building blocks for target oriented synthesis.
We next examined the tandem methodology on gram scale in order to probe scalability of the process. Allyl benzoate was subjected to cross metathesis with 0.5 mol% catalyst Ru-4 in an open vial in an inert atmosphere glove box, followed by dihydroxylation using the standard conditions outside the glove box (Scheme 2). Isolation of the target diol was conveniently achieved without the need for column chromatography: trituration of the crude reaction mixture with ether provided 6c in 66% yield.
Scheme 2.

Gram scale tandem Z-selective metathesis – dihydroxylation
In order to probe the role of Ru in the dihydroxylation step, a series of control experiments were performed. Firstly, Z-2-butenyl 1,4-diacetate 7 was subjected to the standard dihydroxylation conditions in the presence or absence of catalyst Ru-4 (Scheme 3A). Without Ru-4, no conversion was observed, indicating that Ru is a catalyst for both the metathesis and dihydroxylation steps. In the presence of Ru-4 (1 mol%), anti-diol 6b was generated as a single diastereomer, thus confirming the stereospecificity of the dihydroxylation.
Scheme 3.
A) Ruthenium catalyst Ru-4 is required for dihydroxylation. B) Z-4-octene 8 is unreactive in dihydroxylation. C) 8 inhibits the dihydroxylation of 7. Oxidation conditions: NaIO4 (2 eq), CeCl3 (10 mol%), EtOAc:MeCN:H2O (3:3:1), 20 min, 0 °C.
Next, the relative reactivity of electron neutral and deficient internal olefins toward dihydroxylation with Ru-4 was investigated. Z-4-butene 8 was subjected to the standard dihydroxylation conditions with Ru-4 (1 mol%), and no diol was observed (Scheme 3B). Furthermore, when a 1:1 mixture of 8 and Z-2-butenyl 1,4-diacetate 7 was subjected to the same conditions, no diol from either alkene was observed (Scheme 3C). Since 7 is successfully dihydroxylated when it is the only substrate present, this result indicates that 8 is not only unreactive, but also inhibits dihydroxylation of 7. We propose that formation of a stable ruthenate ester from a [3+2] cycloaddition between 8 and a Ru species with at least two oxo ligands sequesters the ruthenium catalyst, making it unavailable for catalysis of dihydroxylation of 7. Hydrolysis of osmate esters is known to be a slow step in the osmium catalyzed dihydroxylation of certain olefins.[26,27] The allylic functional groups could either be acting as electron withdrawing groups to render the Ru center more electrophilic, or as coordinating groups.[28]
In order to probe the inherent reactivity of cross metathesis intermediates containing functionality on only one side of the olefin, we performed the tandem sequence using allyl benzoate and 1-pentene. Under standard conditions, no product was obtained (Scheme 4A). However when the volatiles were removed in vacuo prior to addition of the reagents for dihydroxylation, diol 10 was produced in 33% yield under unoptimized conditions (Scheme 4B). Therefore removal of the inhibitory olefins 4-octene and residual 1-pentene lead to restoration of the dihydroxylation activity, albeit with slightly lower efficiency.[29] This result points the way to expansion of the substrate scope for this tandem transformation.
Scheme 4.

Cross metathesis – dihydroxylation of allyl benzoate and 1-pentene under standard conditions (A) and with removal of volatile intermediates prior to the oxidation step (B).
In summary we have disclosed an assisted tandem catalysis procedure for the Z-selective cross metathesis – dihydroxylation of terminal olefins to yield anti-diols. Ruthenium catalyzes both transformations, and the Z-selectivity observed in the cross metathesis is translated to anti-selectivity via the stereospecific dihydroxylation. Densely functionalized anti-diols with four contiguous heteroatom substituted carbon atoms can be synthesized from simple allyl alcohol and allyl amine derivatives. The behaviour of the in situ generated Ru-based oxidation catalyst was probed with unfunctionalized electron rich alkenes, and these were found to inhibit dihydroxylation. Further studies are ongoing to elucidate details of the reaction mechanism. It is envisioned that this methodology will have applications in target oriented synthesis involving anti-diols, and the mechanistic insights will help to uncover further applications of cyclometalated ruthenium alkylidene catalysts.
Supplementary Material
Table 3.
Z-selective heterocross metathesis – dihydroxylation of allyl substituted terminal olefins
| |||||
|---|---|---|---|---|---|
| Entry | R1 | R2 | Product | 6 | Yield (%) |
| 1 | NHTs | OCOnPr |
|
6i | 63[a] |
| 2 | NHTs | OBz |
|
6j | 39 |
| 3 | NHCbz | OBz |
|
6k | 55 |
| 4 | NHCbz | OCOnPr |
|
6l | 47 |
1.5 mol% Ru-4, 35 °C
Footnotes
This work was financially supported by the ONR (N000141410650), the NIH (5R01GM031332-27) and the NSF (CHE-1212767). Mr. L. M. Henling is thanked for X-ray crystallography. NMR spectra were obtained by instruments supported by the NIH (RR027690). We thank Dr. John Hartung for helpful discussion. Materia, Inc. is thanked for donation of metathesis catalysts.
Supporting information for this article is available on the WWW under http://
References
- 1.a) Fogg DE, dos Santos EN. Coord Chem Rev. 2004;248:2365–2379. [Google Scholar]; b) Ajamian A, Gleason JL. Angew Chem Int Ed. 2004;43:3754–3760. doi: 10.1002/anie.200301727. [DOI] [PubMed] [Google Scholar]
- 2.For metathesis-hydrogenation, see: Louie J, Bielawski CW, Grubbs RH. J Am Chem Soc. 2001;123:11312–11313. doi: 10.1021/ja016431e.Camm KD, Martinez Castro N, Liu Y, Czechura P, Snelgrove JL, Fogg DE. J Am Chem Soc. 2007;129:4168–4169. doi: 10.1021/ja071047o.For metathesis-olefin isomerization, see: Sutton AE, Seigal BA, Finnegan DF, Snapper ML. J Am Chem Soc. 2002;124:13390–13391. doi: 10.1021/ja028044q.Schmidt B. Eur J Org Chem. 2003:816–819.For metathesis-atom transfer radical cyclization see: Schmidt B, Pohler M. J Organomet Chem. 2005;690:5552–5555.Seigal BA, Fajardo C, Snapper ML. J Am Chem Soc. 2005;127:16329–16332. doi: 10.1021/ja055806j.For metathesis-cyclopropanation, see: Kim BG, Snapper ML. J Am Chem Soc. 2006;128:52–53. doi: 10.1021/ja055993l.For metathesis-oxidation, see: Kato H, Ishigame T, Oshima N, Hoshiya N, Shimawaki K, Arisawa M, Shuto S. Adv Synth Catal. 2011;353:2676–2680.Schmidt B, Krehl S. Chem Commun. 2011;47:5879–5881. doi: 10.1039/c1cc11347j.Schmidt B, Krehl S, Hauke S. J Org Chem. 2013;78:5427–5435. doi: 10.1021/jo4005684.For metathesis-dihydroxylation, see: Beligny S, Eibauer S, Maechling S, Blechert S. Angew Chem Int Ed. 2006;45:1900–1903. doi: 10.1002/anie.200503552.Scholte AA, An MH, Snapper ML. Org Lett. 2006;8:4759–4762. doi: 10.1021/ol061837n.Neisius NM, Plietker B. J Org Chem. 2008;73:3218–3227. doi: 10.1021/jo800145x.
- 3.For representative examples of anti-1,2-diol syntheses, see: Kolb HC, VanNieuwenhze MS, Sharpless KB. Chem Rev. 1994;94:2483–2547.Brown HC, Narla G. J Org Chem. 1995;60:4686–4687.Northrup AB, MacMillan DWC. Science. 2004;305:1752–1755. doi: 10.1126/science.1101710.Lim SM, Hill N, Myers AG. J Am Chem Soc. 2009;131:5763–5765. doi: 10.1021/ja901283q.Han SB, Han H, Krische MJ. J Am Chem Soc. 2010;132:1760–1761. doi: 10.1021/ja9097675.Kim D, Lee JS, Kong SB, Han H. Angew Chem Int Ed. 2013;52:4203–4206. doi: 10.1002/anie.201209112.Hartung J, Grubbs RH. Angew Chem Int Ed. 2014;53:3885–3888. doi: 10.1002/anie.201310767.
- 4.Endo K, Grubbs RH. J Am Chem Soc. 2011;133:8525–8527. doi: 10.1021/ja202818v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Keitz BK, Endo K, Patel PR, Herbert MB, Grubbs RH. J Am Chem Soc. 2012;134:693–699. doi: 10.1021/ja210225e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Rosebrugh LE, Herbert MB, Marx VM, Keitz BK, Grubbs RH. J Am Chem Soc. 2013;135:1276–1279. doi: 10.1021/ja311916m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Bronner SM, Herbert MB, Patel PR, Marx VM, Grubbs RH. Chem Sci. 2014;5:4091–4098. doi: 10.1039/C4SC01541J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Koh MJ, Khan RKM, Torker S, Yu M, Mikus MS, Hoveyda AH. Nature. 2015;517:181–186. doi: 10.1038/nature14061. [DOI] [PubMed] [Google Scholar]
- 9.Malcolmson SJ, Meek SJ, Sattely ES, Schrock RR, Hoveyda AH. Nature. 2008;456:933–937. doi: 10.1038/nature07594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Flook MM, Jiang AJ, Schrock RR, Müller P, Hoveyda AH. J Am Chem Soc. 2009;131:7962–7963. doi: 10.1021/ja902738u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Jiang AJ, Zhao Y, Schrock RR, Hoveyda AH. J Am Chem Soc. 2009;131:16630–16631. doi: 10.1021/ja908098t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Peryshkov DV, Schrock RR, Takase MK, Müller P, Hoveyda AH. J Am Chem Soc. 2011;133:20754–20757. doi: 10.1021/ja210349m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Shahane S, Bruneau C, Fischmeister C. ChemCatChem. 2013;5:3436–3459. [Google Scholar]
- 14.For examples of Z-selective cross metathesis, see: Keitz BK, Endo K, Herbert MB, Grubbs RH. J Am Chem Soc. 2011;133:9686–9688. doi: 10.1021/ja203488e.For examples of ROMP, see: Keitz BK, Fedorov A, Grubbs RH. J Am Chem Soc. 2012;134:2040–2043. doi: 10.1021/ja211676y.Rosebrugh LE, Marx VM, Keitz BK, Grubbs RH. J Am Chem Soc. 2013;135:10032–10035. doi: 10.1021/ja405559y.For examples of enantioselective catalysis: Hartung J, Grubbs RH. J Am Chem Soc. 2013;135:10183–10185. doi: 10.1021/ja4046422.Hartung J, Dornan PK, Grubbs RH. J Am Chem Soc. 2014;136:13029–13037. doi: 10.1021/ja506611k.For examples of macrocyclizations, see: Marx VM, Herbert MB, Keitz BK, Grubbs RH. J Am Chem Soc. 2013;135:94–97. doi: 10.1021/ja311241q.
- 15.For mechanistic studies on the origin of Z-selectivity, see: Liu P, Xu X, Dong X, Keitz BK, Herbert MB, Grubbs RH, Houk KN. J Am Chem Soc. 2012;134:1464–1467. doi: 10.1021/ja2108728.Miyazaki H, Herbert MB, Liu P, Dong X, Xu X, Keitz BK, Ung T, Mkrtumyan G, Houk KN, Grubbs RH. J Am Chem Soc. 2013;135:5848–5858. doi: 10.1021/ja4010267.
- 16.Shing TKM, Tai VW-F, Tam EKW. Angew Chem Int Ed Engl. 1994;33:2312–2313. [Google Scholar]
- 17.Shing TKM, Tam EKW, Tai VW-F, Chung IHF, Jiang Q. Chem – Eur J. 1996;2:50–57. [Google Scholar]
- 18.Plietker B, Niggemann M. Org Lett. 2003;5:3353–3356. doi: 10.1021/ol035335a. [DOI] [PubMed] [Google Scholar]
- 19.Plietker B, Niggemann M. J Org Chem. 2005;70:2402–2405. doi: 10.1021/jo048020x. [DOI] [PubMed] [Google Scholar]
- 20.It has been proposed that acids facilitate hydrolysis of a ruthenate ester intermediate.
- 21.Quigley BL, Grubbs RH. Chem Sci. 2013;5:501–506. doi: 10.1039/c3sc52806e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Presumably a larger concentration of ethylene in solution results in slower productive metathesis and more secondary metathesis processes.
- 23.CCDC 1048930 contains the supplementary crystallographic data for compound 6c. This data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
- 24.In all cases, over 90% Z-selectivity was obtained.
- 25.Cannon JS, Grubbs RH. Angew Chem Int Ed. 2013;52:9001–9004. doi: 10.1002/anie.201302724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Sharpless KB, Amberg W, Bennani YL, Crispino GA, Hartung J, Jeong KS, Kwong HL, Morikawa K, Wang ZM. J Org Chem. 1992;57:2768–2771. [Google Scholar]
- 27.Junttila MH, Hormi OEO. J Org Chem. 2007;72:2956–2961. doi: 10.1021/jo070011b. [DOI] [PubMed] [Google Scholar]
- 28.Further studies are ongoing to distinguish these possibilities For directed dihydroxylation see: Donohoe TJ. Synlett. 2002:1223–1232.
- 29.Some olefin intermediate remained after the 20 minute oxidation, resulting in low yield. Therefore dihydroxylation to give diol 10 is likely slower than formation of 6c.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



