Synlett 2019; 30(17): 1988-1994
DOI: 10.1055/s-0039-1690992
cluster
© Georg Thieme Verlag Stuttgart · New York

Trithioorthoester Exchange and Metathesis: New Tools for Dynamic Covalent Chemistry

Michael Bothe
a   Institute of Organic Chemistry I, University of Ulm, Albert-Einstein-Allee 11, 89081 Ulm, Germany   Email: max.vondelius@uni-ulm.de
,
A. Gastón Orrillo
b   Farmacognosia, Facultad de Ciencias Bioquímicas y Farmacéuticas, Universidad Nacional de Rosario-CONICET, S2002LRK Rosario, Argentina   Email: rfurlan@fbioyf.unr.edu.ar
,
Ricardo L. E. Furlan
b   Farmacognosia, Facultad de Ciencias Bioquímicas y Farmacéuticas, Universidad Nacional de Rosario-CONICET, S2002LRK Rosario, Argentina   Email: rfurlan@fbioyf.unr.edu.ar
,
a   Institute of Organic Chemistry I, University of Ulm, Albert-Einstein-Allee 11, 89081 Ulm, Germany   Email: max.vondelius@uni-ulm.de
› Author Affiliations
This work was supported by the Deutsche Forschungsgemeinschaft (Emmy-Noether grant DE1830/2-1), Fondo para la Investigación Científica y Tecnológica (FONCYT) (PICT2015-3574) and a CONICET – BAYLAT Bilateral Cooperation Programme, Level I, Res. 733/15.
Further Information

Publication History

Received: 06 September 2019

Accepted: 11 September 2019

Publication Date:
18 September 2019 (online)

 


Published as part of the Cluster Metathesis beyond Olefins

Abstract

To expand the toolbox of dynamic covalent and systems chemistry, we investigated the acid-catalyzed exchange reaction of trithioorthoesters with thiols. We found that trithioorthoester exchange occurs readily in various solvents in the presence of stoichiometric amounts of strong Brønsted acids or catalytic amounts of certain Lewis acids. The scope of the exchange reaction was explored with various substrates, and conditions were identified that permit clean metathesis reactions between two different trithioorthoesters. One distinct advantage of S,S,S-orthoester exchange over O,O,O-orthoester exchange is that the exchange reaction can kinetically outcompete hydrolysis, thereby making the process less sensitive to residual moisture. We expect that the relatively high stability of the products might be beneficial in future supramolecular receptors or porous materials.


#

Dynamic covalent chemistry (DCC)[1] has emerged in the last two decades as an area that combines the best attributes of organic chemistry (synthesis of stable compounds) with those of supramolecular chemistry (error correction).[2] At the heart of DCC are robust and reliable chemical reactions that, at least in the presence of suitable catalysts,[3] lead to the formation of equilibrium mixtures under relatively mild conditions. Typical examples include the exchange of disulfides with thiols,[4] and the reversible condensation reactions of imines,[5] hydrazones,[6] and oximes.[7] However, in the light of new applications of DCC, especially in the synthesis of porous materials[8] and in the life sciences,[9] there is unabated interest in the development of new dynamic covalent reactions.

To this end, the groups of von Delius and of Furlan recently reported investigations of O,O,O-orthoester exchange[10] and dithioacetal exchange reactions,[11] respectively (Scheme [1]). Both reactions share similarities, such as a requirement for acid catalysis in organic solvents[12] and their widespread use in protective-group chemistry.[13] The tripodal nature of orthoesters makes them uniquely suited for the self-assembly of cage-type architectures.[14] [15] Also, orthoester exchange gives rise to a remarkable level of molecular diversity, because by mixing one orthoester with one alcohol, an equilibrium mixture consisting of four different orthoesters is obtained (Scheme [1]A). In contrast, dithioacetal exchange produces fewer products (Scheme [1]B) and is more suited to the preparation of cyclic hosts.[16] This exchange can be connected to that of disulfides and thioesters to generate multilevel dynamic systems.[17] Compared with orthoesters, dithioacetals are less susceptible to hydrolysis and they demand more-acidic media for exchange.

Zoom Image
Scheme 1 Exchange reactions of (A) orthoesters, (B) dithioacetals, and (C) trithioorthoesters

In this study, we focus on trithioorthoester exchange, an area of chemical space that, from the perspective of topology and reactivity, lies between orthoester exchange and dithioacetal exchange (Scheme [1]C). Trithioorthoesters[18] can be obtained from, inter alia, O,O,O-orthoesters,[19] chloroform,[20] orthothioformates,[21] or dithioacetals,[22] [23] and trithioorthoester groups have been used primarily as protective groups, especially when the product is required to be more stable toward acid hydrolysis than its oxygen counterpart.[13] Here, we present a comprehensive investigation of the conditions required for trithioorthoester exchange and for trithioorthoester metathesis, in the hope that these transformations will prove useful in areas where the advantages of thermodynamic control can be harnessed.[8e,24]

We started by testing the feasibility of trithioorthoester exchange under conditions similar to those used for the activation of O,O,O-orthoesters[10] or dithioacetals.[11] To this end, trifluoroacetic acid (TFA; 1 or 10 equiv) was added to a chloroform-d solution containing tris(phenylthio)methane (A13 ; 50 mM, 1 equiv) and phenylmethanethiol (2; 3 equiv) at room temperature (Table [1]). All solvents were dried over molecular sieves before use to minimize the irreversible hydrolysis of trithioorthoesters to thioesters. The compositions of the reaction mixtures were analyzed by 1H NMR spectroscopy after one hour and 24 hours. The use of ten equivalents of TFA led to equilibration of the mixture within one hour of reaction time. The apparent bias toward trithioorthoesters rich in building block 2 (Table [1], entry 1) can be rationalized in terms of differences in steric demand. The undesired hydrolysis reaction led to ≤2% of thioesters A1 and A2 with respect to the initial trithioorthoesters. A tenfold decrease in the amount of TFA was possible, but led to slower exchange, as indicated by the dominance of starting material A13 after 24 hours of reaction time (Table [1], entry 2). The harsh Brønsted acidic conditions, which are required to equilibrate trithioorthoesters within a reasonable timespan, are similar to those giving rise to dithioacetal exchange, whereas O,O,O-orthoester exchange proceeds under much milder conditions.

Table 1 Scope of Trithioorthoester Exchangea

Entry

Acid (Equiv)

Solvent

Reaction time (h)

A13 /A122/A122 /A23 /(A1A2)b

 1

TFA (10)

CDCl3

 1
24

3:17:38:42:
2:16:38:43:1

 2

TFA (1)

CDCl3

 1
24

97:3:–:–:–
46:30:16:8:–

 3

TFA (10)

C6D6

 1
24

8:17:33:42:–
 3:17:39:41:–

 4

TFA (10)

CD3CN

 1
24

98:2:–:–:–
66:22:8:4:–

 5

MsOH (1)

CD3CN

 1
24

97:3:–:–:–
56:26:13:5:–

 6

PTSA (1)

CD3CN

 1
24

60:24:1:5:–
2:11:37:50:–

 7

H2SO4 (1 )

CD3CN

 1
24

73:18:6:3:–
4:14:39:43:–

 8

TfOH (1)

CD3CN

 1
24

1:18:42:39:–
 2:19:43:36:–

 9

FeCl3 (0.1)

CD3CN

 1
24

35:26:23:16:–
3:15:39:43:–

10

AlCl3 (0.1)

CD3CN

 1
24
48

92:8:–:–:–
24:21:26:29:–
14:15:29:42:–

11

BF3·OEt2 (0.1)

CD3CN

 1
24
48

78:18:4:–:–
9:14:33:44:–
 4:15:38:43:–

12

FeCl3 (0.1)

CDCl3

 1

3:17:40:40:–

a Reaction conditions: A13 (37.5 μmol), 2 (112.5 μmol), TFA (375 μmol), CDCl3 (V total: 750 μL), r.t.

b Product ratios were determined by integration of 1H NMR signals. Equilibrium mixtures are highlighted in bold face; ‘–’ indicates not detected. For full spectra, see Figures S4–S25 in the Supplementary Information.

A range of parameters was next investigated to better understand the reaction. First, various solvents were tested by using TFA in a standard amount of ten equivalents. In benzene-d 6, equilibrium was reached after one hour of reaction (Table [1], entry 3), whereas in acetonitrile-d 3 the exchange was slow and the composition after 24 hours was still far from equilibrium (Table [1], entry 4). No exchange products were observed in DMSO-d 6 or THF-d 8. When stronger Brønsted acids were compared in acetonitrile-d 3 (for solubility reasons), a correlation between the pK a value[25] and the reaction kinetics was observed. Slow exchange was observed with TFA (pK a = 12.7) or methanesulfonic acid (pK a = 10.0) (entries 4 and 5), whereas the mixtures with p-toluenesulfonic acid (pK a = 8.0) or sulfuric acid (pK a = 7.2) equilibrated after 24 hours (entries 6 and 7). Only the mixture with trifluoromethanesulfonic acid (pK a = 2.6) equilibrated after one hour of reaction (entry 8). Trifluoromethanesulfonic acid turned out to be a less effective catalyst in DMSO-d 6, whereas THF-d 8 [26] was unstable in the presence of this acid [see the Supplementary Information (SI)]. Stoichiometric or excess Brønsted acids were shown to be useful for promoting trithioorthoester exchange. These conditions were used in exchange experiments initiated from various starting materials,[1c] [27] which confirmed the reversibility of the reaction (Figures S1–S3, SI).

In the hope of identifying milder and truly catalytic (substoichiometric) conditions, we proceeded to investigate some representative Lewis acids.[28] For reasons of solubility, we chose CD3CN as a solvent to investigate FeCl3, AlCl3, and BF3·OEt2, whereas CDCl3 was used in combination with ­SnCl4, TiCl4, and FeCl3. We found that one equivalent of each of the Lewis acids FeCl3, AlCl3, SnCl4, and BF3·OEt2 led to equilibration after roughly one hour; experiments with 0.1 equivalent of these Lewis acids showed that they can indeed be regarded as catalysts for this reaction (Table [1], entries 9–12) with the following reactivity trend: FeCl3 > BF3·OEt2 > AlCl3. In addition, the use of Lewis acids in CDCl3 was found to be promising. When 0.1 equivalent of FeCl3 in CDCl3 was tested, equilibrium was attained after one hour (entry 12). Although the equilibrium position seems to be unaffected by the presence of the iron cation, signal broadening was observed, which increased with reaction time, preventing accurate integration after 24 hours of reaction. These results show that Lewis acids require a lower catalyst loading than do Brønsted acids. Future work might focus on Lewis acid catalysis in CDCl3 to achieve shorter reaction times when using substoichiometric amounts of catalyst. To this end, alternative analytical tools such as HPLC might be helpful in avoiding interference by the paramagnetic effects induced by metal cations.

To explore the scope of the reaction, we studied the effects of the nucleophile on the kinetics and thermodynamics of exchange. To this end, trithioorthoester A13 was combined with thiols 26 in the presence of TFA (10 equiv) (Scheme [2]). Primary thiols such as phenylmethanethiol (2) or hexane-1-thiol (3) gave compositions shifted toward trithioorthoester products containing more-exchanged units. Secondary and aromatic thiols, such as propane-2-thiol (4) and 4-methylbenzenethiol 5, respectively, led to nearly symmetrical statistical distributions of trithioorthoesters. The use of bulky adamantane-1-thiol (6) led to the formation of only a single exchange product containing one building block 6. The exchange kinetics were not noticeably affected by thiol structures, because all the mixtures were equilibrated after one hour of reaction time. Comparable equilibrium distributions were obtained after 24 hours when A13 was exposed to the nucleophiles in CD3CN in the presence of FeCl3 (0.1 equiv) as catalyst (see SI). These results are in agreement with those observed in O,O,O-orthoester exchange, in which the equilibrium position, but not the equilibration kinetics, depends on the size of the nucleophile.[10] Finally, inspired by recent studies on (crossed) dichalcogenide exchange reactions,[29] we investigated the reaction of trimethyl trithioorthoformate A83 with benzeneselenol 7. After one hour, four signals in the diagnostic trithioorthoester range were observed; these appeared in a statistical proportion, showing the feasibility and the reversibility of the reaction (after 24 hours, selenol oxidation dominates the reaction outcome). To our knowledge, this is the first example of selenol/trithioorthoester exchange.

Zoom Image
Scheme 2 Effect of the structure of the nucleophile on the distribution of trithioorthoesters at equilibrium. Reaction conditions: A13 or A83 (37.5 μmol), nucleophile (112.5 μmol), TFA (375 μmol), CDCl3 (V total: 750 μL), r.t. Product ratios were determined by integration of 1H NMR signals after 1 h. Hydrolysis was negligible (<2%) in all cases except for thiol 6 (7%). a Exchange performed with A13 (CDCl3 or CD3CN). b Exchange performance with A83 (CDCl3). For full spectra, see Figures S26–S36 (SI).

While keeping the ratio of trithioorthoester A13 to thiol 2 at 1:3, and with a constant total amount of TFA, we next investigated the effect of the trithioorthoester concentration (Figure S37). We found that a tenfold decrease in the concentration of A13 from 50 mM to 5 mM did not affect the equilibrium position. This indicates that trithioortho­ester exchange might well be suited to experiments requiring low concentrations of reactants. However, a simultaneous decrease in the concentration of TFA to 10 equivalents (50 mM) slowed the exchange reaction considerably (96% of A13 was present after 1 h).

To take advantage of the low boiling point of methane­thiol (6 °C), A83 was treated with 4-methylbenzenethiol (5) in the presence of excess TFA (10 equiv) under a smooth nitrogen stream while the temperature was increased from r.t. to 40 °C and then to 60 °C to shift the equilibrium completely towards trithioorthoesters containing exchanged side chains.[30] The composition could indeed be shifted almost completely toward the formation of exchange product A53 (Figure S38). This straightforward method might be useful whenever a complete shift in the equilibrium toward the product side is desired.

In a comparative study, we examined the hydrolytic stability of trimethyl trithioorthoformate [(S,S,S)-A83 ], its oxygen-containing counterpart trimethyl orthoformate [(O,O,O)-A93 ], and the dithioacetal bis(methylthio)methane (AH82 ).[31] Increasing amounts of TFA were added to solutions containing each compound with one equivalent of water, and the samples were analyzed by 1H NMR spectroscopy after one hour of reaction (Figure [1]; note the logarithmic scale of the x-axis).

Zoom Image
Figure 1 Effects of increasing amounts of TFA on the hydrolysis of ­orthoester (O,O,O)-A93 , trithioorthoester (S,S,S)-A83 , and dithioacetal AH82 . Reaction conditions: A93 or A83 or AH82 (37.5 μmol), increasing amounts of TFA [from 37.5 nmol (50 μM) to 3.75 mmol (5.0 M)], CDCl3 (V total: 750 μL), r.t., 1 h; internal standard: toluene. For full spectra, see Figures S39–S56 in the SI.

As expected, the trithioorthoester compound was considerably more stable to hydrolysis than was the orthoester, but was less stable than the dithioacetal. We found that the addition of ten equivalents of TFA led to complete hydrolysis of (O,O,O)-A93 , whereas (S,S,S)-A83 and AH82 were mostly stable. The addition of 100 equivalents of TFA was necessary to hydrolyze(S,S,S)-A83 completely, whereas AH82 remained stable under these conditions. These findings indicate that in terms of hydrolytic stability, the trithioorthoester compound is located at an intermediate position between the orthoester and dithioacetal compounds. Most importantly, the results point towards faster exchange versus slower hydrolysis kinetics of trithioorthoesters: over a one hour period, treatment of (S,S,S)-A83 with ten equivalents of TFA led to complete equilibration (Figure S57), but only to 30% hydrolysis. This observation explains why the rigorous exclusion of water is not as essential for (S,S,S)-orthoester exchange as it is for the (O,O,O)-orthoester variant. When trimethyl trithioorthoacetate (C83 ) was hydrolyzed with ten equivalents of TFA, the hydrolysis ratio after one hour was 93%, indicating that the same electronic effects as previously reported account for differences in hydrolytic stability (Figure S58).[15e] The observed differential susceptibility to hydrolysis of orthoesters, trithioorthoesters, and dithioacetals is relevant for the selective removal of protective groups.

In light of the shuttle catalysis concept,[32] there has recently been an increase in interest in exchange reactions between molecules having the same kind of functional group, i.e. Type 1 metathesis.[33] Among the Type 1 metathesis reactions that have been studied from the perspective of dynamic covalent/combinatorial chemistry are disulfide,[4] [34] trithiocarbonate,[35] thiazolidine,[36] acetal,[37] orthoester,[10] and dithioacetal exchange.[17a] We therefore wondered whether a direct metathesis reaction between two trithioorthoesters might be possible.

To answer this question, a set of trithioorthoesters suitable for crossover experiments were synthesized.[38] [39] [40] [41] Treatment of compounds A13 and B83 with TFA (10 equiv) led to a Type I metathesis, as indicated by the appearance in the 1H NMR spectrum of a set of signals corresponding to the eight expected trithioorthoesters (Figure [2] and Table [2], entry 1). As in the case of related orthoester metathesis,[10] we believe that the generation of small quantities of free thiol[42] is responsible for the observed reactivity.

Zoom Image
Figure 2 Trithioorthoester metathesis between A13 and B83 (top) and 1H NMR spectrum measured after one hour (bottom). The two dashed boxes show the corresponding formate and acetoxy hydrogen atoms of trithioorthoesters formed from A13 and B83 . Reaction conditions: A13 (37.5 μmol), B83 (37.5 μmol), TFA (375 μmol), CDCl3 (V total: 750 μL), r.t., 1 h. Product distributions are given by normalized integral values below the corresponding peaks for both sets. See also Table [2], entry 1 and the full spectra in Figure S59 in the SI.

Table 2 Metathesis between Different Trithioorthoesters

Entry

Substrate 1
HR1 3

Substrate 2
XR2 3

Reaction outcome
HR1 3/HR1 2R2/HR1R2 2/HR2 3/thioesters (hydrolysis)a

1b

A13

B83

17:43:33:7:–

2c

A13

A53

13:36:36:13:2

3c

A53

A(10)3

13:35:36:13:3

4d

A13

A83

10:36:37:11:6

5b

A13

A(11)3

12:34:33:10:12

6c

A53

A43

18:28:30:12:12

7b

A(11)3

C13

3:21:42:28:6

8b

A(11)3

D13

ether cleavage

a Product ratios were determined by integration of 1H NMR signals after 1 h of reaction; ‘–’ indicates not detected. For full spectra, see Figures S60–S67 in the SI.

b Reaction conditions: HR1 3 (37.5 μmol), XR2 3 (37.5 μmol), TFA (375 μmol), CDCl3 (V total: 750 μL), r.t.

c Reaction conditions: HR1 3 (15 μmol), XR2 3 (15 μmol,), H2SO4-saturated CDCl3, (V total: 500 μL), r.t.

d Reaction conditions: HR1 3 (37.5 μmol), XR2 3 (37.5 μmol), FeCl3 (3.75 μmol), CD3CN (V total: 750 μL), r.t.

To explore the generality of this finding, additional metathesis reactions were carried out with other pairs of trithioorthoesters. Combinations of trithioorthoesters containing different aromatic substituents on sulfur (Table [2], entry 2) or aromatic and primary alkyl substituents (entries 3–5) led to statistical distributions, whereas trithioorthoesters containing aromatic and secondary thiol side chains (entry 6) led to a slightly biased composition favoring the starting aromatic trithioorthoester. These experiments can also be regarded as competitive hydrolysis experiments: throughout our investigations, we found that hydrolysis of trithioorthoesters with alkyl residues on sulfur is faster than that for aromatic ones (compare Figure S60 with Figures S61–S64; SI). Next, we investigated the influence of electron-withdrawing and electron-donating substituents on trithioorthoester metathesis by conducting metathesis reactions between tris(ethylthio)methane [A(11)3 ] and trithioorthoesters C13 and D13 , containing electron-withdrawing and electron-donating groups, respectively (Figures S65 and S66). After one hour, similar equilibrium distributions for the formate species and a comparable degree of hydrolysis were observed in both cases (entries 7 and 8). However, in the case of D13 , degradation of the starting material, presumably by ether cleavage[43] occurred, representing a notable limitation of this method. Finally, several attempts were made to carry out crossed metathesis reactions between (S,S,S)-A13 and (O,O,O)-A93 ; these were unsuccessful due to instantaneous hydrolysis of O,O,O-orthoester, further confirming the previously observed differences in stability (Figure S67).

In summary, we have described a methodological investigation of the exchange reaction between trithioortho­esters and thiols, as well as the direct trithioorthoester metathesis reaction. These reactions have two appealing properties in the context of other reversible covalent reactions. First, the tripodal structure of S,S,S-orthoesters provides an elegant entry to sulfur-rich three-dimensional architectures with possible applications in the removal of heavy-metal ions. Second, the harsh conditions necessary to initiate these exchange reactions might be advantageous for applications in materials science, such as the (solvothermal) synthesis of porous materials.


#

Acknowledgment

We acknowledge the assistance of Margit Lang with elemental analysis.

Supporting Information

  • References and Notes

    • 1a Lehn J.-M. Angew. Chem. Int. Ed. 2015; 54: 3276
    • 1b Jin Y, Wang Q, Taynton P, Zhang W. Acc. Chem. Res. 2014; 47: 1575
    • 1c Corbett PT, Leclaire J, Vial L, West KR, Wietor J.-L, Sanders JK. M, Otto S. Chem. Rev. 2006; 106: 3652
    • 1d Furlan RL. E, Otto S, Sanders JK. M. Proc. Natl. Acad. Sci. U. S. A. 2002; 99: 4801
    • 2a Lehn J.-M. Top. Curr. Chem. 2011; 322: 1
    • 2b Lehn J.-M. Chem. Soc. Rev. 2007; 36: 151
  • 3 Kölmel DK, Kool ET. Chem. Rev. 2017; 117: 10358
    • 4a Black SP, Sanders JK. M, Stefankiewicz AR. Chem. Soc. Rev. 2014; 43: 1861
    • 4b Fritze UF, von Delius M. Chem. Commun. 2016; 52: 6363
  • 5 Belowich ME, Stoddart JF. Chem. Soc. Rev. 2012; 41: 2003
    • 6a Beeren SR, Sanders JK. M. J. Am. Chem. Soc. 2011; 133: 3804
    • 6b Lam RT. S, Belenguer A, Roberts SL, Naumann C, Jarrosson T, Otto S, Sanders JK. M. Science 2005; 308: 667
    • 6c Furlan RL. E, Ng Y.-F, Otto S, Sanders JK. M. J. Am. Chem. Soc. 2001; 123: 8876
  • 7 Ulrich S, Boturyn D, Marra A, Renaudet O, Dumy P. Chem. Eur. J. 2014; 20: 34
    • 8a Mastalerz M. Acc. Chem. Res. 2018; 51: 2411
    • 8b Ong WJ, Swager TM. Nat. Chem. 2018; 10: 1023
    • 8c Guan X, Li H, Ma Y, Xue M, Fang Q, Yan Y, Valtchev V, Qiu S. Nat. Chem. 2019; 11: 587
    • 8d Ono K, Iwasawa N. Chem. Eur. J. 2018; 24: 17856
    • 8e Ding H, Chen R, Wang C. In Dynamic Covalent Chemistry: Principles, Reactions, and Applications . Zhang W, Jin Y. Wiley; New York: 2018: 165
    • 8f Feng X, Ding X, Jiang D. Chem. Soc. Rev. 2012; 41: 6010
    • 9a Frei P, Hevey R, Ernst B. Chem. Eur. J. 2019; 25: 60
    • 9b Liu Y, Lehn J.-M, Hirsch AK. H. Acc. Chem. Res. 2017; 50: 376
    • 9c Mondal M, Hirsch AK. H. Chem. Soc. Rev. 2015; 44: 2455
  • 10 Brachvogel R.-C, von Delius M. Chem. Sci. 2015; 6: 1399
  • 11 Orrillo AG, Escalante AM, Furlan RL. E. Chem. Eur. J. 2016; 22: 6746
  • 12 Brachvogel R.-C, von Delius M. Eur. J. Org. Chem. 2016; 3662
    • 13a Greene TW, Wuts PG. M. Protective Groups in Organic Synthesis, 2nd ed. Wiley; New York: 1991
    • 13b Kocienski PJ. Protecting Groups, 3rd ed. Thieme; Stuttgart: 1994
  • 14 Collins MS, Shear TA, Smith EK, Strain SM, Zakharov LN, Johnson DW. Chem. Eur. J. 2019; DOI: 10.1002/chem.201903854.
    • 15a Brachvogel R.-C, Hampel F, von Delius M. Nat. Commun. 2015; 6: 7129
    • 15b Brachvogel R.-C, Maid H, von Delius M. Int. J. Mol. Sci. 2015; 16: 20641
    • 15c Shyshov O, Brachvogel R.-C, Bachmann T, Srikantharajah R, Segets D, Hampel F, Puchta R, von Delius M. Angew. Chem. Int. Ed. 2017; 56: 776
    • 15d Wang X, Shyshov O, Hanževački M, Jäger CM, von Delius M. J. Am. Chem. Soc. 2019; 141: 8868
    • 15e Löw H, Mena-Osteritz E, von Delius M. Chem. Sci. 2018; 9: 4785
  • 16 Orrillo AG, Escalante AM, Furlan RL. E. Org. Lett. 2017; 19: 1446
  • 18 Lebel H, Grenon M. In Science of Synthesis . Vol. 22; Charette AB. Thieme; Stuttgart: 2005: 749
    • 19a Seebach D, Geiß K.-H, Beck AK, Graf B, Daum H. Chem. Ber. 1972; 105: 3280
    • 19b Barsamian AL, Blakemore PR. Organometallics 2012; 31: 19
  • 20 Cao Y.-J, Lai Y.-Y, Cao H, Xing X.-N, Wang X, Xiao W.-J. Can. J. Chem. 2006; 84: 1529
  • 21 Seebach D. Angew. Chem. 1967; 79: 468
    • 22a Tlach BC, Tomlinson AL, Ryno AG, Knoble DD, Drochner DL, Krager KJ, Jeffries-EL M. J. Org. Chem. 2013; 78: 6570
    • 22b Corey EJ, Seebach D. Angew. Chem. 1965; 77: 1134
    • 22c Gröbel B.-T, Seebach D. Synthesis 1977; 357

      For syntheses of dithioacetals see:
    • 23a Firouzabadi H, Iranpoor N, Hazarkhani H. J. Org. Chem. 2001; 66: 7527
    • 23b Tazaki M, Takagi M. Chem. Lett. 1979; 8: 767
    • 24a Moulin E, Cormos G, Giuseppone N. Chem. Soc. Rev. 2012; 41: 1031
    • 24b Li J, Nowak P, Otto S. J. Am. Chem. Soc. 2013; 135: 9222
    • 24c Jin Y, Yu C, Denman RJ, Zhang W. Chem. Soc. Rev. 2013; 42: 6634
    • 25a Eckert F, Leito I, Kaljurand I, Kütt A, Klamt A, Diedenhofen M. J. Comput. Chem. 2009; 30: 799
    • 25b Kütt A, Selberg S, Kaljurand I, Tshepelevitsh S, Heering A, Darnell A, Kaupmees K, Piirsalu M, Leito I. Tetrahedron Lett. 2018; 59: 3738
    • 25c Paenurk E, Kaupmees K, Himmel D, Kütt A, Kaljurand I, Koppel IA, Krossing I, Leito I. Chem. Sci. 2017; 8: 6964
  • 26 Pruckmayr G, Wu TK. Macromolecules 1978; 11: 662
    • 27a Otto S, Furlan RL. E, Sanders JK. M. J. Am. Chem. Soc. 2000; 122: 12063
    • 27b Delius M, Geertsema EM, Leigh DA, Slawin AM. Org. Biomol. Chem. 2010; 8: 4617
    • 28a Schmittel M, Levis M. Synlett 1996; 315
    • 28b Kamal A, Laxman E, Reddy PS. M. M. Synlett 2000; 1476
    • 28c Gauthier JY, Martins EO, Young RN, Zamboni RJ. Synlett 2002; 984
    • 28d Kumar V, Dev S. Tetrahedron Lett. 1983; 24: 1289
    • 28e Patney HK. Tetrahedron Lett. 1991; 32: 2259
    • 28f Tamami B, Borujeny KP. Synth. Commun. 2003; 33: 4253
    • 29a Rasmussen B, Sørensen A, Gotfredsen H, Pittelkow M. Chem. Commun. 2014; 50: 3716
    • 29b Ji S, Cao W, Yu Y, Xu H. Angew. Chem. Int. Ed. 2014; 53: 6781
    • 29c Chuard N, Poblador-Bahamonde AI, Zong L, Bartolami E, Hildebrandt J, Weigand W, Sakai N, Matile S. Chem. Sci. 2018; 9: 1860
    • 29d Bartolami E, Basagiannis D, Zong L, Martinent R, Okamoto Y, Laurent Q, Ward TR, Gonzalez-Gaitan M, Sakai N, Matile S. Chem. Eur. J. 2019; 25: 4047
    • 29e Steinmann D, Nauser T, Koppenol WH. J. Org. Chem. 2010; 75: 6696
    • 31a Cordes EH, Bull HG. Chem. Rev. 1974; 74: 581
    • 31b Chiang Y, Kresge AJ, Lahti MO, Weeks DP. J. Am. Chem. Soc. 1983; 105: 6852
    • 31c Repetto SL, Costello JF, Butts CP, Lam JK. W, Ratcliffe NM. Beilstein J. Org. Chem. 2016; 12: 1467
    • 31d Deslongchamps P, Dory YL, Li S. Tetrahedron 2000; 56: 3533
    • 32a Bhawal BN, Morandi B. ACS Catal. 2016; 6: 7528
    • 32b Bhawal BN, Morandi B. Chem. Eur. J. 2017; 23: 12004
  • 33 Bhawal BN, Morandi B. Angew. Chem. Int. Ed. 2019; 58: 10074
    • 34a Otsuka H, Nagano S, Kobashi Y, Maeda T, Takahara A. Chem. Commun. 2010; 46: 1150
    • 34b Caraballo R, Sakulsombat M, Ramström O. Chem. Commun. 2010; 46: 8469
  • 35 Amamoto Y, Kamada J, Otsuka H, Takahara A, Matyjaszewski K. Angew. Chem. Int. Ed. 2011; 50: 1660
  • 36 Saiz C, Wipf P, Manta E, Mahler G. Org. Lett. 2009; 11: 3170
  • 37 Cacciapaglia R, Di Stefano S, Mandolini L. J. Am. Chem. Soc. 2005; 127: 13666
  • 38 Tris[(4-methylbenzene)thio]methane (A53 ), tris(isopropylthio)methane (A43 ), and tris[(2-phenylethyl)thio]methane [A(10)3 ] were synthesized by a modified version of the reported procedure; see Ref. 20. Tris[(4-methylbenzene)thio]methane (A53 ); Typical Procedure CHCl3 (1.48 g, 12.4 mmol, 1.0 equiv), DBU (9.79 g, 64.3 mmol, 5.2 equiv), and 4-methylbenzenethiol (6.14 g, 49.4 mmol, 4.0 equiv) were dissolved in anhyd THF (10 mL), and the mixture was heated to 100 °C for 24 h under N2 in a pressure tube. The mixture was cooled, H2O (30 mL) was added, and the resulting mixture was extracted with Et2O (2 × 50 mL). The combined organic phases were washed with brine (50 mL), dried (Na2SO4), filtered, and concentrated on a rotary evaporator. The residue was purified by column chromatography [silica gel, PE–CH2Cl2 (100:0 to 80:20)]. Subsequent crystallization (CH2Cl2) gave A53 as a white solid; yield: 2.10 g (44%, 5.48 mmol); Rf = 0.3 (silica gel, PE–CH2Cl2, 80:20, UV254). 1H NMR (400 MHz, 298 K, CDCl3): δ = 7.43–7.41 (m, 6 H), 7.17–7.15 (m, 6 H). 5.32 (s, 1 H), 2.37 (s, 9 H). 13C NMR (100 MHz, 298 K, CDCl3): δ = 138.5, 133.4, 130.4, 129.6, 65.9, 21.2. Anal. calcd for C22H22S3: C, 69.07; H, 5.80; S, 25.14. Found: C, 69.00; H, 5.93; S, 25.22.
  • 39 Synthesis of 1,1,1-Tris(methylthio)ethane (B83) This was synthesized by a modified version of the reported procedure.21 Trimethyl trithioorthoformate (A83 ; 5.28 g, 34.2 mmol, 1.0 equiv) was dissolved in anhyd THF (50 mL) and cooled to –78 °C under N2. A 2.5 M solution of BuLi in hexane (20.5 mL, 51.3 mmol, 1.5 equiv) was added dropwise and the mixture was stirred for 80 min. MeI (12.13 g, 85.5 mmol, 2.5 equiv) was added dropwise, and the mixture was allowed to slowly warm to r.t. overnight. Et2O (100 mL) was added and the mixture washed with aq Na2S2O3 (50 mL). The combined extracted organic phases were washed with brine (50 mL), separated, dried (Na2SO4), filtered, and concentrated on a rotary evaporator. Distillation through a short Vigreux column (oil-bath temperature: 120 °C; pressure: 7 mbar; bp 70 °C) gave a colorless oil; yield: 4.55 g (79%, 27.0 mmol); R f = 0.7 (silica gel, PE–EtOAc, 95:5, UV254). 1H NMR (400 MHz, 298 K, CDCl3): δ = 2.16 (s, 9 H), 1.87 (s, 3 H). 13C NMR (100 MHz, 298 K, CDCl3): δ = 64.3, 28.3, 13.6. Anal. calcd for C5H12S3: C, 35.68; H, 7.19; S, 57.14. Found: C, 37.04; H, 7.26; S, 57.81.
  • 40 1-[Bis(phenylthio)methyl]-4-fluorobenzene (CH12 ) and 1-[bis(phenylthio)methyl]-4-methoxybenzene (DH12 ) were synthesized by a modified version of the reported procedure; see Ref 23a. 1-[Bis(phenylthio)methyl]-4-fluorobenzene (CH12); Typical Procedure Iodine (1.29 g, 5.10 mmol, 0.1 equiv) was added to a solution of 4-fluorobenzaldehyde (6.30 g, 50.8 mmol, 1.0 equiv) and benzenethiol (11.7 g, 107 mmol, 2.1 equiv) in CHCl3 (75 mL), and the resulting solution was stirred overnight at r.t. When the reaction was complete, excess I2 was quenched with 0.1 M aq Na2S2O3 (100 mL). The organic phase was separated, washed with H2O (100 mL), dried (Na2SO4), filtered, and concentrated on a rotary evaporator. Purification by flash column chromatography [silica gel, PE–CH2Cl2 (95:5 to 80:20)] and subsequent crystallization from petroleum ether gave CH12 as a white solid; yield: 12.8 g (77%, 39.2 mmol). Rf = 0.30 (silica gel, PE–CH2Cl2, 80:20, UV254). 1H NMR (500 MHz, 298 K, CD2Cl2): δ = 7.42–7.36 (m, 6 H), 7.34–7.28 (m, 6 H), 7.03–6.99 (m, 2 H), 5.53 (s, 1 H). 13C NMR (125 MHz, 298 K, CD2Cl2): δ = 162.7 (d, 1 J CF = 246.5 Hz), 135.9 (d, 4 J CF = 3.2 Hz), 134.5, 132.9, 131.0 (d, 3 J CF = 8.3 Hz), 129.3, 128.3, 114.6 (d, 2 J CF = 21.8 Hz), 59.6. Anal. calcd for C19H15FS2: C, 69.91; H, 4.63; S, 19.64. Found: C, 70.05; H, 4.84; S, 20.05.
  • 41 1-Fluoro-4-[tris(phenylthio)methyl]benzene (C13 ) and 1-methoxy-4-[tris(phenylthio)methyl]benzene (D13 ) were synthesized by a modified version of the reported procedure; see Ref 22a. 1-Fluoro-4-[tris(phenylthio)methyl]benzene (C13); Typical Procedure Dithioacetal DH12 (2.49 g, 7.66 mmol, 1.0 equiv) and TMEDA (2.49 g. 21.4 mmol, 2.8 equiv) were dissolved in anhyd THF (15 mL) under N2. The solution was cooled to –78 °C and a 2.5 M solution of BuLi in hexane (4.29 mL, 10.7 mmol, 1.4 equiv) was added dropwise. The mixture was stirred for 80 min at –78 °C, a solution of diphenyl disulfide (5.02 g, 23.0 mmol, 3.0 equiv) in anhyd THF (10 mL) was added slowly, and the mixture was allowed to warm to r.t overnight. The mixture was then cooled to 0 °C and the reaction was carefully quenched with several drops of H2O. The resulting mixture was extracted with Et2O (2 × 70 mL), washed with H2O (100 mL) and brine (100 mL), and the organic layer was separated, dried (Na2SO4), filtered, and concentrated on a rotary evaporator. Purification by flash column chromatography [silica gel, PE–CH2Cl2 (100:0 to 80:20)] and subsequent crystallization from PE–CH2Cl2 gave C13 as a white solid; yield: 2.30 g (69%, 5.30 mmol); Rf = 0.32 (silica gel, PE–CH2Cl2, 80:20, UV254). 1H NMR (500 MHz, 298 K, CD2Cl2): δ = 7.68–7.60 (m, 2 H), 7.34–7.18 (m, 15 H), 6.90–6.82 (m, 2 H). 13C NMR (125 MHz, 298 K, CD2Cl2): δ = 162.6 (d, 1 J CF = 248.0 Hz), 135.7 (d, 4 J CF = 3.1 Hz), 135.0, 133.1, 131.2 (d, 3 J CF = 8.3 Hz), 129.0, 128.7, 114.9 (d, 2 J CF = 21.6 Hz), 76.7. Anal. calcd for C25H19FS3: C, 69.09; H, 4.41; S, 22.13. Found: C, 68.96; H, 4.61; S, 22.11.
  • 42 In the case of O,O,O-orthoesters, hydrolysis is presumably the source of the free nucleophile, whereas with S,S,S-orthoesters, our observation of a bright-pink color might indicate that a thiol/thionium pair is formed, even without participation of water.
    • 43a Fan Y, An Z, Pan X, Liu X, Bao X. Chem. Commun. 2009; 7488
    • 43b Node M, Nishide K, Fuji K, Fujita E. J. Org. Chem. 1980; 45: 4275

  • References and Notes

    • 1a Lehn J.-M. Angew. Chem. Int. Ed. 2015; 54: 3276
    • 1b Jin Y, Wang Q, Taynton P, Zhang W. Acc. Chem. Res. 2014; 47: 1575
    • 1c Corbett PT, Leclaire J, Vial L, West KR, Wietor J.-L, Sanders JK. M, Otto S. Chem. Rev. 2006; 106: 3652
    • 1d Furlan RL. E, Otto S, Sanders JK. M. Proc. Natl. Acad. Sci. U. S. A. 2002; 99: 4801
    • 2a Lehn J.-M. Top. Curr. Chem. 2011; 322: 1
    • 2b Lehn J.-M. Chem. Soc. Rev. 2007; 36: 151
  • 3 Kölmel DK, Kool ET. Chem. Rev. 2017; 117: 10358
    • 4a Black SP, Sanders JK. M, Stefankiewicz AR. Chem. Soc. Rev. 2014; 43: 1861
    • 4b Fritze UF, von Delius M. Chem. Commun. 2016; 52: 6363
  • 5 Belowich ME, Stoddart JF. Chem. Soc. Rev. 2012; 41: 2003
    • 6a Beeren SR, Sanders JK. M. J. Am. Chem. Soc. 2011; 133: 3804
    • 6b Lam RT. S, Belenguer A, Roberts SL, Naumann C, Jarrosson T, Otto S, Sanders JK. M. Science 2005; 308: 667
    • 6c Furlan RL. E, Ng Y.-F, Otto S, Sanders JK. M. J. Am. Chem. Soc. 2001; 123: 8876
  • 7 Ulrich S, Boturyn D, Marra A, Renaudet O, Dumy P. Chem. Eur. J. 2014; 20: 34
    • 8a Mastalerz M. Acc. Chem. Res. 2018; 51: 2411
    • 8b Ong WJ, Swager TM. Nat. Chem. 2018; 10: 1023
    • 8c Guan X, Li H, Ma Y, Xue M, Fang Q, Yan Y, Valtchev V, Qiu S. Nat. Chem. 2019; 11: 587
    • 8d Ono K, Iwasawa N. Chem. Eur. J. 2018; 24: 17856
    • 8e Ding H, Chen R, Wang C. In Dynamic Covalent Chemistry: Principles, Reactions, and Applications . Zhang W, Jin Y. Wiley; New York: 2018: 165
    • 8f Feng X, Ding X, Jiang D. Chem. Soc. Rev. 2012; 41: 6010
    • 9a Frei P, Hevey R, Ernst B. Chem. Eur. J. 2019; 25: 60
    • 9b Liu Y, Lehn J.-M, Hirsch AK. H. Acc. Chem. Res. 2017; 50: 376
    • 9c Mondal M, Hirsch AK. H. Chem. Soc. Rev. 2015; 44: 2455
  • 10 Brachvogel R.-C, von Delius M. Chem. Sci. 2015; 6: 1399
  • 11 Orrillo AG, Escalante AM, Furlan RL. E. Chem. Eur. J. 2016; 22: 6746
  • 12 Brachvogel R.-C, von Delius M. Eur. J. Org. Chem. 2016; 3662
    • 13a Greene TW, Wuts PG. M. Protective Groups in Organic Synthesis, 2nd ed. Wiley; New York: 1991
    • 13b Kocienski PJ. Protecting Groups, 3rd ed. Thieme; Stuttgart: 1994
  • 14 Collins MS, Shear TA, Smith EK, Strain SM, Zakharov LN, Johnson DW. Chem. Eur. J. 2019; DOI: 10.1002/chem.201903854.
    • 15a Brachvogel R.-C, Hampel F, von Delius M. Nat. Commun. 2015; 6: 7129
    • 15b Brachvogel R.-C, Maid H, von Delius M. Int. J. Mol. Sci. 2015; 16: 20641
    • 15c Shyshov O, Brachvogel R.-C, Bachmann T, Srikantharajah R, Segets D, Hampel F, Puchta R, von Delius M. Angew. Chem. Int. Ed. 2017; 56: 776
    • 15d Wang X, Shyshov O, Hanževački M, Jäger CM, von Delius M. J. Am. Chem. Soc. 2019; 141: 8868
    • 15e Löw H, Mena-Osteritz E, von Delius M. Chem. Sci. 2018; 9: 4785
  • 16 Orrillo AG, Escalante AM, Furlan RL. E. Org. Lett. 2017; 19: 1446
  • 18 Lebel H, Grenon M. In Science of Synthesis . Vol. 22; Charette AB. Thieme; Stuttgart: 2005: 749
    • 19a Seebach D, Geiß K.-H, Beck AK, Graf B, Daum H. Chem. Ber. 1972; 105: 3280
    • 19b Barsamian AL, Blakemore PR. Organometallics 2012; 31: 19
  • 20 Cao Y.-J, Lai Y.-Y, Cao H, Xing X.-N, Wang X, Xiao W.-J. Can. J. Chem. 2006; 84: 1529
  • 21 Seebach D. Angew. Chem. 1967; 79: 468
    • 22a Tlach BC, Tomlinson AL, Ryno AG, Knoble DD, Drochner DL, Krager KJ, Jeffries-EL M. J. Org. Chem. 2013; 78: 6570
    • 22b Corey EJ, Seebach D. Angew. Chem. 1965; 77: 1134
    • 22c Gröbel B.-T, Seebach D. Synthesis 1977; 357

      For syntheses of dithioacetals see:
    • 23a Firouzabadi H, Iranpoor N, Hazarkhani H. J. Org. Chem. 2001; 66: 7527
    • 23b Tazaki M, Takagi M. Chem. Lett. 1979; 8: 767
    • 24a Moulin E, Cormos G, Giuseppone N. Chem. Soc. Rev. 2012; 41: 1031
    • 24b Li J, Nowak P, Otto S. J. Am. Chem. Soc. 2013; 135: 9222
    • 24c Jin Y, Yu C, Denman RJ, Zhang W. Chem. Soc. Rev. 2013; 42: 6634
    • 25a Eckert F, Leito I, Kaljurand I, Kütt A, Klamt A, Diedenhofen M. J. Comput. Chem. 2009; 30: 799
    • 25b Kütt A, Selberg S, Kaljurand I, Tshepelevitsh S, Heering A, Darnell A, Kaupmees K, Piirsalu M, Leito I. Tetrahedron Lett. 2018; 59: 3738
    • 25c Paenurk E, Kaupmees K, Himmel D, Kütt A, Kaljurand I, Koppel IA, Krossing I, Leito I. Chem. Sci. 2017; 8: 6964
  • 26 Pruckmayr G, Wu TK. Macromolecules 1978; 11: 662
    • 27a Otto S, Furlan RL. E, Sanders JK. M. J. Am. Chem. Soc. 2000; 122: 12063
    • 27b Delius M, Geertsema EM, Leigh DA, Slawin AM. Org. Biomol. Chem. 2010; 8: 4617
    • 28a Schmittel M, Levis M. Synlett 1996; 315
    • 28b Kamal A, Laxman E, Reddy PS. M. M. Synlett 2000; 1476
    • 28c Gauthier JY, Martins EO, Young RN, Zamboni RJ. Synlett 2002; 984
    • 28d Kumar V, Dev S. Tetrahedron Lett. 1983; 24: 1289
    • 28e Patney HK. Tetrahedron Lett. 1991; 32: 2259
    • 28f Tamami B, Borujeny KP. Synth. Commun. 2003; 33: 4253
    • 29a Rasmussen B, Sørensen A, Gotfredsen H, Pittelkow M. Chem. Commun. 2014; 50: 3716
    • 29b Ji S, Cao W, Yu Y, Xu H. Angew. Chem. Int. Ed. 2014; 53: 6781
    • 29c Chuard N, Poblador-Bahamonde AI, Zong L, Bartolami E, Hildebrandt J, Weigand W, Sakai N, Matile S. Chem. Sci. 2018; 9: 1860
    • 29d Bartolami E, Basagiannis D, Zong L, Martinent R, Okamoto Y, Laurent Q, Ward TR, Gonzalez-Gaitan M, Sakai N, Matile S. Chem. Eur. J. 2019; 25: 4047
    • 29e Steinmann D, Nauser T, Koppenol WH. J. Org. Chem. 2010; 75: 6696
    • 31a Cordes EH, Bull HG. Chem. Rev. 1974; 74: 581
    • 31b Chiang Y, Kresge AJ, Lahti MO, Weeks DP. J. Am. Chem. Soc. 1983; 105: 6852
    • 31c Repetto SL, Costello JF, Butts CP, Lam JK. W, Ratcliffe NM. Beilstein J. Org. Chem. 2016; 12: 1467
    • 31d Deslongchamps P, Dory YL, Li S. Tetrahedron 2000; 56: 3533
    • 32a Bhawal BN, Morandi B. ACS Catal. 2016; 6: 7528
    • 32b Bhawal BN, Morandi B. Chem. Eur. J. 2017; 23: 12004
  • 33 Bhawal BN, Morandi B. Angew. Chem. Int. Ed. 2019; 58: 10074
    • 34a Otsuka H, Nagano S, Kobashi Y, Maeda T, Takahara A. Chem. Commun. 2010; 46: 1150
    • 34b Caraballo R, Sakulsombat M, Ramström O. Chem. Commun. 2010; 46: 8469
  • 35 Amamoto Y, Kamada J, Otsuka H, Takahara A, Matyjaszewski K. Angew. Chem. Int. Ed. 2011; 50: 1660
  • 36 Saiz C, Wipf P, Manta E, Mahler G. Org. Lett. 2009; 11: 3170
  • 37 Cacciapaglia R, Di Stefano S, Mandolini L. J. Am. Chem. Soc. 2005; 127: 13666
  • 38 Tris[(4-methylbenzene)thio]methane (A53 ), tris(isopropylthio)methane (A43 ), and tris[(2-phenylethyl)thio]methane [A(10)3 ] were synthesized by a modified version of the reported procedure; see Ref. 20. Tris[(4-methylbenzene)thio]methane (A53 ); Typical Procedure CHCl3 (1.48 g, 12.4 mmol, 1.0 equiv), DBU (9.79 g, 64.3 mmol, 5.2 equiv), and 4-methylbenzenethiol (6.14 g, 49.4 mmol, 4.0 equiv) were dissolved in anhyd THF (10 mL), and the mixture was heated to 100 °C for 24 h under N2 in a pressure tube. The mixture was cooled, H2O (30 mL) was added, and the resulting mixture was extracted with Et2O (2 × 50 mL). The combined organic phases were washed with brine (50 mL), dried (Na2SO4), filtered, and concentrated on a rotary evaporator. The residue was purified by column chromatography [silica gel, PE–CH2Cl2 (100:0 to 80:20)]. Subsequent crystallization (CH2Cl2) gave A53 as a white solid; yield: 2.10 g (44%, 5.48 mmol); Rf = 0.3 (silica gel, PE–CH2Cl2, 80:20, UV254). 1H NMR (400 MHz, 298 K, CDCl3): δ = 7.43–7.41 (m, 6 H), 7.17–7.15 (m, 6 H). 5.32 (s, 1 H), 2.37 (s, 9 H). 13C NMR (100 MHz, 298 K, CDCl3): δ = 138.5, 133.4, 130.4, 129.6, 65.9, 21.2. Anal. calcd for C22H22S3: C, 69.07; H, 5.80; S, 25.14. Found: C, 69.00; H, 5.93; S, 25.22.
  • 39 Synthesis of 1,1,1-Tris(methylthio)ethane (B83) This was synthesized by a modified version of the reported procedure.21 Trimethyl trithioorthoformate (A83 ; 5.28 g, 34.2 mmol, 1.0 equiv) was dissolved in anhyd THF (50 mL) and cooled to –78 °C under N2. A 2.5 M solution of BuLi in hexane (20.5 mL, 51.3 mmol, 1.5 equiv) was added dropwise and the mixture was stirred for 80 min. MeI (12.13 g, 85.5 mmol, 2.5 equiv) was added dropwise, and the mixture was allowed to slowly warm to r.t. overnight. Et2O (100 mL) was added and the mixture washed with aq Na2S2O3 (50 mL). The combined extracted organic phases were washed with brine (50 mL), separated, dried (Na2SO4), filtered, and concentrated on a rotary evaporator. Distillation through a short Vigreux column (oil-bath temperature: 120 °C; pressure: 7 mbar; bp 70 °C) gave a colorless oil; yield: 4.55 g (79%, 27.0 mmol); R f = 0.7 (silica gel, PE–EtOAc, 95:5, UV254). 1H NMR (400 MHz, 298 K, CDCl3): δ = 2.16 (s, 9 H), 1.87 (s, 3 H). 13C NMR (100 MHz, 298 K, CDCl3): δ = 64.3, 28.3, 13.6. Anal. calcd for C5H12S3: C, 35.68; H, 7.19; S, 57.14. Found: C, 37.04; H, 7.26; S, 57.81.
  • 40 1-[Bis(phenylthio)methyl]-4-fluorobenzene (CH12 ) and 1-[bis(phenylthio)methyl]-4-methoxybenzene (DH12 ) were synthesized by a modified version of the reported procedure; see Ref 23a. 1-[Bis(phenylthio)methyl]-4-fluorobenzene (CH12); Typical Procedure Iodine (1.29 g, 5.10 mmol, 0.1 equiv) was added to a solution of 4-fluorobenzaldehyde (6.30 g, 50.8 mmol, 1.0 equiv) and benzenethiol (11.7 g, 107 mmol, 2.1 equiv) in CHCl3 (75 mL), and the resulting solution was stirred overnight at r.t. When the reaction was complete, excess I2 was quenched with 0.1 M aq Na2S2O3 (100 mL). The organic phase was separated, washed with H2O (100 mL), dried (Na2SO4), filtered, and concentrated on a rotary evaporator. Purification by flash column chromatography [silica gel, PE–CH2Cl2 (95:5 to 80:20)] and subsequent crystallization from petroleum ether gave CH12 as a white solid; yield: 12.8 g (77%, 39.2 mmol). Rf = 0.30 (silica gel, PE–CH2Cl2, 80:20, UV254). 1H NMR (500 MHz, 298 K, CD2Cl2): δ = 7.42–7.36 (m, 6 H), 7.34–7.28 (m, 6 H), 7.03–6.99 (m, 2 H), 5.53 (s, 1 H). 13C NMR (125 MHz, 298 K, CD2Cl2): δ = 162.7 (d, 1 J CF = 246.5 Hz), 135.9 (d, 4 J CF = 3.2 Hz), 134.5, 132.9, 131.0 (d, 3 J CF = 8.3 Hz), 129.3, 128.3, 114.6 (d, 2 J CF = 21.8 Hz), 59.6. Anal. calcd for C19H15FS2: C, 69.91; H, 4.63; S, 19.64. Found: C, 70.05; H, 4.84; S, 20.05.
  • 41 1-Fluoro-4-[tris(phenylthio)methyl]benzene (C13 ) and 1-methoxy-4-[tris(phenylthio)methyl]benzene (D13 ) were synthesized by a modified version of the reported procedure; see Ref 22a. 1-Fluoro-4-[tris(phenylthio)methyl]benzene (C13); Typical Procedure Dithioacetal DH12 (2.49 g, 7.66 mmol, 1.0 equiv) and TMEDA (2.49 g. 21.4 mmol, 2.8 equiv) were dissolved in anhyd THF (15 mL) under N2. The solution was cooled to –78 °C and a 2.5 M solution of BuLi in hexane (4.29 mL, 10.7 mmol, 1.4 equiv) was added dropwise. The mixture was stirred for 80 min at –78 °C, a solution of diphenyl disulfide (5.02 g, 23.0 mmol, 3.0 equiv) in anhyd THF (10 mL) was added slowly, and the mixture was allowed to warm to r.t overnight. The mixture was then cooled to 0 °C and the reaction was carefully quenched with several drops of H2O. The resulting mixture was extracted with Et2O (2 × 70 mL), washed with H2O (100 mL) and brine (100 mL), and the organic layer was separated, dried (Na2SO4), filtered, and concentrated on a rotary evaporator. Purification by flash column chromatography [silica gel, PE–CH2Cl2 (100:0 to 80:20)] and subsequent crystallization from PE–CH2Cl2 gave C13 as a white solid; yield: 2.30 g (69%, 5.30 mmol); Rf = 0.32 (silica gel, PE–CH2Cl2, 80:20, UV254). 1H NMR (500 MHz, 298 K, CD2Cl2): δ = 7.68–7.60 (m, 2 H), 7.34–7.18 (m, 15 H), 6.90–6.82 (m, 2 H). 13C NMR (125 MHz, 298 K, CD2Cl2): δ = 162.6 (d, 1 J CF = 248.0 Hz), 135.7 (d, 4 J CF = 3.1 Hz), 135.0, 133.1, 131.2 (d, 3 J CF = 8.3 Hz), 129.0, 128.7, 114.9 (d, 2 J CF = 21.6 Hz), 76.7. Anal. calcd for C25H19FS3: C, 69.09; H, 4.41; S, 22.13. Found: C, 68.96; H, 4.61; S, 22.11.
  • 42 In the case of O,O,O-orthoesters, hydrolysis is presumably the source of the free nucleophile, whereas with S,S,S-orthoesters, our observation of a bright-pink color might indicate that a thiol/thionium pair is formed, even without participation of water.
    • 43a Fan Y, An Z, Pan X, Liu X, Bao X. Chem. Commun. 2009; 7488
    • 43b Node M, Nishide K, Fuji K, Fujita E. J. Org. Chem. 1980; 45: 4275

Zoom Image
Scheme 1 Exchange reactions of (A) orthoesters, (B) dithioacetals, and (C) trithioorthoesters
Zoom Image
Scheme 2 Effect of the structure of the nucleophile on the distribution of trithioorthoesters at equilibrium. Reaction conditions: A13 or A83 (37.5 μmol), nucleophile (112.5 μmol), TFA (375 μmol), CDCl3 (V total: 750 μL), r.t. Product ratios were determined by integration of 1H NMR signals after 1 h. Hydrolysis was negligible (<2%) in all cases except for thiol 6 (7%). a Exchange performed with A13 (CDCl3 or CD3CN). b Exchange performance with A83 (CDCl3). For full spectra, see Figures S26–S36 (SI).
Zoom Image
Figure 1 Effects of increasing amounts of TFA on the hydrolysis of ­orthoester (O,O,O)-A93 , trithioorthoester (S,S,S)-A83 , and dithioacetal AH82 . Reaction conditions: A93 or A83 or AH82 (37.5 μmol), increasing amounts of TFA [from 37.5 nmol (50 μM) to 3.75 mmol (5.0 M)], CDCl3 (V total: 750 μL), r.t., 1 h; internal standard: toluene. For full spectra, see Figures S39–S56 in the SI.
Zoom Image
Figure 2 Trithioorthoester metathesis between A13 and B83 (top) and 1H NMR spectrum measured after one hour (bottom). The two dashed boxes show the corresponding formate and acetoxy hydrogen atoms of trithioorthoesters formed from A13 and B83 . Reaction conditions: A13 (37.5 μmol), B83 (37.5 μmol), TFA (375 μmol), CDCl3 (V total: 750 μL), r.t., 1 h. Product distributions are given by normalized integral values below the corresponding peaks for both sets. See also Table [2], entry 1 and the full spectra in Figure S59 in the SI.