ACS Publications. Most Trusted. Most Cited. Most Read
NHC-CDI Betaine Adducts and Their Cationic Derivatives as Catalyst Precursors for Dichloromethane Valorization
My Activity

Figure 1Loading Img
  • Open Access
Article

NHC-CDI Betaine Adducts and Their Cationic Derivatives as Catalyst Precursors for Dichloromethane Valorization
Click to copy article linkArticle link copied!

  • David Sánchez-Roa
    David Sánchez-Roa
    Departamento de Química Orgánica y Química Inorgánica, Instituto de Investigación en Química “Andrés M. del Río” (IQAR) Universidad de Alcalá, Campus Universitario, Alcala de Henares, Madrid 28871, Spain
  • Marta E. G. Mosquera*
    Marta E. G. Mosquera
    Departamento de Química Orgánica y Química Inorgánica, Instituto de Investigación en Química “Andrés M. del Río” (IQAR) Universidad de Alcalá, Campus Universitario, Alcala de Henares, Madrid 28871, Spain
    *Email: [email protected]
  • Juan Cámpora*
    Juan Cámpora
    Instituto de Investigaciones Químicas, CSIC-Universidad de Sevilla, C/Américo Vespucio, 49, Sevilla 41092, Spain
    *Email: [email protected]
Open PDFSupporting Information (1)

The Journal of Organic Chemistry

Cite this: J. Org. Chem. 2021, 86, 23, 16725–16735
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.joc.1c01971
Published November 1, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Zwitterionic adducts of N-heterocyclic carbene and carbodiimide (NHC-CDI) are an emerging class of organic compounds with promising properties for applications in various fields. Herein, we report the use of the ICyCDI(p-Tol) betaine adduct (1a) and its cationic derivatives 2a and 3a as catalyst precursors for the dichloromethane valorization via transformation into high added value products CH2Z2 (Z = OR, SR or NR2). This process implies selective chloride substitution of dichloromethane by a range of nucleophiles Na+Z (preformed or generated in situ from HZ and an inorganic base) to yield formaldehyde-derived acetals, dithioacetals, or aminals with full selectivity. The reactions are conducted in a multigram-scale under very mild conditions, using dichloromethane both as a reagent and solvent, and very low catalyst loading (0.01 mol %). The CH2Z2 derivatives were isolated in quantitative yields after filtration and evaporation, which facilitates recycling the dichloromethane excess. Mechanistic studies for the synthesis of methylal CH2(OMe)2 rule out organocatalysis as being responsible for the CH2 transfer, and a phase-transfer catalysis mechanism is proposed instead. Furthermore, we observed that 1a and 2a react with NaOMe to form unusual isoureate ethers, which are the actual phase-transfer catalysts, with a strong preference for sodium over other alkali metal nucleophiles.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2021 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Dichloromethane (DCM) is a relatively inert compound that is widely applied as a solvent in organic synthesis and separation procedures. It can be regarded as a low polarity, nonprotic, and noncoordinating solvent, which, however, dissolves many polar or even ionic species. (1) The heavier dihalomethane congeners, dibromomethane and diiodomethane, are considerably more reactive than DCM. Those are rarely used as a solvent but often employed as a methylene source via nucleophilic substitution reactions. (2) Indeed, along with formaldehyde, dihalomethanes represent an important C1 chemical building block. (3) DCM could be a convenient, readily available, and inexpensive alternative to these compounds; nevertheless, it is seldom used as a chemical reagent. The chemical stability of C–Cl bonds, which makes DCM such an excellent solvent, usually leads to difficult chloride substitution processes. (4) The latter are often plagued by side reactions, such as the HCl elimination whenever the nucleophiles behave as strong bases or free radical chain processes. (4a,5) These unwanted reactions can entail explosion hazards in some cases. (6) Yet, the use of DCM as the source of the CH2 fragment could help improve many chemical transformations beyond the mere cost reduction. Among other advantages, DCM allows much safer handling than formaldehyde, a widely used CH2 precursor in classic organic synthesis. (7) Although the intensive usage of organohalogens can lead to environmental issues, there are strategies that can be followed in order to optimize their application and minimize this impact. (8) For instance, DCM can be used simultaneously as solvent and reagent, and due to its low boiling point and its poor miscibility with water and aqueous mixtures, it can be readily separated from polar or heavier substances and recycled for further uses. (9) Among others, clean chlorine substitution in DCM could lead to significant improvements in the syntheses of a wide variety of methylene-bridged derivatives, such as formaldehyde-derived acetals, dithioacetals, or aminals. Formaldehyde acetals or formals are chemicals with many practical applications. Those have proven useful as fuels or fuel additives, (10) and also as reagents for the introduction of an alkoxy- or aryloxymethylene group. (11) On the other hand, dithioacetals and aminals are less used in the industry than formals but have important applications in synthesis (12) and coordination chemistry. (13) Although transition metal catalysts (14) (e.g., Ni (14a) or Cu (14b) complexes) have been occasionally used for coupling dichloromethane with nucleophiles, the usefulness of such processes is usually limited by the need for high metal loadings and relatively low yields.
In this contribution, we disclose a practical application of stable adducts of NHC carbenes and carbodiimides (NHC-CDI) as suitable catalysts for nucleophilic chloride substitution on DCM, providing a convenient route for a variety of symmetrical, methylene-bridged derivatives CH2Z2. NHC-CDI’s are an emerging class of dipolar compounds that belong to the chemical class of betaines, electroneutral zwitterions that cannot be represented by any resonance form with full charge cancellation. (15) NHC-CDI’s bearing aryl substituents in the CDI part are readily prepared and exhibit enhanced stability even under the open air, despite the strongly basic and nucleophilic character of their amidinate moiety. (16) Due to these properties, stable NHC-CDI adducts have a promising potential for application in many different fields, including metal ligands in coordination and organometallic chemistry, (17) nanocatalyst design, (18) building blocks for polymers, (19) or in the development of new types of persistent free radicals. (20)
Recently, we have reported that the ICyCDI(p-Tol) betaine adduct (1a), an NHC-CDI derivative containing 1,3-dicyclohexylimidazolylidene (ICy) and di-p-tolylcarbodiimide (CDI(p-Tol)) as the NHC and CDI parts, respectively, reacts in DCM solution to afford the ionic salt [CH2(1a)2]2+[2Cl] (2a·2Cl). Despite the considerable steric bulk of 1a, this process, which implies the cleavage of both C–Cl bonds of DCM, proceeds in quantitative yield and full selectivity, without any detectable byproducts or intermediates (Scheme 1). (21) Compound 2a can be envisaged as a “[CH2]2+” fragment trapped by two neutral 1a units. We were intrigued by the possibility that the central core of the 2a dication could be activated for nucleophilic substitution, delivering this group to mild nucleophiles. In a catalytic version of this process, DCM could be applied as a source of the [CH2]2+ synthon, and 1a (or some other related NHC-CDI derivatives) might act as an organocatalyst in an unusual class of organocatalyzed nucleophilic substitution reaction. Herein we show that this is indeed the case, although further investigation of the details shows that the mechanism of this catalytic transformation differs largely from our initial intuition.

Scheme 1

Scheme 1. Synthesis of 1a and 2a

Results and Discussion

Click to copy section linkSection link copied!

To test our proposal, we first used solid sodium hydroxide as a nucleophile. This reaction led to the substitution of Cl for OH, affording hydrated formaldehyde oligomers, which could be detected in the 1H NMR spectra of the mixture (22) (broad signal at 4.88 ppm, see Figure S25) and also because of the characteristic smell of free formaldehyde, but the efficiency of the reaction was hard to quantify. Therefore, we shifted our trial to solid sodium methoxide, which would enable the synthesis of methylal through the substitution of chlorine atoms of DCM by methoxy groups (Scheme 2). Methylal or dimethoxymethane is an interesting target since it can be used as a green solvent (23) (far less toxic than DCM itself) or as a safer surrogate of formaldehyde, preventing the use of this toxic substance or its insoluble polymer, paraformaldehyde. (24) Yet, it has been known for a very old time that neat dichloromethane does not react with sodium methoxide, or it does only reluctantly under forcing conditions to afford complex mixtures of products. (25) We performed a series of tests with different amounts of solid sodium methoxide suspended in neat DCM at 25 and 60 °C, using a sealed ampule with a magnetic stirrer as a reactor. We confirmed that, whereas hardly any detectable quantity of methylal was formed in the absence of 2a, the addition of a catalytic amount of the latter (from 2 to 0.2 mol %) led to the quantitative conversion of NaOMe into methylal and NaCl. Due to the similar boiling points of DCM and methylal, we made no attempt to separate the product from the excess solvent. However, spectroscopic analyses of the crude mixtures showed that the transformation is very clean. (26)

Scheme 2

Scheme 2. Synthesis of Methylal Using 2a as a Catalyst
The NMR spectrum of a drop of the liquid phase in CDCl3 displayed only major signals for methylal (3.30 and 4.52 ppm) and unreacted DCM (5.32 ppm), along with very minor signals from the catalyst. We then decided to scale-up the process to show the potential of this reaction. Hence, the synthesis of methylal was carried out starting from 2 g of sodium methoxide and 20 mL of DCM, at 60 °C, using 0.2 mol % of catalyst 2a, and full conversion was achieved. Finally, the catalyst loading was decreased to 0.01 mol % (100 ppm) and NaOMe quantity scaled up to 20 g, leading to an apparent TON figure in the order of 104 without any loss of purity or selectivity. The use of such a small proportion of a metal-free catalyst is remarkable. (27)
NMR analyses of samples removed at regular intervals from a stirred suspension of solid NaOMe in DCM containing 2a (0.2 mol % vs NaOMe) at 40 °C showed that the methylal content of samples increased linearly up to 70% conversion (68% within 7 h) and was consistent with full conversion when a final aliquot was measured after 24 h (Figure 1). These data indicate that the activity of 2a does not show any significant decay over long periods of time, at a turnover frequency (TOF) of 45 h–1 (referred to DCM, or 90 h–1 for NaOMe) under such mild conditions. Moreover, the addition of a second 2 g portion of sodium methoxide to a completed reaction mixture containing an excess of unreacted DCM led again to full conversion within the expected 24 h period, suggesting that this catalyst remains indefinitely active under the specified conditions.

Figure 1

Figure 1. Yield vs time plot for the reaction of DCM with NaOMe catalyzed by 2a. Conditions: 40 °C, NaOMe, 2 g (31 mmol); DCM (neat) 40 mL (626 mmol); 2a, 37 mg (0.2 mol % with regard to NaOMe). The last check after 24 h was consistent with full NaOMe consumption.

With this first accomplishment in hand, we set out to explore other suitable nucleophiles to define the scope of this interesting reaction. In particular, we analyzed whether other O, S, or N-based nucleophiles would perform in a similar way in order to obtain formaldehyde-derived acetals, dithioacetals, or aminals, respectively. The results are displayed in Table 1. We first explored the reaction of DCM with solid sodium phenoxide and sodium benzyloxide under the same conditions developed for the synthesis of methylal (entries 2 and 3). We selected these alkoxides because they should exhibit milder reactivity than sodium methoxide. Yet, both of them reacted with DCM under mild conditions, affording the corresponding products in a highly selective manner. Essentially pure products were isolated after a very simple workup involving filtration through a silica pad (to remove NaCl and the remaining catalyst) and evaporation under a vacuum. The purity and quantitative isolated yield confirmed the high selectivity appreciated in the NMR analyses of the crude mixtures.
Table 1. Catalyzed Nucleophilic Substitution Reactions in Dichloromethane with Various Nucleophiles and Conditionsa
entryproduct CH2Z2bsubstrate/baseccat.conversiond/yielde
1iNaOMe2a>99/>99
2iiNaOPh2a>99/92
3iiiNaOBn2a>99/87
4iMeOH/NaOH2a>99/89
5iiPhOH/NaOH2a>99/89
6iiiBnOH/NaOH2a>99/86
7ivPTBPf/NaOH2a>99/85
8vAAg/NaOH2a>99/83
9vii-PrOH/NaH2a74/67
10viiEtSH/NaOH2a>99
11viiiHPzh/NaH2a89/72
12iNaOMe1a>99/>99
13iiNaOPh1a>99/86
14iiiNaOBn1a>99/87
15iNaOMe3a>99/>99
a

Reaction conditions: catalyst loading 0.2 mol %, 60 °C, 24 h.

b

See Scheme 3 for the detailed structure of the products.

c

Insoluble bases (NaOH, NaH), added in excess.

d

Spectroscopic yield from 1H NMR.

e

Isolated yields (calculated on the basis of the starting substrate), unless otherwise specified.

f

p-tert-Butylphenol.

g

Allyl alcohol.

h

Pyrazole.

In view of the excellent results of transformations with solid alkoxides, we examined a more practical setup for this catalytic reaction (Scheme 3). Thus, methylal was obtained using methanol combined with solid sodium hydroxide, a much cheaper and safer base than NaOMe. Although the spectroscopic yield of this reaction was slightly lower, the selectivity was excellent again since methylal was the only organic product detected, along with unreacted DCM. Despite the fact that 2a also catalyzes the reaction of DCM with NaOH, no traces of formaldehyde or formaldehyde oligomers were detected in the NMR spectra of the reaction mixture. The apparent losses in the mass balance can be attributed to the poor mechanical properties of the sticky solid formed by wet sodium salts, which could occlude some methanol.

Scheme 3

Scheme 3. Catalytic Transformation of DCM into CH2Z2 Using Combinations of Weak Protic Acids and Suitable Bases (NaOH or NaH)
In a similar fashion, 2a efficiently catalyzed the reaction of several alcohols with DCM at 0.2 mol % catalyst loading in the presence of suitable bases. Among the substrates, those that gave the best conversions with NaOH were primary alcohols and phenol derivatives. Thus, benzyl alcohol, allyl alcohol (AA), phenol, and p-tert-butylphenol (PTBP), combined with NaOH, were cleanly transformed into the corresponding acetals with perfect selectivity and excellent yields. NaOH was not effective for 2-propanol, a less acidic secondary alcohol. Nevertheless, when the stronger base NaH was used, the corresponding acetal was produced. Surprisingly enough, the acidic p-chlorophenol turned out to be unreactive, either in the presence of NaOH or as the solid sodium salt, which we attribute to the low nucleophilicity of the corresponding anion. Regarding the synthesis of dithioacetals, we used ethanethiol as a reference for thiols, which was successfully converted to di(ethylthio)methane using NaOH as a base. For the synthesis of aminals, we performed the synthesis of bis(pyrazolyl)methane. This target was selected in view of the wide application of pyrazolyl-based molecules as chelating ligands for metal complexes. (13) Also, in this case, the functionalization of both C–Cl bonds of DCM and the subsequent formation of new C–N was achieved, but NaH was required for the deprotonation of pyrazole (HPz).

Mechanistic Studies

According to our initial ideas, the role of dication 2a would be to act as a source of the [CH2]2+ synthon, regenerating the precursor betaine 1a in each turn of the catalytic cycle. Accordingly, not only 2a but also betaine 1a would be active catalysts for this reaction. Similar reasoning led us to conclude that cationic derivatives, [ZCH2·1a]+, would participate as short-lived intermediates in the catalytic cycle, preceding the introduction of the second Z fragment. To test these hypotheses, we synthesized the salt [MeOCH2·1a]+[Br] (3a) as shown in Scheme 4 and compared its catalytic performance with those of 1a and 2a in the synthesis of methylal. Gratifyingly, as can be seen in Table 1, all three compounds perform as catalysts in the synthesis of methylal, with similar efficiency. Moreover, 1a was tested in the synthesis of diphenoxymethane and dibenzyloxymethane with identical results as 2a (Table 1, entries 12–15). However, we realized that the TON and TOF figures of our system were atypical for a regular organocatalytic mechanism involving systematic C–N bond formation and cleavage. On the basis of these data and previous qualitative observations, (21) it can be easily shown that this reaction is at least 2–3 orders of magnitude faster than the reaction of 1a with DCM at the same temperature, (28) which rules out our preliminary proposals.

Scheme 4

Scheme 4. Synthesis of the Catalyst Precursor 3a
An alternative phenomenon that would account for the high activity levels observed in our system is solid–liquid phase-transfer catalysis (PTC). PTC allows the solubilization of highly polar or ionic species in nonpolar solvents. Insoluble reagents that normally would be forced to react through the solid–liquid interphase are transported into the solution, where they react under rather diluted conditions. Therefore, solid–liquid PTC enhances the reactivity of solids, and at the same time improves selectivity, by reducing side processes that are common when highly reactive species are dissolved in a higher concentration (Figure 2). (29) Solid–liquid PTC is often regarded as a green methodology, as reagents and saline products are kept in the solid state and can be readily separated by physical methods. (30) Alkylation of alcohols and other nucleophilic substrates under PTC conditions is not a novel procedure, but the potential of this methodology using DCM both as a solvent and as alkylating reagent remains almost unexplored. (31) Although DCM has been found an excellent solvent for liquid–solid PTC, (32) in the rare cases when DCM is used as an electrophile, liquid–liquid PTC methods in biphasic media (with concentrated aqueous NaOH), or highly polar solvents like N-methyl-2-pyrrolidone have been preferred. (33) Interestingly, the above-mentioned transition metal catalysts also used DCM both as electrophile and solvent. (14)

Figure 2

Figure 2. Solid–liquid phase-transfer catalysis.

Several considerations point to PTC as the mechanism that best explains the high activity of our catalysts. First of all, a process controlled by the phase transfer step is consistent with the linear kinetic plot shown in Figure 1, which indicates a zero-order dependency on the reagents well over 50% conversion of NaOMe. Second, to put our results in the correct perspective, we decided to compare the performance of 1a, 2a, or 3a with those of a set of typical phase-transfer agents, like onium-type salts and crown ethers for the synthesis of methylal under the same experimental conditions (Table 2). To enable a more accurate comparison, we ran the experiments under the conditions required by 2a (0.2 mol %) to drive the reaction of DCM with NaOMe to ca. 50% conversion. The only difference with the preparative conditions was that, in order to facilitate data collection, this set of experiments was performed at a slightly lower temperature, 40 °C, at which the half-conversion time would be ca. 4 h. As shown in Table 2, no methylal was formed in the absence of any PTC reagent, confirming the inertness of DCM against solid NaOMe reported in old literature sources cited above. (25) Nevertheless, when we added the same catalyst loading of NBu4Cl, 18-crown-6, 15-crown-5, or the imidazolium salt ICy·HBF4 (the precursor of the NHC carbene ICy, one of the constitutive parts of 1a), DCM conversion was observed, albeit with significant differences. Remarkably, 2a proved the most active of the set. NBu4Cl, one of the best-known and widely used PTC reagents, is also active, with slightly lower activity, and the imidazolium salt showed a rather poor conversion. Moreover, we tested whether CDI(p-Tol), the carbodiimide moiety of 1a, was also active, but no conversion was achieved. At any rate, the efficiency of both 1a and 2a are comparable even to NBu4Cl and are clearly superior to a typical crown ether, such as 18-crown-6.
Table 2. Catalytic studies of 1a, 2a, 3a, and Other PTC Catalysts toward MOMe (M = Li, Na, K)
entrysubstratecatalystaconversionb (%)
1NaOMe  
2NaOMeNBu4Cl44
3NaOMe18-crown-620
4NaOMe15-crown-539
5NaOMeICy·HBF48
6NaOMeCDI(p-Tol) 
7NaOMe1a46
8NaOMe3a43
9NaOMe2a55
10LiOMe2a 
11KOMe2a5
a

Catalyst or catalyst precursor (see below).

b

An aliquot was taken after 4 h to calculate conversion by 1H NMR.

The mechanism responsible for the solubilization of insoluble nucleophiles depends largely on the nature of the PTC catalyst. Whereas tetraalkylammonium salts, like NBu4Cl, rely on the high solubility of its cation in low polarity solvents, the action of 18-crown-6 is due to the sequestration of the alkali metal cation in the cavity of the molecule. (29,30) Hence, the lower efficacy of 18-crown-6 in the reaction with NaOMe could be due to its moderate affinity for Na+ (e.g., the affinity ratio K+/Na+ for 18-crown-6 in the gas phase is 10:6). (34) It is well-known that the efficacy of complexing PTCs shows significant dependency on the size of the alkali metal. (35) As shown in entry 4, 15-crown-5, whose smaller cavity is best suited for Na+, performs better than 18-crown-6, but it is still inferior to 2a. Similarly, compound 2a shows a strong preference for Na+, performing its action in a much more efficient manner with NaOMe than with KOMe, whereas LiOMe turned out to be ineffective (Table 2, entries 9–11). These observations suggest that, in addition to the cation exchange solubilization mechanism, our catalysts may simultaneously be performing as selective ligands for the alkali metal. This conclusion is a somewhat unexpected result since the positive charge of the betaine derivatives (double in the case of 2a) should reduce their capacity to coordinate to cationic alkali metal centers and, in consequence, would be expected to behave in a nonspecific manner, like quaternary ammonium salts.
A significant detail observed during the reaction of NaOMe and DCM catalyzed by either 1a or the cations 2a and 3a is the evident intensification of the color within the initial hours of the reactions, leaving a strong reddish-orange tone that persists once the reaction is finished. Strongly colored reaction mixtures were also observed with other nucleophiles, independently on whether 1a or 2a were used as catalysts, but, revealingly, no color change was observed when the catalysis was unsuccessful, as in the reactions with p-chlorophenoxide or with lithium methoxide. In an attempt to identify the colored materials formed with NaOMe, we recorded the 1H and 13C{1H} spectra of the deeply colored oils remaining after evaporation of the reaction mixtures catalyzed with either 1a or 2a, by dissolving in CD2Cl2. The NMR spectra of the residue left by 2a was particularly clean, showing that the original signals of the methylene-bridged dication had fully disappeared and replaced with a new spectrum of what looked like a mixture containing a strongly prevalent species 2b (see Figure S20). The DOSY spectrum of the mixture (see Figure S19) allows a clear distinction of the signals of the main product, 2b, from those of background species, which have all significantly higher diffusion coefficient or, what is the same, lower molecular weight. The spectra of 2b correspond to a symmetrical species akin to 2a, which has no imidazolium moieties. The broad and ill-defined signals of the cyclohexyl fragments are all associated with the low molecular weight region, consistent with the degradation of the ICy unit under the reaction conditions. Concerning 2b, a single resonance of relative intensity for 6H is found at 3.53 ppm, corresponding to two equivalent methoxy groups, along with another singlet of 2H for the central CH2 unit at 5.15 ppm. The relative simplicity of the spectra of 2b enabled us to fully assign by 1H and 13C NMR the unusual isoureate ether structure shown in Scheme 5. The NMR spectra of the residue obtained from the 1a-catalyzed reaction (see Figure S17) indicates the presence of the prevalent species, 1b, different from the starting betaine and from 2b, which contains two resonances MeO (3.30 and 3.75 ppm) and a single CH2 unit at 4.69 ppm. Interestingly, the DOSY spectrum of this mixture (see Figure S16) gave slightly lower size to the molecules of 1b than to the broad background arising from the imidazolium fragment. These data clearly point to the unprecedented processes shown in Scheme 5, giving rise to 1b and 2b, rare examples of isoureate ethers, which are likely responsible for most of the PTC activity in the catalytic reaction. Accordingly, ESI-MS of partially purified extracts in wet MeOH showed intense signals for the corresponding monoprotonated ions (m/z = 299 and 521, respectively), as well as their corresponding Na+ complexes (321 and 543). Ion compositions were confirmed in the corresponding HR-ESI spectra (see Experimental part). Note that 1b is the result of trapping the reactive chloromethyl intermediate [ClCH2·1a]+[Cl] by a methoxide, giving the same characteristic methoxymethylene fragment, also present in 3a. This was confirmed by comparing the NMR spectra of 3a with the crude residue left after evaporation of DCM/methylal solutions (see Figure S24). Thus, 1b cannot be converted in 2b during the reaction. In consequence (and in contrast with our initial expectations), each of the catalytic reactions initiated by precursors 1a and 2a are due to a different catalytic species. According to this proposal, the catalyst generated from 3a is, very likely the same formed from 1a, namely, 1b.

Scheme 5

Scheme 5. Transformation of 2a and 1a into 2b and 1b, Respectively, in Catalytic Media
Attempts to separate 1b and 2b from the mixture of products resulting from the degradation of the NHC units have been unsuccessful so far, apparently due to the sensitivity of the isoureate ethers, which seem to hydrolyze on attempted purification by typical chromatographic workup. However, since the imidazolium precursor ICy·HBF4 shows very poor activity as a PTC catalyst, the effectivity of both 1a and 2a should be entirely ascribed to the electroneutral isoureate ethers, that, as commented above, must have a significant affinity for the Na+ cation. This is in accordance with the facile ionization of the neutral molecules with traces of Na+ in the methanol solvent. The low catalytic activity found for the imidazolium salt is consistent with these observations, as this is expected to be deprotonated to the same ICy moiety subsequently degrades under the reaction conditions. Likewise, the experiment performed with CDI(p-Tol) as a catalyst (Table 2, entry 6) demonstrates not only that the free carbodiimide is not be involved in the process, but also that the active isoureate 2b cannot be generated directly by the successive reaction of CDI(p-Tol) with NaOMe and DCM. To confirm this point, we attempted the reaction of CDI(p-Tol) with NaOMe in DCM but, as expected, no reaction was observed. Thus, the only plausible pathway to generate the catalytically active species 1b and 2b is through betaine 1a and its cationic derivatives 2a and 3a.

Conclusions

Click to copy section linkSection link copied!

In summary, we have shown that efficient catalytic systems generated from betaine 1a or their cationic alkyl derivatives 2a and 3a provide a versatile and clean method for the conversion of DCM in a range of formaldehyde derivatives CH2Z2. We have successfully applied this transformation to aliphatic and aromatic alcohols, thiols, and N-heterocycles, which can be used either as insoluble sodium salts or directly, in combination with suitable bases like NaOH or NaH, depending on their acid strengths. This process only generates sodium chloride and water or hydrogen as byproducts, whose straightforward elimination greatly facilitates the purification of the products. Furthermore, DCM transformation was achieved at very low catalyst loadings and required no transition metals. The volatility of dichloromethane allows its facile recovery and recirculation, enabling facile scale-up of the reactions. Mechanistic studies have ruled out an organocatalytic cycle involving the reversible formation and cleavage of C–N bonds of 2a, which rather acts as a precursor for electroneutral, highly efficient Na+ complexing isoureate ethers arising from an unusual nucleophilic displacement of the ICy (the NHC fragment of the betaine) by the nucleophilic reagent. We are currently extending our studies to other NHC-CDI derivatives and investigating the potential uses of these compounds in catalysis, either free or complexed to abundant and low-toxicity metals.

Experimental Section

Click to copy section linkSection link copied!

General Considerations

All manipulations were performed under an inert atmosphere using Schlenk-line techniques (O2 < 3 ppm) and a glovebox (O2 < 0.6 ppm) MBraun MB-20G. Solvents were purified using an MBraun Solvent Purification System, except dichloromethane (DCM) and THF, which were distilled with CaH2 and Na, respectively. All NMR scale experiments were carried out in sample tubes with airtight PTFE valves. Deuterated solvents were degassed and stored in the glovebox in the presence of molecular sieves (4 Å). NMR spectra were recorded with a Bruker 400 Ultrashield (1H 400 MHz, 13C 101 MHz) at 25 °C. All chemical shifts were determined using residual signals of solvents and were referenced with regard to external SiMe4. Assignments of spectral signals were helped with 2D (1H–13C HSQC and HMBC) and diffusion (DOSY) NMR experiments. Elemental analysis and ESI-MS spectra of samples were carried out by the Analytical Services of the Institute for Chemical Research (Seville, Spain) using an LECO CHNS-TruSpec and Bruker Ion Trap Bruker Esquire 6000, respectively. HR-ESI spectra were recorded by the Mass Spectrometry Service (CITIUS, University of Seville) in a Thermo Scientific Orbitrap Elite hybrid mass spectrometer, operating in direct injection mode with an ESI ion source and ion-trap analyzer. Unless otherwise specified, all commercial reagents were purchased from Sigma-Aldrich and used as received. Imidazolium salt ICy·HBF4 and catalysts 1a and 2a were prepared according to literature and our synthetic procedures, previously reported. (17a,21,36) Warning! Although dichloromethane does not react vigorously with NaH, NaOH, or sodium alkoxides or aryloxides described in this work, strongly basic reagents like potassium tert-butoxide or neat sodium, which are known to react violently with this solvent, should be avoided.

General Procedure

A regular magnetic bar sufficed to achieve a satisfactory mixing of the suspensions throughout the whole experiment. DCM, containing the required amount of catalyst, was added to a gastight glass Teflon valve glass ampule containing solid Na+Z (either preformed or generated in situ from HZ and an inorganic base), and the mixture was stirred for a prescribed time in an oil bath preset at the specified temperature. In all experiments, the mixture gradually takes an intense yellow-orange color, as the suspended solid gradually becomes a thinner precipitate (NaCl). At the prescribed time, the mixture was allowed to settle, and a small sample (0.1–0.05 mL) was taken and dissolved in CDCl3. The conversion was deduced from the relative ratios of the central CH2 signal of the product and an internal standard (hexamethylbenzene). See below for more details and the purification procedures.

Preliminary Experiments

With NaOH, an excess (1 g) of NaOH was added to a vial, and 10 mL of a solution of 20 mg of catalyst 2a in neat dichloromethane was added to the solid. After 24 h, a free formaldehyde smell was detected, and an aliquot of the mixture was taken. The NMR spectrum of the mixture in CDCl3 showed a broad signal for hydrated formaldehyde oligomers (see Figure S25), but the efficiency of the reaction was hard to quantify.
With NaOMe in an NMR tube, catalyst 2a (5 mg, 2 mol %) was dissolved in CD2Cl2, and then anhydrous NaOMe (27 mg) was added inside the NMR tube. It was not soluble and remained in the bottom of the tube. After 2 h at room temperature, the solution became light orange, and it got more intense over time. It was also perceived a change in the texture of the solid, which became thinner. The 1H NMR spectrum showed an intense singlet at 3.30 ppm, which was assigned to 1H of methoxy groups from dimethoxymethane (methylal). As the methylene source was CD2Cl2, the methylene fragment of the product was not observed.

Optimization of the General Reaction Conditions for the Syntheses of Methylal from DCM and NaOMe with Catalyst 2a

In the general procedure (see Table S1 for details), a regular magnetic bar sufficed to achieve a satisfactory mixing of the suspensions throughout the whole experiment. DCM, containing the required amount of 2a, was added to a Schlenk flask containing solid NaOMe, and the mixture was stirred for a prescribed time. In the experiments conducted at 60 °C, the mixture was transferred to a gastight glass Teflon valve glass ampule, which was closed and magnetically stirred in an oil bath preset at the specified temperature. In all experiments, the mixture gradually takes an intense yellow-orange color, as the suspended solid gradually becomes a thinner precipitate (NaCl). At the prescribed time, the mixture was allowed to settle, and a small sample (0.1–0.05 mL) was taken and dissolved in CDCl3. The conversion was deduced from the relative ratios of the CH2 signals of unreacted dichloromethane and methylal. The mixture was then allowed to settle, filtered with a cannula, and vacuum-transferred to a cold trap, which affords a clean, colorless solution of methylal in the remaining dichloromethane. The mixture was a colorless liquid. 1H NMR (CDCl3, 400 MHz): δ 4.52 (s, 2H), 3.30 (s, 6H). 13C{1H} NMR (CDCl3, 100 MHz): δ 97.7, 55.2. Spectroscopic data is in accordance with data reported in literature. (37a)

Monitoring the Reaction of NaOMe with CH2Cl2

A 50 mL glass reactor provided with a Young Teflon screw stopcock and a nitrogen-purged liner to allow sample removal was charged with solid NaOMe (2 g, 37 mmol), DCM (35 mL), and 5 mL of a DCM solution containing 37 mg of catalyst 2a (0.2 mol %), and an accurately weighed amount (30 mg) of hexamethylbenzene to be used as an internal standard, weigh to ±0.1 mg accuracy in an analytical balance. The mixture was stirred at 40 °C in an oil bath preset at the specified temperature, taking aliquots of ca. 0.1 mL with a long-needle syringe through the sample removal port. Each sample was diluted with 0.5 mL of CDCl3, filtered, and analyzed using NMR to determine the reaction advance. The NMR data were positively compared with those listed in the SDBS spectral database in the same solvent (CD3Cl). (37a)

Comparative Assessment of Catalyst Efficiency in the Methylal Synthesis Using Different Catalysts and Alkali Metal Cations

In order to study the effect of the catalyst in the obtention of methylal, we repeated the reaction under the same conditions described in the previous section, using the same molar amount of a different catalyst in each of them: NBu4Cl, 18-crown-6, 15-crown-5, 1a, 3a or ICy·HBF4, CDI(p-Tol). These reactions were not monitored but stopped after a fixed reaction time (4 h at 40 °C, when 50% conversion NaOMe was estimated using 2a), and their progress was determined using the same NMR-based methodology to determine the amount of methylal formed. Likewise, the efficiency of 2a with different alkali metal cations was assessed using equivalent amounts (37 mmol) of the corresponding alkali methoxides MOMe (M = Li, Na, or K) and a 4 h reaction time. The characteristic dark orange color associated with these reactions was not observed when LiOMe was used as the methoxide source.

Procedure for the Generation of Solid Sodium Benzyloxide and Sodium Phenoxide

Sodium benzyloxide or phenoxide was generated prior to use, following a modified procedure taken from the literature: (38) a 100 mL Schlenk flask equipped with a stirring bar was charged with NaH (82.5 mmol) in mineral oil and 25 mL of THF. The flask was brought to an ice–water bath, and the solid was suspended with magnetic stirring. A solution containing 75 mmol of the corresponding alcohol or phenol in 25 mL of THF was kept cool with vigorous stirring. The addition caused the evolution of H2 gas, more vigorous in the case of phenol. Once the addition was completed, the mixture was kept stirring at room temperature for 4 h. The solvent was removed under a vacuum, and the sticky white solid obtained was washed three times with hexane, in order to remove the mineral oil, to afford a white powder. No further purification was performed.

Synthesis of CH2(OPh)2 from Sodium Phenoxide

Sodium phenoxide (3.83 g, 33 mmol), generated from phenol and NaH as described above, was added to an ampule with an airtight PTFE valve and suspended in 15 mL of dried dichloromethane. A solution of 33 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. After a few minutes at 60 °C, an intense yellow-orange color in the solution was observed. Once the reaction finished, the stirring was stopped. A light white solid (NaCl) remained in the bottom of the flask, leaving the colored solution clear to perform the purification. The solution was separated from the solid by filtration via cannula. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst, and afterward, the solvent was removed under a vacuum to afford a slightly pale-yellow liquid with a yield of 92% (3.04 g). The same procedure was conducted with 16.5 mg of 1a as a catalyst, with a yield of 86% (2.84 g). 1H NMR (CDCl3, 400 MHz): δ 7.32 (m, spin system AA′BB′C, 4H), 7.13 (m, spin system AA′BB′C, 4H), 7.05 (tt, J = 7.4 Hz, J = 1.1 Hz, 2H), 5.75 (s, 2H). 13C{1H} NMR (CDCl3, 100 MHz): δ 157.2, 129.7, 122.6, 116.7, 91.4. IR (cm–1): 3041 (ν CAr─H), 2973 (ν Csp3─H), 1588 (ν CAr═CAr), 1490 (ν CAr═CAr), 1200 (νas CAr─O-C), 1012 (νs CAr–O-C). Spectroscopic data is in accordance with data reported in literature. (37b)

Synthesis of CH2(OBn)2 from Sodium Benzyloxide

Sodium benzyloxide (2.18 g, 16.5 mmol), generated from benzyl alcohol and NaH as described above, was added to an ampule with an airtight PTFE valve and suspended in 15 mL of dried dichloromethane. A solution of 16.5 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. After a few minutes at 60 °C, an intense orange-red color in the solution was observed. Once the reaction finished, the stirring was stopped. A light solid white solid (NaCl) remained in the bottom of the flask, leaving the colored solution clear to perform the purification. The solution was separated from the solid by filtration via cannula. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst, and afterward, the solvent was removed under a vacuum to afford a slightly pale-yellow liquid, with a yield of 87% (1.64 g). The same procedure was conducted with 8.25 mg of 1a as a catalyst with a yield of 87% (1.64 g). 1H NMR (CDCl3,400 MHz): δ 7.37 (m, 8H). 7.32 (m, 2H), 4.86 (s, 2H), 4.67 (s, 4H). 13C{1H} NMR (100 MHz, CDCl3, 25 °C): δ 138.0, 128.5, 128.1, 127.8, 94.1, 69.7. IR (cm─1): 3029 (ν CAr─H), 2879 (ν Csp3–H), 1497 (ν CAr═CAr), 1102 (νas C─O─C), 1041 (νs C─O–C). Spectroscopic data is in accordance with data reported in literature. (37c)

Synthesis of CH2(OMe)2 from Methanol and Sodium Hydroxide

Sodium hydroxide (1.7 g, 44.5 mmol) was ground and added to an ampule with an airtight PTFE valve and suspended on 25 mL of dried dichloromethane. Afterward, 1.19 g of methanol (37 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium methoxide. A solution of 37 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. The stirring was quite hindered at the beginning due to the formation of a water phase. After a few minutes at 60 °C, an intense yellow-orange color in the solution was observed. Once the reaction finished, the stirring was stopped. The organic layer was separated from the solid by filtration via cannula. The yellow mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst to afford a colorless solution of CH2(OMe2)2 in dichloromethane with a yield of 89%. 1H NMR (CDCl3, 400 MHz): δ 4.52 (s, 2H), 3.30 (s, 6H). 13C{1H} NMR (CDCl3, 100 MHz): δ 97.7, 55.2. Spectroscopic data is in accordance with data reported in literature. (37a)

Synthesis of CH2(OPh)2 from Phenol and Sodium Hydroxide

Sodium hydroxide (0.37 g, 9.25 mmol) was ground and added to an ampule with an airtight PTFE valve and suspended in 15 mL of dried dichloromethane. Afterward, 0.78 g of phenol (8.25 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium phenoxide. A solution of 8.25 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. The stirring was quite hindered at the beginning due to the formation of a water phase. After a few minutes at 60 °C, an intense yellow-orange color in the solution was observed. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst. Afterward, the solvent was removed under a vacuum to afford a white powder in 89% yield (0.74 g). 1H NMR (CDCl3, 400 MHz): δ 7.32 (m, 4H). 7.13 (m, 4H), 7.05 (tt, J = 7.4 Hz, J = 1.1 Hz, 2H), 5.75 (s, 2H).13C{1H} NMR (CDCl3, 100 MHz): δ 157.2, 129.7, 122.6, 116.7, 91.4. IR (cm–1): 3041 (ν CAr─H), 2973 (ν Csp3─H), 1588 (ν CAr═CAr), 1490 (ν CAr═CAr), 1200 (νas CAr─O─C), 1012 (νs CAr─O─C). Spectroscopic data is in accordance with data reported in literature. (37b)

Synthesis of CH2(OBn)2 from Benzyl Alcohol and Sodium Hydroxide

Sodium hydroxide (0.37 g, 9.25 mmol) was ground and added to an ampule with an airtight PTFE valve and suspended in 15 mL of dried dichloromethane. Afterward, 0.89 g of benzyl alcohol (8.25 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium benzyloxide. A solution of 8.25 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. The stirring was quite hindered at the beginning due to the formation of a water phase. After a few minutes at 60 °C, an intense yellow-orange color in the solution was observed. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst. Afterward, the solvent was removed under a vacuum to afford a white powder in 89% yield (0.84 g). 1H NMR (CDCl3, 400 MHz): δ 7.37 (m, 8H), 7.32 (m, 2H), 4.86 (s, 2H), 4.67 (s, 4H). 13C{1H} NMR (100 MHz, CDCl3, 25 °C): δ 138.0, 128.5, 128.1, 127.8, 94.1, 69.7. IR (cm─1): 3029 (ν CAr─H), 2879 (ν Csp3–H), 1497 (ν CAr═CAr), 1102 (νas C─O─C), 1041 (νs C─O–C). Spectroscopic data is in accordance with data reported in literature. (37c)

Synthesis of CH2(O-p-C6H4t-Bu)2 from p-tert-Butylphenol and Sodium Hydroxide

Sodium hydroxide (0.37 g, 9.25 mmol) was ground and added to an ampule with an airtight PTFE valve and suspended in 15 mL of dried dichloromethane. Afterward, 1.15 g of p-tert-butylphenol (7.66 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium p-tert-butylphenoxide. A solution of 15 mg of catalyst 2a (0.4 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. The stirring was quite hindered at the beginning due to the formation of a water phase. After a few minutes at 60 °C, an intense yellow-orange color in the solution. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst. Afterward, the solvent was removed under a vacuum to afford a white powder, in 85% yield (1.02 g). 1H NMR (CDCl3, 400 MHz): δ 7.31 (m, 4H), 7.04 (m, 4H), 5.69 (s, 2H, CH2), 1.29 (s, 18H). 13C{1H} NMR (CDCl3, 100 MHz): δ 155.0,145.3,126.5, 116.1, 91.6, 34.3, 31.6. IR (cm─1): 3041 (ν CAr─H), 2957 (ν Csp3─H), 1509 (ν CAr═CAr), 1211 (νas CAr─O─C), 1016 (νs CAr─O─C). Anal. Calcd for C21O28O2: C, 80.73%; H, 9.03%. Found: C, 81.16%; H, 8.87%.

Synthesis of CH2(OCH2CH═CH2)2 from Allyl Alcohol and Sodium Hydroxide

Sodium hydroxide (0.40 g, 9.25 mmol) was ground and added to an ampule with an airtight PTFE valve and suspended in 15 mL of dried dichloromethane. Afterward, 0.48 g of allyl alcohol (8.25 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium allyloxide. A solution of 8.25 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. The stirring was quite hindered at the beginning due to the formation of a water phase. After a few minutes at 60 °C, an intense yellow-orange color in the solution was observed. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. In order to remove the unreacted alcohol, 500 mg of sodium hydroxide was added to the solution, and it was stirred for 1 h. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst. Afterward, the solvent was removed by heating slightly to afford a pale-yellow liquid in 83% yield (0.44 g). 1H NMR (CDCl3, 400 MHz): δ 5.92 (ddt, J = 17.2 Hz, J = 10.4, J = 5.6 Hz), 5.29 (m, J = 17.2 Hz, J = 1.6 Hz, 2H), 5.18 (m, J = 10.4 Hz, J = 1.6 Hz, 2H), 4.72 (s, 2H), 4.08 (m, J = 5.6 Hz, J = 1.5 Hz, 4H). 13C{1H} NMR (CDCl3, 100 MHz): δ 134.5,117.2, 93.9, 68.5. IR (cm─1): 3081 (ν Csp2─H), 2881 (ν Csp3─H), 1647 (ν C═C), 1105 (νas C─O─C), 1041 (νs C─O─C), 918 (δ C═CH2). Spectroscopic data is in accordance with data reported in literature. (37c)

Synthesis of CH2(O-i-Pr)2 from 2-Propanol and Sodium Hydride

Sodium hydride (0.40 g, 9.25 mmol) was added to an ampule with an airtight PTFE valve and washed with hexane in order to remove the oil suspension. The dried solid was suspended in 15 mL of dichloromethane. Afterward, 0.50 g of 2-propanol (8.25 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium 2-propoxide. A solution of 8.25 mg of catalyst 2a (0.2 mol %) in 5 mL of dried dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. After a few minutes at 60 °C, an intense yellow-orange color in the solution was observed. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. In order to remove the unreacted alcohol, 500 mg of sodium hydroxide was added to the solution, and it was stirred for 1 h. The mixture was filtrated through a pad of silica gel in order to remove the traces of the catalyst. Afterward, the solvent was removed by heating slightly to afford a pale-yellow liquid in 67% yield (0.36 g). 1H NMR (CDCl3, 400 MHz): δ 4.72 (s, 2H), 3.90 (sep, J = 6.2, 2H), 1.17 (d, J = 6.2, 12H). 13C{1H} NMR (CDCl3, 100 MHz): δ 90.9,68.6, 22.6. IR (cm–1): 2921 (ν Csp3─H), 1033 (νs C─O─C). Spectroscopic data is in accordance with data reported in literature. (37e)

Synthesis of CH2(SEt)2 from Ethanethiol and Sodium Hydroxide

Sodium hydroxide (0.55 g, 13.75 mmol) was ground and added to an ampule with an airtight PTFE valve and suspended in 15 mL of dichloromethane. Afterward, 0.80 g of ethanethiol (12.9 mmol) was added to the mixture, which caused the appearance of a white sticky solid, sodium ethanethiolate. A solution of 25 mg of catalyst 2a (0.4 mol %) in 5 mL of dichloromethane was added with a syringe directly upon the suspension. The resulting colorless mixture was stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. The stirring was quite hindered at the beginning due to the formation of a water phase. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. No further purification was performed. Conversion >99% by NMR was found. 1H NMR (CDCl3, 400 MHz): δ 3.69 (s, 2H), 2.65 (q, J = 7.4 Hz, 4H), 1.26 (t, J = 7.4 Hz, 6H). Spectroscopic data is in accordance with data reported in literature. (37f)

Synthesis of CH2Pz2 from Pyrazole and Sodium Hydride

Sodium hydride (0.60 g, 24.9 mmol) was added to an ampule with an airtight PTFE valve and washed with hexane in order to remove the oil suspension. The dried solid was suspended in 15 mL of dichloromethane. Afterward, 1.13 g (16.6 mmol) of pyrazole was added to the suspension. The addition led to the formation of sodium pyrazolate and hydrogen. A solution of 16.5 mg of catalyst 2a (0.2 mol %) in 5 mL of dichloromethane was added with a syringe directly upon the suspension, and it is stirred at 60 °C for 24 h in an oil bath preset at the specified temperature. After a few minutes at 60 °C, a brownish white color in the solution was observed. Once the reaction finished, the stirring was stopped. The organic solution was separated from the solid by filtration via cannula. The solvent was removed under a vacuum to afford a pale-brown solid. The solid is purified by crystallization in heptane, where the product is soluble at high temperatures, and the impurities remain insoluble, to afford a white solid, in 72% yield (0.88 g). 1H NMR (CDCl3, 400 MHz): δ 7.65 (dd, J = 2.4 Hz, J = 0.6 Hz, 2H), 7.55 (dd, J = 1.9 Hz, J = 0.6 Hz, 2H), 6.31 (s, 2H), 6.29 (dd, J = 2.4 Hz, J = 1.9 Hz, 2H). 13C{1H} NMR (CDCl3, 100 MHz): δ 141.0, 129.8, 107.3, 65.5. IR (cm─1): 3101 (ν CAr–H), 2924 (ν Csp3─H), 1610 (ν CAr═N), 1514 (ν CAr═CAr), 1435 (ν CAr═CAr). Spectroscopic data is in accordance with data reported in literature. (37g)

Synthesis of [MeOCH2·1a]+[Br] (3a)

To a 50 mL Schlenk flask was added 100 mg (0.22 mmol) of 1a. The yellow solid was dissolved in 10 mL of dried dichloromethane to afford a clear yellow solution. Then, 20 μL (0.22 mmol, 90%) of bromomethyl methyl ether was added, and the solution was stirred for 30 min while the color faded. Afterward, 10 mL of dried hexane was added to the solution, and the solvent was removed under a vacuum. The solid was washed with more hexane, to afford yellowish solid with a yield of 94% (0.12 g). 1H NMR (CD2Cl2, 400 MHz): δ 8.30 (s, 2H), 7.15 (m, 2H), 7.04 (m, 4H), 6.43 (m, 2H), 5.25 (br s, 2H), 4.06 (m, 2H), 3.47 (s, 3H), 2.30 (s, 3H), 2.25 (s, 3H), 1.90 (m, 2H), 1.79 (m, 4H), 1.67 (m, 4H), 1.50 (m, 2H), 1.34 (m, 4H), 1.23 (m, 2H), 0.99 (m, 2H). 13C{1H} NMR (CD2Cl2, 100 MHz): δ 143.3, 138.4, 136.1, 134.1, 130.8, 130.7, 124.7, 123.1, 121.0, 82.6, 60.4, 57.5, 33.5, 33.3, 25.9, 25.8, 24.7, 21.0, 20.9. Elemental analysis calcd for C32H43ON4Br: C, 66.31; H, 7.48; N, 9.67. Found: C, 66.46; H, 7.22; N, 9.43. ESI-MS (1 μM in anhydrous THF): m/z 499.56 (MeOCH2·1a+). HRMS (ESI) m/z: [M]+ calcd for C32H43N4O, 499.3431; found, 499.3423.

Spectroscopic Data of 1b

In order to detect and characterize 1b, the general procedure of catalysis was applied to obtain methylal starting from NaOMe. Once the reaction showed an intense change of color to a deep orange-yellow, the mixture was filtered, and the liquid phase was dried under a vacuum. An oily deep red solid was obtained. No further purification was performed. 1H NMR (CD2Cl2, 400 MHz): δ 6.98 (m, 2H). 6.93 (m, 2H), 6.90 (m, 2H), 6.65 (m, 2H), 4.69 (s, 2H), 3.75 (s, 3H), 3.30 (s, 3H), 2.24 (s, 3H), 2.20 (s, 3H). 13C{1H} NMR (CD2Cl2, 100 MHz): δ 152.5, 145.2, 141.8, 134.3, 131.4, 129.5, 123.5, 121.9, 82.4, 55.7,55.7, 20.8, 20.8. ESI-MS (1 μM THF/wet methanol 1:100): m/z 299.30 (1b·H+, 100%); 321.30 (1b·Na+, 73%); 337.33 (1b·K+, 13%). HRMS (ESI) m/z: [M + H]+ calcd for C18H23N2O2, 299.1754; found, 299.1753. HRMS (ESI) m/z: [M + Na]+ calcd for C18H22N2O2Na, 321.1573; found, 321.1770.

Spectroscopic Data of 2b

In order to detect and characterize 2b, the general procedure of catalysis was applied to obtain methylal starting from NaOMe. Once the reaction showed an intense change of color to a deep orange-yellow, the mixture was filtered, and the liquid phase was dried under a vacuum. An oily, deep red solid was obtained. No further purification was performed. 1H NMR (CD2Cl2, 400 MHz): δ 6.97 (m, 4H), 6.88 (m, 4H), 6.86 (m, 4H), 6.54 (m, 4H), 5.15 (s, 2H), 3.53 (s, 6H), 2.25 (s, 6H), 2.20 (s, 6H). 13C{1H} NMR (CD2Cl2, 100 MHz): δ 152.5, 145.5, 141.1, 135.2, 131.2, 129.5, 129.4, 125.9, 122.2, 67.1, 55.9, 20.9, 20.8. ESI-MS (1 μM in THF/wet methanol 1:100): m/z 521.51 (2b·H+, 100%); 543.50 (2b·Na+, 19%); 559.45 (2b·K+, 7%). HRMS (ESI) m/z: [M + H]+ calcd for C33H37N4O2, 521.2911; found, 521.2902. HRMS (ESI) m/z: [M + Na]+ calcd for C33H36N4O2Na, 343.2730; found, 343.2719.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c01971.

  • Additional experimental procedures and spectroscopic data (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Author
    • David Sánchez-Roa - Departamento de Química Orgánica y Química Inorgánica, Instituto de Investigación en Química “Andrés M. del Río” (IQAR) Universidad de Alcalá, Campus Universitario, Alcala de Henares, Madrid 28871, Spain
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was supported by the Spanish Research Agency (AEI), the European Union (Feder Funds), Junta de Andalucía, and the University of Alcalá through projects PGC2018-095768-B-100, RTI2018-094840-B-C31, UAH-AE-2017-2, and PY20_00104. D.S.R. thanks the Spanish Government for a Predoctoral Fellowship. We gratefully acknowledge Tomás G. Santiago, M. Sc. (IIQ), and Dr. Gloria Guitérrez Alcalá (Mass Spectrometry Service, IIQ) for their skillful assistance in sample preparation and recording ESI-MS spectra.

Abbreviations

Click to copy section linkSection link copied!

NHC

N-heterocyclic carbene

CDI

carbodiimide

DCM

dichloromethane

ICy

1,3-dicyclohexylimidazolylidene

PTBP

p-tert-butylphenol

AA

allyl alcohol

HPz

pyrazole

TON

turnover number

TOF

turnover frequency

NMR

nuclear magnetic resonance

DOSY

diffusion-ordered spectroscopy

HSQC

heteronuclear single quantum coherence spectroscopy

HMBC

heteronuclear multiple bond correlation

References

Click to copy section linkSection link copied!

This article references 38 other publications.

  1. 1
    (a) Rossberg, M.; Lendle, W.; Pfleiderer, G.; Tögel, A.; Torkelson, T. R.; Beutel, K. K. Chloromethanes. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley, 2011. DOI:  DOI: 10.1002/14356007.a06_233.pub3 .
    (b) Tyner, T.; Francis, J. Dichloromethane. In ACS Reagent Chemicals; American Chemical Society, 2017. DOI:  DOI: 10.1021/acsreagents.4117 .
  2. 2
    (a) Lee, H. M.; Lu, C. Y.; Chen, C. Y.; Chen, W. L.; Lin, H. C.; Chiu, P. L.; Cheng, P. Y. Palladium complexes with ethylene-bridged bis(N-heterocyclic carbene) for C–C coupling reactions. Tetrahedron 2004, 60, 58075825,  DOI: 10.1016/j.tet.2004.04.070
    (b) Noujeim, N.; Leclercq, L. C.; Schmitzer, A. R. N,Nʼ-Disubstituted Methylenediimidazolium Salts: A Versatile Guest for Various Macrocycles. J. Org. Chem. 2008, 73, 37843790,  DOI: 10.1021/jo702683c
    (c) Cheong, M.; Kim, M.-N.; Shim, J. Y. Rhodium-catalyzed double carbonylation of diiodomethane in the presence of triethylorthoformate. J. Organomet. Chem. 1996, 520, 253255,  DOI: 10.1016/0022-328X(96)06349-8
    (d) Zadykowicz, J.; Potvin, P. G. N-(2-tetrahydrofuranyl)azole nucleoside analogs by reactions of azoles with dihalomethanes in tetrahydrofuran. J. Heterocycl. Chem. 1999, 36, 623626,  DOI: 10.1002/jhet.5570360308
    (e) Zidan, A.; Garrec, J.; Cordier, M.; El-Naggar, A. M.; Abd El-Sattar, N. E. A.; Ali, A. K.; Hassan, M. A.; El Kaim, L. β-Lactam Synthesis through Diodomethane Addition to Amide Dianions. Angew. Chem., Int. Ed. 2017, 56, 1217912183,  DOI: 10.1002/anie.201706315
  3. 3
    (a) Durandetti, S.; Sibille, S.; Perichon, J. Electrochemical cyclopropanation of alkenes using dibromomethane and zinc in dichloromethane/DMF mixture. J. Org. Chem. 1991, 56, 32553258,  DOI: 10.1021/jo00010a015
    (b) Hon, Y.-S.; Hsieh, C.-H.; Liu, Y.-W. Dibromomethane as one-carbon source in organic synthesis: total synthesis of (±)- and (−)-methylenolactocin. Tetrahedron 2005, 61, 27132723,  DOI: 10.1016/j.tet.2005.01.057
    (c) Brunner, G.; Eberhard, L.; Oetiker, J.; Schröder, F. Cyclopropanation with Dibromomethane under Grignard and Barbier Conditions. Synthesis 2009, 21, 37083718,  DOI: 10.1055/s-0029-1216999
    (d) Raposo, C. D. C. B. G. Diiodomethane: A Versatile C1 Building Block. Synlett 2013, 24, 17371738,  DOI: 10.1055/s-0033-1338964
  4. 4

    Though rather inert, DCM can undergo different reactions when used as a solvent, e.g., through free radical routes; see:

    (a) Levina, I. I.; Klimovich, O. N.; Vinogradov, D. S.; Podrugina, T. A.; Bormotov, D. S.; Kononikhin, A. S.; Dement’eva, O. V.; Senchikhin, I. N.; Nikolaev, E. N.; Kuzmin, V. A.; Nekipelova, T. D. Dichloromethane as solvent and reagent: a case study of photoinduced reactions in mixed phosphonium-iodonium ylide. J. Phys. Org. Chem. 2018, 31, e3844,  DOI: 10.1002/poc.3844

    and references cited therein

    In addition, DCM is known to react violently with some strong bases like potassium t-butoxide or some reactive metals (e.g., Li or Na/K alloy); see:

    (b) Lewis, R. J., Sr., Ed.; Sax’s Dangerous Properties of Industrial Materials, 11th ed.; Wiley-Interscience, Wiley & Sons, Inc: Hoboken, NJ, 2004; p 2436.
    (c) Anhydrous DCM is not corrosive for most metals, but in the presence of water, it slowly hydrolyzes releasing small amounts of HCl, which can attack stainless steel, aluminum, or even copper pipes. See any dichlorometane Safety Data Sheet (MSDS, e.g., Fisher, ACC no. 14930).

    Moreover, there are reports concerning dichloromethane C–Cl bond substitutions but are usually very slow processes, see:

    (d) Rudine, A. B.; Walter, M. G.; Wamser, C. C. Reaction of Dichloromethane with Pyridine Derivatives under Ambient Conditions. J. Org. Chem. 2010, 75, 42924295,  DOI: 10.1021/jo100276m
  5. 5
    (a) Closs, G. L. Carbenes from Alkyl Halides and Organolithium Compounds. IV. Formation of Alkylcarbenes from Methylene Chloride and Alkyllithium Compounds. J. Am. Chem. Soc. 1962, 84, 809813,  DOI: 10.1021/ja00864a026
    (b) Böhm, A.; Bach, T. Radical Reactions Induced by Visible Light in Dichloromethane Solutions of Hünig’s Base: Synthetic Applications and Mechanistic Observations. Chem. - Eur. J. 2016, 22, 1592115928,  DOI: 10.1002/chem.201603303
    (c) Sturala, J.; Ambrosi, A.; Sofer, Z.; Pumera, M. Covalent Functionalization of Exfoliated Arsenic with Chlorocarbene. Angew. Chem., Int. Ed. 2018, 57, 1483714840,  DOI: 10.1002/anie.201809341
  6. 6
    Young, J. A. Dichloromethane. J. Chem. Educ. 2004, 81, 1415,  DOI: 10.1021/ed081p1415
  7. 7
    (a) Salthammer, T.; Mentese, S.; Marutzky, R. Formaldehyde in the Indoor Environment. Chem. Rev. 2010, 110, 25362572,  DOI: 10.1021/cr800399g
    (b) Li, W.; Wu, X.-F. The Applications of (Para)formaldehyde in Metal-Catalyzed Organic Synthesis. Adv. Synth. Catal. 2015, 357, 33933418,  DOI: 10.1002/adsc.201500753
    (c) Pathak, D. D.; Gerald, J. J. An Efficient and Convenient Method for the Synthesis of Dialkoxymethanes Using Kaolinite as a Catalyst. Synth. Commun. 2003, 33, 15571561,  DOI: 10.1081/SCC-120018774
  8. 8
    Hossaini, R.; Chipperfield, M. P.; Montzka, S. A.; Leeson, A. A.; Dhomse, S. S.; Pyle, J. A. The increasing threat to stratospheric ozone from dichloromethane. Nat. Commun. 2017, 8, 15962,  DOI: 10.1038/ncomms15962
  9. 9
    (a) Kane, A.; Giraudet, S.; Vilmain, J.-B.; Le Cloirec, P. Intensification of the temperature-swing adsorption process with a heat pump for the recovery of dichloromethane. J. Environ. Chem. Eng. 2015, 3, 734743,  DOI: 10.1016/j.jece.2015.02.021
    (b) Wu, X.-F. W.; Tlili, A.; Schranck, J.; Wu, X.-F. W. The Application of Dichloromethane and Chloroform as Reagents in Organic Synthesis. Solvents as Reagents in Organic Synthesis 2017, 125159,  DOI: 10.1002/9783527805624.ch4
  10. 10
    (a) Lautenschütz, L.; Oestreich, D.; Seidenspinner, P.; Arnold, U.; Dinjus, E.; Sauer, J. Physico-chemical properties and fuel characteristics of oxymethylene dialkyl ethers. Fuel 2016, 173, 129137,  DOI: 10.1016/j.fuel.2016.01.060
    (b) Thenert, K.; Beydoun, K.; Wiesenthal, J.; Leitner, W.; Klankermayer, J. Ruthenium-Catalyzed Synthesis of Dialkoxymethane Ethers Utilizing Carbon Dioxide and Molecular Hydrogen. Angew. Chem., Int. Ed. 2016, 55, 1226612269,  DOI: 10.1002/anie.201606427
    (c) Schieweck, B. G.; Klankermayer, J. Tailor-made Molecular Cobalt Catalyst System for the Selective Transformation of Carbon Dioxide to Dialkoxymethane Ethers. Angew. Chem., Int. Ed. 2017, 56, 1085410857,  DOI: 10.1002/anie.201702905
  11. 11
    (a) El-Brollosy, N. R.; Jørgensen, P. T.; Dahan, B.; Boel, A. M.; Pedersen, E. B.; Nielsen, C. Synthesis of Novel N-1 (Allyloxymethyl) Analogues of 6-Benzyl-1-(ethoxymethyl)-5-isopropyluracil (MKC-442, Emivirine) with Improved Activity Against HIV-1 and Its Mutants. J. Med. Chem. 2002, 45, 57215726,  DOI: 10.1021/jm020949r
    (b) Wamberg, M.; Pedersen, E. B.; El-Brollosy, N. R.; Nielsen, C. Synthesis of 6-arylvinyl analogues of the HIV drugs SJ-3366 and Emivirine. Bioorg. Med. Chem. 2004, 12, 11411149,  DOI: 10.1016/j.bmc.2003.11.032
    (c) Loksha, Y. M.; Pedersen, E. B.; Loddo, R.; Sanna, G.; Collu, G.; Giliberti, G.; Colla, P. L. Synthesis of Novel Fluoro Analogues of MKC442 as Microbicides. J. Med. Chem. 2014, 57, 51695178,  DOI: 10.1021/jm500139a
  12. 12
    (a) Makosza, M.; Sypniewski, M. Reaction of sulfonium salts of formaldehyde dithioacetals with aromatic aldehydes and rearrangements of the produced thioalkyl oxiranes. Tetrahedron 1995, 51, 1059310600,  DOI: 10.1016/0040-4020(95)00632-I
    (b) Luh, T.-Y. Recent advances on the synthetic applications of the dithioacetal functionality. J. Organomet. Chem. 2002, 653, 209214,  DOI: 10.1016/S0022-328X(02)01151-8
  13. 13
    (a) Otero, A.; Fernández-Baeza, J.; Antiñolo, A.; Tejeda, J.; Lara-Sánchez, A. Heteroscorpionate ligands based on bis(pyrazol-1-yl)methane: design and coordination chemistry. Dalton Trans. 2004, 10, 14991510,  DOI: 10.1039/B401425A
    (b) Krieck, S.; Koch, A.; Hinze, K.; Müller, C.; Lange, J.; Görls, H.; Westerhausen, M. s-Block Metal Complexes with Bis- and Tris(pyrazolyl)methane and -methanide Ligands. Eur. J. Inorg. Chem. 2016, 15, 23322348,  DOI: 10.1002/ejic.201501263
  14. 14
    (a) Csok, Z.; Vechorkin, O.; Harkins, S. B.; Scopelliti, R.; Hu, X. Nickel Complexes of a Pincer NN2 Ligand: Multiple Carbon-Chloride Activation of CH2Cl2 and CHCl3 Leads to Selective Carbon-Carbon Bond Formation. J. Am. Chem. Soc. 2008, 130, 81568157,  DOI: 10.1021/ja8025938
    (b) Zhan, L.; Pan, R.; Xing, P.; Jiang, B. An efficient method for the preparation of dialkoxymethanes from dichloromethane with alcohols catalyzed by a Cu-NHC complex. Tetrahedron Lett. 2016, 57, 40364038,  DOI: 10.1016/j.tetlet.2016.07.056
  15. 15
    Lamb, J. R.; Brown, C. M.; Johnson, J. A. N-Heterocyclic carbene–carbodiimide (NHC–CDI) betaine adducts: synthesis, characterization, properties, and applications. Chem. Sci. 2021, 12, 26992715,  DOI: 10.1039/D0SC06465C
  16. 16
    Baishya, A.; Kumar, L.; Barman, M. K.; Peddarao, T.; Nembenna, S. Air Stable N-Heterocyclic Carbene-Carbodiimide (“NHC-CDI”) Adducts: Zwitterionic Type Bulky Amidinates. ChemistrySelect 2016, 1, 498503,  DOI: 10.1002/slct.201600019
  17. 17
    (a) Márquez, A.; Ávila, E.; Urbaneja, C.; Álvarez, E.; Palma, P.; Cámpora, J. Copper(I) Complexes of Zwitterionic Imidazolium-2-Amidinates, a Promising Class of Electroneutral, Amidinate-Type Ligands. Inorg. Chem. 2015, 54, 1100711017,  DOI: 10.1021/acs.inorgchem.5b02141
    (b) Baishya, A.; Kumar, L.; Barman, M. K.; Biswal, H. S.; Nembenna, S. N-Heterocyclic Carbene–Carbodiimide (“NHC–CDI”) Adduct or Zwitterionic-Type Neutral Amidinate-Supported Magnesium(II) and Zinc(II) Complexes. Inorg. Chem. 2017, 56, 95359546,  DOI: 10.1021/acs.inorgchem.7b00879
  18. 18
    (a) Martínez-Prieto, L. M.; Cano, I.; Márquez, A.; Baquero, E. A.; Tricard, S.; Cusinato, L.; del Rosal, I.; Poteau, R.; Coppel, Y.; Philippot, K.; Chaudret, B.; Cámpora, J.; van Leeuwen, P. W. N. M. Zwitterionic amidinates as effective ligands for platinum nanoparticle hydrogenation catalysts. Chem. Sci. 2017, 8, 29312941,  DOI: 10.1039/C6SC05551F
    (b) López-Vinasco, A. M.; Martínez-Prieto, L. M.; Asensio, J. M.; Lecante, P.; Chaudret, B.; Cámpora, J.; van Leeuwen, P. W. N. M. Novel nickel nanoparticles stabilized by imidazolium-amidinate ligands for selective hydrogenation of alkynes. Catal. Sci. Technol. 2020, 10, 342350,  DOI: 10.1039/C9CY02172H
  19. 19
    Gallagher, N. M.; Zhukhovitskiy, A. V.; Nguyen, H. V. T.; Johnson, J. A. Main-Chain Zwitterionic Supramolecular Polymers Derived from N-Heterocyclic Carbene–Carbodiimide (NHC–CDI) Adducts. Macromolecules 2018, 51, 30063016,  DOI: 10.1021/acs.macromol.8b00579
  20. 20
    Gallagher, N. M.; Ye, H.-Z.; Feng, S.; Lopez, J.; Zhu, Y. G.; Van Voorhis, T.; Shao-Horn, Y.; Johnson, J. A. An N-Heterocyclic-Carbene-Derived Distonic Radical Cation. Angew. Chem., Int. Ed. 2020, 59, 39523955,  DOI: 10.1002/anie.201915534
  21. 21
    Sánchez-Roa, D.; Santiago, T. G.; Fernández-Millán, M.; Cuenca, T.; Palma, P.; Cámpora, J.; Mosquera, M. E. G. Interaction of an imidazolium-2-amidinate (NHC-CDI) zwitterion with zinc dichloride in dichloromethane: role as ligands and C–Cl activation promoters. Chem. Commun. 2018, 54, 1258612589,  DOI: 10.1039/C8CC07661H
  22. 22
    (a) Rivlin, M.; Eliav, U.; Navon, G. NMR Studies of the Equilibria and Reaction Rates in Aqueous Solutions of Formaldehyde. J. Phys. Chem. B 2015, 119, 44794487,  DOI: 10.1021/jp513020y
    (b) Dankelman, W.; Daemen, J. M. H. Gas Chromatographic and Nuclear Magnetic Resonance Determination of Linear Formaldehyde Oligomers in Formaline. Anal. Chem. 1976, 48, 401404,  DOI: 10.1021/ac60366a030
  23. 23
    (a) Li, M.; Long, Y.; Deng, Z.; Zhang, H.; Yang, X.; Wang, G. Ruthenium trichloride as a new catalyst for selective production of dimethoxymethane from liquid methanol with molecular oxygen as sole oxidant. Catal. Commun. 2015, 68, 4648,  DOI: 10.1016/j.catcom.2015.04.031
    (b) Thavornprasert, K.-a.; Capron, M.; Jalowiecki-Duhamel, L.; Dumeignil, F. One-pot 1,1-dimethoxymethane synthesis from methanol: a promising pathway over bifunctional catalysts. Catal. Sci. Technol. 2016, 6, 958970,  DOI: 10.1039/C5CY01858G
    (c) Chao, Y.; Lai, J.; Yang, Y.; Zhou, P.; Zhang, Y.; Mu, Z.; Li, S.; Zheng, J.; Zhu, Z.; Tan, Y. Visible light-driven methanol dehydrogenation and conversion into 1,1-dimethoxymethane over a non-noble metal photocatalyst under acidic conditions. Catal. Sci. Technol. 2018, 8, 33723378,  DOI: 10.1039/C8CY01030G
  24. 24
    Łojewska, J.; Wasilewski, J.; Terelak, K.; Łojewski, T.; Kołodziej, A. Selective oxidation of methylal as a new catalytic route to concentrated formaldehyde: Reaction kinetic profile in gradientless flow reactor. Catal. Commun. 2008, 9, 18331837,  DOI: 10.1016/j.catcom.2008.02.014
  25. 25
    (a) Arnhold, M. Zur Kenntniss des dreibasischen Ameisensäureäthers und verschiedener Methylale. Justus Liebigs Ann. Chem. 1887, 240, 192208,  DOI: 10.1002/jlac.18872400204
    (b) Löbering, J.; Fleischmann, A. Über Mono- und Dioxymethylen-dimethyläther. Ber. Dtsch. Chem. Ges. B 1937, 70, 16801683,  DOI: 10.1002/cber.19370700807
  26. 26
    Direct separation of methylal from DCM is a technical problem that can be efficiently addressed by special distillation techniques. See, for example:Berg, L. Separation of methylene vhloride from methylal by extractive distillation. US Patent US5051153A. Sept. 24, 1991.
  27. 27
    Giacalone, F.; Gruttadauria, M.; Agrigento, P.; Noto, R. Low-loading asymmetric organocatalysis. Chem. Soc. Rev. 2012, 41, 24062447,  DOI: 10.1039/C1CS15206H
  28. 28
    The TOF figure allows an estimation of the pseudo-first-order rate for the dichloromethane activation in the order k′ ≈ 10–2 s–1 (=45 h–1/3600 s h–1). We have shown before that the reaction of 1a with DCM is very slow at room temperature but completes within 24 h at 60 °C. Thus, taking the latter value as an indicative value for the half-life of 1a in DCM at 40 °C, the pseudo-first-order rate for this reaction should be in the order of 10–5 s–1 (k′ = Ln(2)/t1/2).
  29. 29
    (a) Dehmlow, E. V. Advances in Phase-Transfer Catalysis [New synthetic methods (20)]. Angew. Chem., Int. Ed. Engl. 1977, 16, 493505,  DOI: 10.1002/anie.197704933
    (b) Makosza, M.; Fedoryński, M.; Eley, D. D.; Pines, H.; Weisz, P. B. Catalysis in Two-Phase Systems: Phase Transfer and Related Phenomena. Adv. Catal. 1987, 35, 375422,  DOI: 10.1016/S0360-0564(08)60097-8
  30. 30
    (b) Makosza, M. Phase-transfer catalysis. A general green methodology in organic synthesis. Pure Appl. Chem. 2000, 72, 13991403,  DOI: 10.1351/pac200072071399
    (a) Liotta, C. L.; Berkner, J.; Wright, J.; Fair, B. Mechanisms and Applications of Solid─Liquid Phase-Transfer Catalysis. Phase-Transfer Catalysis 1997, 659, 2940,  DOI: 10.1021/bk-1997-0659.ch003
  31. 31
    Díez-Barra, E.; Hoz, A.; Sánchez-Migallón, A.; Tejeda, J. Phase transfer catalysis without solvent: synthesis of bisazolylalkanes. Heterocycles 1992, 34, 13651373,  DOI: 10.3987/COM-92-6024
  32. 32
    Blümel, M.; Crocker, R. D.; Harper, J. B.; Enders, D.; Nguyen, T. V. N-Heterocyclic olefins as efficient phase-transfer catalysts for base-promoted alkylation reactions. Chem. Commun. 2016, 52, 79587961,  DOI: 10.1039/C6CC03771B
  33. 33
    (a) Wu, H.-S.; Jou, S.-H. Kinetics of formation of diphenoxymethane from phenol and dichloromethane using phase-transfer catalysis. J. Chem. Technol. Biotechnol. 1995, 64, 325330,  DOI: 10.1002/jctb.280640403
    (b) Cornélis, A.; Laszlo, P. Clay-Supported Reagents; II. Quaternary Ammonium-Exchanged Montmorillonite as Catalyst in the Phase-Transfer Preparation of Symmetrical Formaldehyde Acetals. Synthesis 1982, 1982, 162163,  DOI: 10.1055/s-1982-29732
    (c) Wu, H.-S.; Fang, T.-R.; Meng, S.-S.; Hu, K.-H. Equilibrium and extraction of quaternary salt in an organic solvent/alkaline solution: effect of NaOH concentration. J. Mol. Catal. A: Chem. 1998, 136, 135146,  DOI: 10.1016/S1381-1169(98)00054-5
    (d) Liu, W.; Szewczyk, J.; Waykole, L.; Repic, O.; Blacklock, T. J. Practical Synthesis of Diaryloxymethanes. Synth. Commun. 2003, 33, 17511754,  DOI: 10.1081/SCC-120018936
  34. 34
    Liou, C.-C.; Brodbelt, J. S. Determination of orders of relative alkali metal ion affinities of crown ethers and acyclic analogs by the kinetic method. J. Am. Soc. Mass Spectrom. 1992, 3, 543548,  DOI: 10.1016/1044-0305(92)85031-E
  35. 35
    Gobbi, A.; Landini, D.; Maia, A.; Secci, D. Metal Ion Catalysis in Nucleophilic Substitution Reactions Promoted by Complexes of Polyether Ligands with Alkali Metal Salts. J. Org. Chem. 1995, 60, 59545957,  DOI: 10.1021/jo00123a036
  36. 36
    Archer, R. H.; Carpenter, J. R.; Hwang, S.-J.; Burton, A. W.; Chen, C.-Y.; Zones, S. I.; Davis, M. E. Physicochemical Properties and Catalytic Behavior of the Molecular Sieve SSZ-70. Chem. Mater. 2010, 22, 25632572,  DOI: 10.1021/cm9035677
  37. 37
    (a) https://sdbs.db.aist.go.jp/sdbs/cgi-bin/cre_index.cgi.

    For diphenoxymethane, see:

    (b) Salvador, T. K.; Arnett, C. H.; Kundu, S.; Sapiezynski, N. G.; Bertke, J. A.; Raghibi Boroujeni, M.; Warren, T. H. Copper Catalyzed sp3 C–H Etherification with Acyl Protected Phenols. J. Am. Chem. Soc. 2016, 138, 1658016583,  DOI: 10.1021/jacs.6b09057

    For dibenzyloxymethane, see:

    (c) Cacciapaglia, R.; Di Stefano, S.; Mandolini, L. Metathesis Reaction of Formaldehyde Acetals: An Easy Entry into the Dynamic Covalent Chemistry of Cyclophane Formation. J. Am. Chem. Soc. 2005, 127, 1366613671,  DOI: 10.1021/ja054362o

    For diallyloxymethane, see:

    (d) El-Brollosy, N. R.; Jørgensen, P. T.; Dahan, B.; Boel, A. M.; Pedersen, E. B.; Nielsen, C. Synthesis of Novel N-1 (Allyloxymethyl) Analogues of 6-Benzyl-1-(ethoxymethyl)-5-isopropyluracil (MKC-442, Emivirine) with Improved Activity Against HIV-1 and Its Mutants. J. Med. Chem. 2002, 45, 57215726,  DOI: 10.1021/jm020949r

    For diisopropoxymethane, see:

    (e) Berkefeld, A.; Piers, W. E.; Parvez, M.; Castro, L.; Maron, L.; Eisenstein, O. Decamethylscandocinium-hydrido-(perfluorophenyl)borate: fixation and tandem tris(perfluorophenyl)borane catalysed deoxygenative hydrosilation of carbon dioxide. Chem. Sci. 2013, 4, 21522162,  DOI: 10.1039/c3sc50145k

    For di(ethylthio)methane, see:

    (f) Zaidi, J. H.; Naeem, F.; Khan, K. M.; Iqbal, R.; Zia-Ullah Synthesis of Dithioacetals and Oxathioacetals with Chiral Auxiliaries. Synth. Commun. 2004, 34, 26412653,  DOI: 10.1081/SCC-200025627

    For bis(pyrazolyl)methane, see:

    (g) Field, L. D.; Messerle, B. A.; Rehr, M.; Soler, L. P.; Hambley, T. W. Cationic Iridium(I) Complexes as Catalysts for the Alcoholysis of Silanes. Organometallics 2003, 22, 23872395,  DOI: 10.1021/om020938w
  38. 38
    Liang, L. C.; Chien, P. S.; Lee, P. Y.; Lin, J. M.; Huang, Y. L. Terminal nickel(ii) amide, alkoxide, and thiolate complexes containing amido diphosphine ligands of the type [N(o-C6H4PR2)2] (R = Ph, iPr, Cy). Dalton Trans. 2008, 25, 33203327,  DOI: 10.1039/b719894a

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 8 publications.

  1. Joshua T. Kamps, Bruce E. Kirkpatrick, Sean P. Keyser, Connor E. Miksch, Benjamin R. Nelson, John F. Rynk, Benjamin D. Fairbanks, Kristi S. Anseth, Christopher N. Bowman. Linear and Network-Forming Acetal Polymerization of Multifunctional Alcohols with Dichloromethane for Degradable and Recyclable Materials. Macromolecules 2025, 58 (3) , 1578-1584. https://doi.org/10.1021/acs.macromol.4c03078
  2. David Sánchez-Roa, Valentina Sessini, Marta E. G. Mosquera, Juan Cámpora. N-Heterocyclic Carbene-Carbodiimide (NHC-CDI) Betaines as Organocatalysts for β-Butyrolactone Polymerization: Synthesis of Green PHB Plasticizers with Tailored Molecular Weights. ACS Catalysis 2024, 14 (4) , 2487-2501. https://doi.org/10.1021/acscatal.3c05357
  3. Matthew J. Cullen, Matthew G. Davidson, Matthew D. Jones, Jack A. Stewart. Valorization of polyoxymethylene (POM) waste as a C 1 synthon for industrially relevant dialkoxymethanes and cyclic aminals. RSC Sustainability 2025, 3 (3) , 1114-1121. https://doi.org/10.1039/D4SU00547C
  4. Jyoti Sharma, Suman Das, Archana Jain, Tarun K. Panda. Structural Diversity of Cadmium Complexes with N‐Heterocyclic Carbene‐Isothiocyanate Zwitterionic Ligand. Zeitschrift für anorganische und allgemeine Chemie 2024, 650 (24) https://doi.org/10.1002/zaac.202400169
  5. Xinyu Zhang, Qing Luo, Fengying Zhang, Xinye Zhao, Ying Li, Ning Yang, Liangshan Feng. Preparation of PLA Nanoparticles and Study of Their Influencing Factors. Molecules 2024, 29 (23) , 5566. https://doi.org/10.3390/molecules29235566
  6. Le Dung Pham, Red O. Smith‐Sweetser, Briana Krupinsky, Carolyn E. Dewey, Jessica R. Lamb. Switchable Organocatalysis from N ‐heterocyclic Carbene‐Carbodiimide Adducts with Tunable Release Temperature**. Angewandte Chemie 2023, 135 (48) https://doi.org/10.1002/ange.202314376
  7. Le Dung Pham, Red O. Smith‐Sweetser, Briana Krupinsky, Carolyn E. Dewey, Jessica R. Lamb. Switchable Organocatalysis from N ‐heterocyclic Carbene‐Carbodiimide Adducts with Tunable Release Temperature**. Angewandte Chemie International Edition 2023, 62 (48) https://doi.org/10.1002/anie.202314376
  8. Nidhi Jangir, Saneha Sohu, Surendra Kumar Bagaria, Kartik Kumar, Kapil Pareek, Dinesh Kumar Jangid. Optimization of Catalytic Bromination of Acetophenones Using Taguchi's Method. ChemistrySelect 2023, 8 (20) https://doi.org/10.1002/slct.202204870

The Journal of Organic Chemistry

Cite this: J. Org. Chem. 2021, 86, 23, 16725–16735
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.joc.1c01971
Published November 1, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

2389

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Synthesis of 1a and 2a

    Scheme 2

    Scheme 2. Synthesis of Methylal Using 2a as a Catalyst

    Figure 1

    Figure 1. Yield vs time plot for the reaction of DCM with NaOMe catalyzed by 2a. Conditions: 40 °C, NaOMe, 2 g (31 mmol); DCM (neat) 40 mL (626 mmol); 2a, 37 mg (0.2 mol % with regard to NaOMe). The last check after 24 h was consistent with full NaOMe consumption.

    Scheme 3

    Scheme 3. Catalytic Transformation of DCM into CH2Z2 Using Combinations of Weak Protic Acids and Suitable Bases (NaOH or NaH)

    Scheme 4

    Scheme 4. Synthesis of the Catalyst Precursor 3a

    Figure 2

    Figure 2. Solid–liquid phase-transfer catalysis.

    Scheme 5

    Scheme 5. Transformation of 2a and 1a into 2b and 1b, Respectively, in Catalytic Media
  • References


    This article references 38 other publications.

    1. 1
      (a) Rossberg, M.; Lendle, W.; Pfleiderer, G.; Tögel, A.; Torkelson, T. R.; Beutel, K. K. Chloromethanes. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley, 2011. DOI:  DOI: 10.1002/14356007.a06_233.pub3 .
      (b) Tyner, T.; Francis, J. Dichloromethane. In ACS Reagent Chemicals; American Chemical Society, 2017. DOI:  DOI: 10.1021/acsreagents.4117 .
    2. 2
      (a) Lee, H. M.; Lu, C. Y.; Chen, C. Y.; Chen, W. L.; Lin, H. C.; Chiu, P. L.; Cheng, P. Y. Palladium complexes with ethylene-bridged bis(N-heterocyclic carbene) for C–C coupling reactions. Tetrahedron 2004, 60, 58075825,  DOI: 10.1016/j.tet.2004.04.070
      (b) Noujeim, N.; Leclercq, L. C.; Schmitzer, A. R. N,Nʼ-Disubstituted Methylenediimidazolium Salts: A Versatile Guest for Various Macrocycles. J. Org. Chem. 2008, 73, 37843790,  DOI: 10.1021/jo702683c
      (c) Cheong, M.; Kim, M.-N.; Shim, J. Y. Rhodium-catalyzed double carbonylation of diiodomethane in the presence of triethylorthoformate. J. Organomet. Chem. 1996, 520, 253255,  DOI: 10.1016/0022-328X(96)06349-8
      (d) Zadykowicz, J.; Potvin, P. G. N-(2-tetrahydrofuranyl)azole nucleoside analogs by reactions of azoles with dihalomethanes in tetrahydrofuran. J. Heterocycl. Chem. 1999, 36, 623626,  DOI: 10.1002/jhet.5570360308
      (e) Zidan, A.; Garrec, J.; Cordier, M.; El-Naggar, A. M.; Abd El-Sattar, N. E. A.; Ali, A. K.; Hassan, M. A.; El Kaim, L. β-Lactam Synthesis through Diodomethane Addition to Amide Dianions. Angew. Chem., Int. Ed. 2017, 56, 1217912183,  DOI: 10.1002/anie.201706315
    3. 3
      (a) Durandetti, S.; Sibille, S.; Perichon, J. Electrochemical cyclopropanation of alkenes using dibromomethane and zinc in dichloromethane/DMF mixture. J. Org. Chem. 1991, 56, 32553258,  DOI: 10.1021/jo00010a015
      (b) Hon, Y.-S.; Hsieh, C.-H.; Liu, Y.-W. Dibromomethane as one-carbon source in organic synthesis: total synthesis of (±)- and (−)-methylenolactocin. Tetrahedron 2005, 61, 27132723,  DOI: 10.1016/j.tet.2005.01.057
      (c) Brunner, G.; Eberhard, L.; Oetiker, J.; Schröder, F. Cyclopropanation with Dibromomethane under Grignard and Barbier Conditions. Synthesis 2009, 21, 37083718,  DOI: 10.1055/s-0029-1216999
      (d) Raposo, C. D. C. B. G. Diiodomethane: A Versatile C1 Building Block. Synlett 2013, 24, 17371738,  DOI: 10.1055/s-0033-1338964
    4. 4

      Though rather inert, DCM can undergo different reactions when used as a solvent, e.g., through free radical routes; see:

      (a) Levina, I. I.; Klimovich, O. N.; Vinogradov, D. S.; Podrugina, T. A.; Bormotov, D. S.; Kononikhin, A. S.; Dement’eva, O. V.; Senchikhin, I. N.; Nikolaev, E. N.; Kuzmin, V. A.; Nekipelova, T. D. Dichloromethane as solvent and reagent: a case study of photoinduced reactions in mixed phosphonium-iodonium ylide. J. Phys. Org. Chem. 2018, 31, e3844,  DOI: 10.1002/poc.3844

      and references cited therein

      In addition, DCM is known to react violently with some strong bases like potassium t-butoxide or some reactive metals (e.g., Li or Na/K alloy); see:

      (b) Lewis, R. J., Sr., Ed.; Sax’s Dangerous Properties of Industrial Materials, 11th ed.; Wiley-Interscience, Wiley & Sons, Inc: Hoboken, NJ, 2004; p 2436.
      (c) Anhydrous DCM is not corrosive for most metals, but in the presence of water, it slowly hydrolyzes releasing small amounts of HCl, which can attack stainless steel, aluminum, or even copper pipes. See any dichlorometane Safety Data Sheet (MSDS, e.g., Fisher, ACC no. 14930).

      Moreover, there are reports concerning dichloromethane C–Cl bond substitutions but are usually very slow processes, see:

      (d) Rudine, A. B.; Walter, M. G.; Wamser, C. C. Reaction of Dichloromethane with Pyridine Derivatives under Ambient Conditions. J. Org. Chem. 2010, 75, 42924295,  DOI: 10.1021/jo100276m
    5. 5
      (a) Closs, G. L. Carbenes from Alkyl Halides and Organolithium Compounds. IV. Formation of Alkylcarbenes from Methylene Chloride and Alkyllithium Compounds. J. Am. Chem. Soc. 1962, 84, 809813,  DOI: 10.1021/ja00864a026
      (b) Böhm, A.; Bach, T. Radical Reactions Induced by Visible Light in Dichloromethane Solutions of Hünig’s Base: Synthetic Applications and Mechanistic Observations. Chem. - Eur. J. 2016, 22, 1592115928,  DOI: 10.1002/chem.201603303
      (c) Sturala, J.; Ambrosi, A.; Sofer, Z.; Pumera, M. Covalent Functionalization of Exfoliated Arsenic with Chlorocarbene. Angew. Chem., Int. Ed. 2018, 57, 1483714840,  DOI: 10.1002/anie.201809341
    6. 6
      Young, J. A. Dichloromethane. J. Chem. Educ. 2004, 81, 1415,  DOI: 10.1021/ed081p1415
    7. 7
      (a) Salthammer, T.; Mentese, S.; Marutzky, R. Formaldehyde in the Indoor Environment. Chem. Rev. 2010, 110, 25362572,  DOI: 10.1021/cr800399g
      (b) Li, W.; Wu, X.-F. The Applications of (Para)formaldehyde in Metal-Catalyzed Organic Synthesis. Adv. Synth. Catal. 2015, 357, 33933418,  DOI: 10.1002/adsc.201500753
      (c) Pathak, D. D.; Gerald, J. J. An Efficient and Convenient Method for the Synthesis of Dialkoxymethanes Using Kaolinite as a Catalyst. Synth. Commun. 2003, 33, 15571561,  DOI: 10.1081/SCC-120018774
    8. 8
      Hossaini, R.; Chipperfield, M. P.; Montzka, S. A.; Leeson, A. A.; Dhomse, S. S.; Pyle, J. A. The increasing threat to stratospheric ozone from dichloromethane. Nat. Commun. 2017, 8, 15962,  DOI: 10.1038/ncomms15962
    9. 9
      (a) Kane, A.; Giraudet, S.; Vilmain, J.-B.; Le Cloirec, P. Intensification of the temperature-swing adsorption process with a heat pump for the recovery of dichloromethane. J. Environ. Chem. Eng. 2015, 3, 734743,  DOI: 10.1016/j.jece.2015.02.021
      (b) Wu, X.-F. W.; Tlili, A.; Schranck, J.; Wu, X.-F. W. The Application of Dichloromethane and Chloroform as Reagents in Organic Synthesis. Solvents as Reagents in Organic Synthesis 2017, 125159,  DOI: 10.1002/9783527805624.ch4
    10. 10
      (a) Lautenschütz, L.; Oestreich, D.; Seidenspinner, P.; Arnold, U.; Dinjus, E.; Sauer, J. Physico-chemical properties and fuel characteristics of oxymethylene dialkyl ethers. Fuel 2016, 173, 129137,  DOI: 10.1016/j.fuel.2016.01.060
      (b) Thenert, K.; Beydoun, K.; Wiesenthal, J.; Leitner, W.; Klankermayer, J. Ruthenium-Catalyzed Synthesis of Dialkoxymethane Ethers Utilizing Carbon Dioxide and Molecular Hydrogen. Angew. Chem., Int. Ed. 2016, 55, 1226612269,  DOI: 10.1002/anie.201606427
      (c) Schieweck, B. G.; Klankermayer, J. Tailor-made Molecular Cobalt Catalyst System for the Selective Transformation of Carbon Dioxide to Dialkoxymethane Ethers. Angew. Chem., Int. Ed. 2017, 56, 1085410857,  DOI: 10.1002/anie.201702905
    11. 11
      (a) El-Brollosy, N. R.; Jørgensen, P. T.; Dahan, B.; Boel, A. M.; Pedersen, E. B.; Nielsen, C. Synthesis of Novel N-1 (Allyloxymethyl) Analogues of 6-Benzyl-1-(ethoxymethyl)-5-isopropyluracil (MKC-442, Emivirine) with Improved Activity Against HIV-1 and Its Mutants. J. Med. Chem. 2002, 45, 57215726,  DOI: 10.1021/jm020949r
      (b) Wamberg, M.; Pedersen, E. B.; El-Brollosy, N. R.; Nielsen, C. Synthesis of 6-arylvinyl analogues of the HIV drugs SJ-3366 and Emivirine. Bioorg. Med. Chem. 2004, 12, 11411149,  DOI: 10.1016/j.bmc.2003.11.032
      (c) Loksha, Y. M.; Pedersen, E. B.; Loddo, R.; Sanna, G.; Collu, G.; Giliberti, G.; Colla, P. L. Synthesis of Novel Fluoro Analogues of MKC442 as Microbicides. J. Med. Chem. 2014, 57, 51695178,  DOI: 10.1021/jm500139a
    12. 12
      (a) Makosza, M.; Sypniewski, M. Reaction of sulfonium salts of formaldehyde dithioacetals with aromatic aldehydes and rearrangements of the produced thioalkyl oxiranes. Tetrahedron 1995, 51, 1059310600,  DOI: 10.1016/0040-4020(95)00632-I
      (b) Luh, T.-Y. Recent advances on the synthetic applications of the dithioacetal functionality. J. Organomet. Chem. 2002, 653, 209214,  DOI: 10.1016/S0022-328X(02)01151-8
    13. 13
      (a) Otero, A.; Fernández-Baeza, J.; Antiñolo, A.; Tejeda, J.; Lara-Sánchez, A. Heteroscorpionate ligands based on bis(pyrazol-1-yl)methane: design and coordination chemistry. Dalton Trans. 2004, 10, 14991510,  DOI: 10.1039/B401425A
      (b) Krieck, S.; Koch, A.; Hinze, K.; Müller, C.; Lange, J.; Görls, H.; Westerhausen, M. s-Block Metal Complexes with Bis- and Tris(pyrazolyl)methane and -methanide Ligands. Eur. J. Inorg. Chem. 2016, 15, 23322348,  DOI: 10.1002/ejic.201501263
    14. 14
      (a) Csok, Z.; Vechorkin, O.; Harkins, S. B.; Scopelliti, R.; Hu, X. Nickel Complexes of a Pincer NN2 Ligand: Multiple Carbon-Chloride Activation of CH2Cl2 and CHCl3 Leads to Selective Carbon-Carbon Bond Formation. J. Am. Chem. Soc. 2008, 130, 81568157,  DOI: 10.1021/ja8025938
      (b) Zhan, L.; Pan, R.; Xing, P.; Jiang, B. An efficient method for the preparation of dialkoxymethanes from dichloromethane with alcohols catalyzed by a Cu-NHC complex. Tetrahedron Lett. 2016, 57, 40364038,  DOI: 10.1016/j.tetlet.2016.07.056
    15. 15
      Lamb, J. R.; Brown, C. M.; Johnson, J. A. N-Heterocyclic carbene–carbodiimide (NHC–CDI) betaine adducts: synthesis, characterization, properties, and applications. Chem. Sci. 2021, 12, 26992715,  DOI: 10.1039/D0SC06465C
    16. 16
      Baishya, A.; Kumar, L.; Barman, M. K.; Peddarao, T.; Nembenna, S. Air Stable N-Heterocyclic Carbene-Carbodiimide (“NHC-CDI”) Adducts: Zwitterionic Type Bulky Amidinates. ChemistrySelect 2016, 1, 498503,  DOI: 10.1002/slct.201600019
    17. 17
      (a) Márquez, A.; Ávila, E.; Urbaneja, C.; Álvarez, E.; Palma, P.; Cámpora, J. Copper(I) Complexes of Zwitterionic Imidazolium-2-Amidinates, a Promising Class of Electroneutral, Amidinate-Type Ligands. Inorg. Chem. 2015, 54, 1100711017,  DOI: 10.1021/acs.inorgchem.5b02141
      (b) Baishya, A.; Kumar, L.; Barman, M. K.; Biswal, H. S.; Nembenna, S. N-Heterocyclic Carbene–Carbodiimide (“NHC–CDI”) Adduct or Zwitterionic-Type Neutral Amidinate-Supported Magnesium(II) and Zinc(II) Complexes. Inorg. Chem. 2017, 56, 95359546,  DOI: 10.1021/acs.inorgchem.7b00879
    18. 18
      (a) Martínez-Prieto, L. M.; Cano, I.; Márquez, A.; Baquero, E. A.; Tricard, S.; Cusinato, L.; del Rosal, I.; Poteau, R.; Coppel, Y.; Philippot, K.; Chaudret, B.; Cámpora, J.; van Leeuwen, P. W. N. M. Zwitterionic amidinates as effective ligands for platinum nanoparticle hydrogenation catalysts. Chem. Sci. 2017, 8, 29312941,  DOI: 10.1039/C6SC05551F
      (b) López-Vinasco, A. M.; Martínez-Prieto, L. M.; Asensio, J. M.; Lecante, P.; Chaudret, B.; Cámpora, J.; van Leeuwen, P. W. N. M. Novel nickel nanoparticles stabilized by imidazolium-amidinate ligands for selective hydrogenation of alkynes. Catal. Sci. Technol. 2020, 10, 342350,  DOI: 10.1039/C9CY02172H
    19. 19
      Gallagher, N. M.; Zhukhovitskiy, A. V.; Nguyen, H. V. T.; Johnson, J. A. Main-Chain Zwitterionic Supramolecular Polymers Derived from N-Heterocyclic Carbene–Carbodiimide (NHC–CDI) Adducts. Macromolecules 2018, 51, 30063016,  DOI: 10.1021/acs.macromol.8b00579
    20. 20
      Gallagher, N. M.; Ye, H.-Z.; Feng, S.; Lopez, J.; Zhu, Y. G.; Van Voorhis, T.; Shao-Horn, Y.; Johnson, J. A. An N-Heterocyclic-Carbene-Derived Distonic Radical Cation. Angew. Chem., Int. Ed. 2020, 59, 39523955,  DOI: 10.1002/anie.201915534
    21. 21
      Sánchez-Roa, D.; Santiago, T. G.; Fernández-Millán, M.; Cuenca, T.; Palma, P.; Cámpora, J.; Mosquera, M. E. G. Interaction of an imidazolium-2-amidinate (NHC-CDI) zwitterion with zinc dichloride in dichloromethane: role as ligands and C–Cl activation promoters. Chem. Commun. 2018, 54, 1258612589,  DOI: 10.1039/C8CC07661H
    22. 22
      (a) Rivlin, M.; Eliav, U.; Navon, G. NMR Studies of the Equilibria and Reaction Rates in Aqueous Solutions of Formaldehyde. J. Phys. Chem. B 2015, 119, 44794487,  DOI: 10.1021/jp513020y
      (b) Dankelman, W.; Daemen, J. M. H. Gas Chromatographic and Nuclear Magnetic Resonance Determination of Linear Formaldehyde Oligomers in Formaline. Anal. Chem. 1976, 48, 401404,  DOI: 10.1021/ac60366a030
    23. 23
      (a) Li, M.; Long, Y.; Deng, Z.; Zhang, H.; Yang, X.; Wang, G. Ruthenium trichloride as a new catalyst for selective production of dimethoxymethane from liquid methanol with molecular oxygen as sole oxidant. Catal. Commun. 2015, 68, 4648,  DOI: 10.1016/j.catcom.2015.04.031
      (b) Thavornprasert, K.-a.; Capron, M.; Jalowiecki-Duhamel, L.; Dumeignil, F. One-pot 1,1-dimethoxymethane synthesis from methanol: a promising pathway over bifunctional catalysts. Catal. Sci. Technol. 2016, 6, 958970,  DOI: 10.1039/C5CY01858G
      (c) Chao, Y.; Lai, J.; Yang, Y.; Zhou, P.; Zhang, Y.; Mu, Z.; Li, S.; Zheng, J.; Zhu, Z.; Tan, Y. Visible light-driven methanol dehydrogenation and conversion into 1,1-dimethoxymethane over a non-noble metal photocatalyst under acidic conditions. Catal. Sci. Technol. 2018, 8, 33723378,  DOI: 10.1039/C8CY01030G
    24. 24
      Łojewska, J.; Wasilewski, J.; Terelak, K.; Łojewski, T.; Kołodziej, A. Selective oxidation of methylal as a new catalytic route to concentrated formaldehyde: Reaction kinetic profile in gradientless flow reactor. Catal. Commun. 2008, 9, 18331837,  DOI: 10.1016/j.catcom.2008.02.014
    25. 25
      (a) Arnhold, M. Zur Kenntniss des dreibasischen Ameisensäureäthers und verschiedener Methylale. Justus Liebigs Ann. Chem. 1887, 240, 192208,  DOI: 10.1002/jlac.18872400204
      (b) Löbering, J.; Fleischmann, A. Über Mono- und Dioxymethylen-dimethyläther. Ber. Dtsch. Chem. Ges. B 1937, 70, 16801683,  DOI: 10.1002/cber.19370700807
    26. 26
      Direct separation of methylal from DCM is a technical problem that can be efficiently addressed by special distillation techniques. See, for example:Berg, L. Separation of methylene vhloride from methylal by extractive distillation. US Patent US5051153A. Sept. 24, 1991.
    27. 27
      Giacalone, F.; Gruttadauria, M.; Agrigento, P.; Noto, R. Low-loading asymmetric organocatalysis. Chem. Soc. Rev. 2012, 41, 24062447,  DOI: 10.1039/C1CS15206H
    28. 28
      The TOF figure allows an estimation of the pseudo-first-order rate for the dichloromethane activation in the order k′ ≈ 10–2 s–1 (=45 h–1/3600 s h–1). We have shown before that the reaction of 1a with DCM is very slow at room temperature but completes within 24 h at 60 °C. Thus, taking the latter value as an indicative value for the half-life of 1a in DCM at 40 °C, the pseudo-first-order rate for this reaction should be in the order of 10–5 s–1 (k′ = Ln(2)/t1/2).
    29. 29
      (a) Dehmlow, E. V. Advances in Phase-Transfer Catalysis [New synthetic methods (20)]. Angew. Chem., Int. Ed. Engl. 1977, 16, 493505,  DOI: 10.1002/anie.197704933
      (b) Makosza, M.; Fedoryński, M.; Eley, D. D.; Pines, H.; Weisz, P. B. Catalysis in Two-Phase Systems: Phase Transfer and Related Phenomena. Adv. Catal. 1987, 35, 375422,  DOI: 10.1016/S0360-0564(08)60097-8
    30. 30
      (b) Makosza, M. Phase-transfer catalysis. A general green methodology in organic synthesis. Pure Appl. Chem. 2000, 72, 13991403,  DOI: 10.1351/pac200072071399
      (a) Liotta, C. L.; Berkner, J.; Wright, J.; Fair, B. Mechanisms and Applications of Solid─Liquid Phase-Transfer Catalysis. Phase-Transfer Catalysis 1997, 659, 2940,  DOI: 10.1021/bk-1997-0659.ch003
    31. 31
      Díez-Barra, E.; Hoz, A.; Sánchez-Migallón, A.; Tejeda, J. Phase transfer catalysis without solvent: synthesis of bisazolylalkanes. Heterocycles 1992, 34, 13651373,  DOI: 10.3987/COM-92-6024
    32. 32
      Blümel, M.; Crocker, R. D.; Harper, J. B.; Enders, D.; Nguyen, T. V. N-Heterocyclic olefins as efficient phase-transfer catalysts for base-promoted alkylation reactions. Chem. Commun. 2016, 52, 79587961,  DOI: 10.1039/C6CC03771B
    33. 33
      (a) Wu, H.-S.; Jou, S.-H. Kinetics of formation of diphenoxymethane from phenol and dichloromethane using phase-transfer catalysis. J. Chem. Technol. Biotechnol. 1995, 64, 325330,  DOI: 10.1002/jctb.280640403
      (b) Cornélis, A.; Laszlo, P. Clay-Supported Reagents; II. Quaternary Ammonium-Exchanged Montmorillonite as Catalyst in the Phase-Transfer Preparation of Symmetrical Formaldehyde Acetals. Synthesis 1982, 1982, 162163,  DOI: 10.1055/s-1982-29732
      (c) Wu, H.-S.; Fang, T.-R.; Meng, S.-S.; Hu, K.-H. Equilibrium and extraction of quaternary salt in an organic solvent/alkaline solution: effect of NaOH concentration. J. Mol. Catal. A: Chem. 1998, 136, 135146,  DOI: 10.1016/S1381-1169(98)00054-5
      (d) Liu, W.; Szewczyk, J.; Waykole, L.; Repic, O.; Blacklock, T. J. Practical Synthesis of Diaryloxymethanes. Synth. Commun. 2003, 33, 17511754,  DOI: 10.1081/SCC-120018936
    34. 34
      Liou, C.-C.; Brodbelt, J. S. Determination of orders of relative alkali metal ion affinities of crown ethers and acyclic analogs by the kinetic method. J. Am. Soc. Mass Spectrom. 1992, 3, 543548,  DOI: 10.1016/1044-0305(92)85031-E
    35. 35
      Gobbi, A.; Landini, D.; Maia, A.; Secci, D. Metal Ion Catalysis in Nucleophilic Substitution Reactions Promoted by Complexes of Polyether Ligands with Alkali Metal Salts. J. Org. Chem. 1995, 60, 59545957,  DOI: 10.1021/jo00123a036
    36. 36
      Archer, R. H.; Carpenter, J. R.; Hwang, S.-J.; Burton, A. W.; Chen, C.-Y.; Zones, S. I.; Davis, M. E. Physicochemical Properties and Catalytic Behavior of the Molecular Sieve SSZ-70. Chem. Mater. 2010, 22, 25632572,  DOI: 10.1021/cm9035677
    37. 37
      (a) https://sdbs.db.aist.go.jp/sdbs/cgi-bin/cre_index.cgi.

      For diphenoxymethane, see:

      (b) Salvador, T. K.; Arnett, C. H.; Kundu, S.; Sapiezynski, N. G.; Bertke, J. A.; Raghibi Boroujeni, M.; Warren, T. H. Copper Catalyzed sp3 C–H Etherification with Acyl Protected Phenols. J. Am. Chem. Soc. 2016, 138, 1658016583,  DOI: 10.1021/jacs.6b09057

      For dibenzyloxymethane, see:

      (c) Cacciapaglia, R.; Di Stefano, S.; Mandolini, L. Metathesis Reaction of Formaldehyde Acetals: An Easy Entry into the Dynamic Covalent Chemistry of Cyclophane Formation. J. Am. Chem. Soc. 2005, 127, 1366613671,  DOI: 10.1021/ja054362o

      For diallyloxymethane, see:

      (d) El-Brollosy, N. R.; Jørgensen, P. T.; Dahan, B.; Boel, A. M.; Pedersen, E. B.; Nielsen, C. Synthesis of Novel N-1 (Allyloxymethyl) Analogues of 6-Benzyl-1-(ethoxymethyl)-5-isopropyluracil (MKC-442, Emivirine) with Improved Activity Against HIV-1 and Its Mutants. J. Med. Chem. 2002, 45, 57215726,  DOI: 10.1021/jm020949r

      For diisopropoxymethane, see:

      (e) Berkefeld, A.; Piers, W. E.; Parvez, M.; Castro, L.; Maron, L.; Eisenstein, O. Decamethylscandocinium-hydrido-(perfluorophenyl)borate: fixation and tandem tris(perfluorophenyl)borane catalysed deoxygenative hydrosilation of carbon dioxide. Chem. Sci. 2013, 4, 21522162,  DOI: 10.1039/c3sc50145k

      For di(ethylthio)methane, see:

      (f) Zaidi, J. H.; Naeem, F.; Khan, K. M.; Iqbal, R.; Zia-Ullah Synthesis of Dithioacetals and Oxathioacetals with Chiral Auxiliaries. Synth. Commun. 2004, 34, 26412653,  DOI: 10.1081/SCC-200025627

      For bis(pyrazolyl)methane, see:

      (g) Field, L. D.; Messerle, B. A.; Rehr, M.; Soler, L. P.; Hambley, T. W. Cationic Iridium(I) Complexes as Catalysts for the Alcoholysis of Silanes. Organometallics 2003, 22, 23872395,  DOI: 10.1021/om020938w
    38. 38
      Liang, L. C.; Chien, P. S.; Lee, P. Y.; Lin, J. M.; Huang, Y. L. Terminal nickel(ii) amide, alkoxide, and thiolate complexes containing amido diphosphine ligands of the type [N(o-C6H4PR2)2] (R = Ph, iPr, Cy). Dalton Trans. 2008, 25, 33203327,  DOI: 10.1039/b719894a
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.1c01971.

    • Additional experimental procedures and spectroscopic data (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.