Next Article in Journal
Electrochemical Activity of Original and Infiltrated Fe-Doped Ba(Ce,Zr,Y)O3-Based Electrodes to Be Used for Protonic Ceramic Fuel Cells
Previous Article in Journal
ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tetrachlorocobaltate-Catalyzed Methane Oxidation to Methyl Trifluoroacetate

1
Clean Energy Research Center, Korea Institute of Science and Technology (KIST), Seoul 02792, Korea
2
Division of Energy & Environment Technology, KIST School, Korea University of Science and Technology, Seoul 02792, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1419; https://doi.org/10.3390/catal12111419
Submission received: 17 October 2022 / Revised: 4 November 2022 / Accepted: 7 November 2022 / Published: 11 November 2022
(This article belongs to the Section Catalysis for Sustainable Energy)

Abstract

:
In ongoing attempts to efficiently utilize abundant natural gas, there has been steady scientific and industrial interest in using an environmentally benign and inexpensive oxidant (dioxygen O2) for the direct catalytic oxidation of methane to oxygenate products under mild conditions. Here, we report the homogeneous bis(tetramethylammonium) tetrachlorocobaltate ([Me4N]2CoCl4)-catalyzed methane oxidation to methyl trifluoroacetate (MeTFA) with dioxygen O2 in trifluoroacetic acid (HTFA) media. [Me4N]2CoCl4 had the highest catalytic activity among previously reported homogeneous cobalt-based catalyst systems; the turnover of methane to MeTFA reached 8.26 molester molmetal−1h−1 at 180 °C. Results suggest that the ionic form of the catalyst makes the Co species more soluble in the HTFA media; consequently, an active catalyst form, [CoTFAxCly]2−, can form very rapidly. Furthermore, chloride anions dissociated from CoCl42− appear to suppress oxidation of the solvent HTFA, thereby driving the reaction toward methane oxidation. The effects of reaction time, catalyst concentration, O2 and methane pressure, and reaction temperature on MeTFA production were also investigated.

Graphical Abstract

1. Introduction

Methane is the primary constituent of natural gas, a vital part of the world’s energy supply. Methanol demand and production have also been steadily increasing over the last two decades for use as a platform chemical and fuel [1,2]. Developing new methods to convert an affordable starting material (methane) into a strategic product (methanol) is an important process in the chemical industry. While several commercial technologies are currently used to convert methane to syngas and for the subsequent transformation of syngas to value-added chemicals such as methanol [3,4,5], this is an indirect and energy-intensive pathway. Finding a way to directly convert methane to oxygenates under benign conditions is a challenge that has captivated both scientific and industrial communities. However, high-value product formation and selectivity via direct oxidation are limited because of the stability of the C-H bond [6], and this can result in the over-oxidation of methanol to carbon dioxide [7]. To avoid this problem, in recent decades, various methods have been proposed and studied to protect the methanol in the form of an ester, such as methyl bisulfate [8,9,10,11,12] or methyl trifluoroacetate (MeTFA) [13,14,15,16,17,18,19,20,21,22]. Trifluoroacetic acid (HTFA)-based MeTFA synthesis from methane has received attention because it occurs in non-super acidic media under milder conditions; and because of its low boiling temperature (43 °C), methyl trifluoroacetate can be easily separated from the mixture by simple distillation after oxidation.
KIO3, KIO4, K2S2O8, and O2 are frequently used as oxidants to directly synthesize MeTFA from methane in HTFA media. Among them, the most promising oxidant candidate is dioxygen O2 since it is inexpensive, plentiful, non-harmful, and simple to eliminate after the reaction. Accordingly, in recent decades, many efforts have been dedicated to utilizing dioxygen O2 for methane oxidation in HTFA [16,23,24,25,26,27,28,29,30].
For example, the one-pot aerobic oxidation of methane to MeTFA using a catalyst system consisting of three redox couples, Pd2+/Pd0, Q/H2Q (p-benzoquinone/p-hydroquinone), and NO2/NO, was reported by Z. An et al. [25]. The turnover frequency with this catalyst system was only 0.7 per hour due to the limitation of the initial activation step. To improve its catalytic performance, J. Yuan et al. suggested adding perfluorooctane (C8F18) into the Pd(OAc)2/benzoquinone/H5PMo10V2O40 system [26]. The resulting increase in the oxygen solubility of the HTFA/C8F18 media reduced the tension of the liquid/gas interface and resulted in a dramatic increase in MeTFA formation; the turnover number (TON) reached over 100 at 80 °C after 8 h.
Compared to the Pd-based multi-redox couple oxidation system, Co-based catalysts are known to act as a direct oxidation catalyst for hydrocarbon in HTFA. R. Tang et al. reported that Co(III) reacted instantaneously as an electron-transfer oxidant for the activation of the C-H bond in cyclohexane, and O2 was an essential contributor to the regeneration cycle between the Co(II) and Co(III) species [31]. M.N. Vargaftik et al. tested various metals, including Pd, Mn, Fe, Co, Cu, and Pb, on aerobic methane oxidation and showed Co(TFA)2 gave the highest activity for MeTFA formation with a TON of 4 at 180 °C after 4 h [32]. Nevertheless, the precipitation of Co species into inactive CoF2 during the reaction was reported to decrease the rate of MeTFA production.
T. Strassner and co-workers investigated the activities of various cobalt salts, such as Co(OAc)2, CoCl2, Co(NO3)2, Co2O3, CoBr2, Co(acac)3, and CoSO4, on aerobic methane oxidation and found the catalytic activities were highly dependent on the anions of cobalt salts [24]. Among the tested catalysts, Co(NO3)2 and Co(OAc)2 were the most potent for this reaction; TONs of 8.4 and 7.3 were obtained at 160 °C after 24 h, respectively. Furthermore, catalyst deactivation to CoF2 could be prevented by the addition of trifluoroacetic anhydride (TFAA). They also noted that CoBr2, CoF2, and CoSO4 did not show any activity in the reaction due to their poor solubilities in HTFA.
Many inorganic metal compounds have been disregarded as potential homogeneous catalysts because they are insoluble in typical organic solvents, aqueous solutions, or concentrated acids. One way of improving their solubilities is by transforming inorganic metal compounds into organometallic or ionic complexes. For example, in our previous research, we found that transforming neutral PdCl2 to anionic PdCl42− led to greatly enhanced catalytic performance (from 7.3% to 33%) as methane oxidation proceeded in a K2S2O8 oxidant and HTFA solvent system [21]. CoCl2 is the most basic form of Co catalyst, and its performance on aerobic methane oxidation was reported to be TON 2–4.7 after a reaction time of 3–72 h at 180 °C, which corresponds to 0.6 to 0.06 molester molmetal−1h−1 [24].
Here, we introduce an anionic Co complex, bis(tetramethylammonium) tetrachlorocobaltate ([Me4N]2CoCl4), as a methane oxidation catalyst in an O2 oxidant system. Like PdCl42− mentioned above, by changing CoCl2 to an anionic form, CoCl42−, the catalytic activity can be greatly increased to 8.26 molester molmetal−1h−1, which corresponds to the highest value among the homogeneous cobalt catalysts reported. To understand the structure of the real catalytic species, ICP, XPS, and FT-IR analyses were conducted, and the results showed that chloride in CoCl42− reversibly exchanged with trifluoroacetate to form CoClxTFAy2− in the HTFA solvent. The effect of reaction time, catalyst concentration, and dioxygen amount were also investigated.

2. Results and Discussion

[Me4N]2CoCl4 can be made from CoCl2 and Me4NCl by dissolving them in ethanol, as described in the Experimental section. Elemental Analysis of the CHN contents and ICP analysis of the Co content revealed the [Me4N]2CoCl4 was successfully synthesized. Figure 1 provides an X-ray diffractogram of the prepared catalyst which matches the reference [Me4N]2CoCl4 well [33].
Using the synthesized metal catalyst and other conventional Co catalysts, a methane oxidation reaction was conducted in the presence of O2 as an oxidant, trifluoroacetic acid (HTFA) as a solvent, and trifluoroacetic anhydride (TFAA) as a dehydrating agent at 180 °C for 1 h (Scheme 1).
It can be expected that Co(II) is oxidized to Co(III) by dioxygen with the liberation of water, which will be removed quickly by TFAA. After that, the oxidation of methane by Co(III) will take place via radical mechanism in the HTFA media to produce MeTFA, as previously reported [24,32,34].
Table 1 shows that, among the various cobalt catalysts tested, the synthesized [Me4N]2CoCl4 gave the highest activity; 1.00 mmol of [Me4N]2CoCl4 produced 8.26 mmol of MeTFA (Table 1, Entry 1) during a 1 h reaction at 180 °C. Table S1 in Supplementary Materials reveals that, among the O2-based oxidant system, a productivity (molester molmetal−1h−1) of 8.26 for MeTFA formation was the highest value among the homogeneous aerobic catalysts reported (Table S1 in Supplementary Materials). The same reaction conducted in the absence of methane (Entry 2) gave no MeTFA, suggesting the tetramethylammonium cation in [Me4N]2CoCl4 does not lead to the formation of MeTFA. Similarly, using MeTFA as a starting material (Entry 3) showed that MeTFA is quite stable under this reaction condition. CoCl2, the precursor of [Me4N]2CoCl4, only produced a negligible amount of MeTFA (0.33 mmol). Interestingly, the in situ addition of Me4NCl to CoCl2 prior to the oxidation reaction appeared to be effective, increasing the MeTFA formation to some extent; the addition of 2 equiv. of Me4NCl resulted in an increase in MeTFA formation to 5.52 mmol (Entry 5).
On the other hand, Co(NO3)2 and Co(TFA)2 resulted in MeTFA formations of 4.38 mmol and 4.74 mmol, respectively. The other cobalt salts, such as CoBr2 and CoSO4, exhibited very low catalytic performance, with only 0.12 and 0.10 mmol of MeTFA, respectively.

2.1. [Me4N]2CoCl4-Catalyzed Methane Oxidation

As T. Strassner et al. mentioned, the solubility of inorganic cobalt compounds such as CoCl2, CoBr2, and CoSO4 are very poor in HTFA [24]. However, as expected, [Me4N]2CoCl4 was instantly dissolved in HTFA, even at room temperature, forming a pink-colored solution (Figure 2b). Furthermore, this pink-colored solution was also observed during the reaction (Figure 2c); however, only a sky-blue heterogeneous solid remained after the reaction (Figure 2d). When this pink solution and sky-blue solid were used as a catalyst for this methane oxidation, respectively, the quantitation pink solution showed some MeTFA formation, but no product was observed after the sky-blue solid-catalyzed reaction, indicating active catalysts remained in the pink solution phase.
To understand the catalyst form in HTFA in the [Me4N]2CoCl4-catalyzed reaction, the HTFA solvent was removed by vacuum from the pink solution after the reaction, and a pink powder was obtained. Characterization of the obtained pink solid using XPS, IR, and elemental analysis revealed the isolated solid was [Me4N]2CoTFA4, suggesting CoCl42− was turned into CoTFA42− in HTFA.
The XPS data shown in Figure 3 reveals that Co, Cl, C, and N atoms exist in the fresh [Me4N]2CoCl4, as expected, while the Cl peak disappeared with the appearance of an F peak in the solid isolated from the pink solution produced by the [Me4N]2CoCl4-catalyzed reaction.
The detailed C1s and Co2p XPS analyses exhibited in Figure 4 suggest the transformation of CoCl42− to CoTFA42− in HTFA during the reaction. According to the C1s XPS analysis, apart from a peak at 286.5 eV, which was ascribed to a methyl group in the Me4N+ cation, the pink powder obtained from the [Me4N]2CoCl4-catalyzed reaction showed two new peaks at 288.8 eV and 292.5 eV, which correspond to C=O and CF3, respectively. These peaks were also observed in the [Me4N]2CoTFA4 synthesized from [Me4N]2CoCl4 and AgTFA, according to the literature [35,36]. Moreover, in the Co2p XPS analysis data, similar peak positions at 781.7 and 781.6 eV were observed both in the isolated solid after the reaction and in the synthesized [Me4N]2CoTFA4, a clear difference from the 780.8 eV of [Me4N]2CoCl4.
The FT-IR pattern of the pink solid was also very similar to that of synthesized [Me4N]2CoTFA4 (Figure 5b,c). In addition to two peaks at 1489 and 949 cm−1, which were assigned to the CH3 groups in [Me4N]+, a strong peak at 1685 cm−1 was observed in both the isolated pink solid and synthesized [Me4N]2CoTFA4, which can be interpreted to be the carbonyl group (C=O) in TFA (CF3CO2) binding to a Co metal center. Furthermore, two strong peaks at 1185 cm−1 and 1130 cm−1, assigned to the C-F stretching vibration, and two medium intensity peaks at 840 cm−1 and 794 cm−1, assigned to the C-C stretching and CF3 symmetric stretching vibrations, respectively, were also observed in the pink solid and synthesized [Me4N]2CoTFA4 [37,38].
Interestingly, the ligand exchange of CoCl42− to CoTFA42− seems to happen very quickly, even at room temperature. As Figure 5d shows, the IR spectra of the solid obtained from just-dissolved [Me4N]2CoCl4 in HTFA and synthesized [Me4N]2CoTFA4 are almost similar.
However, although [Me4N]2CoCl4 turns to [Me4N]2CoTFA4 in HTFA, the catalytic activity of [Me4N]2CoTFA4 for the methane oxidation reaction was lower than that of [Me4N]2CoCl4; 5.40 mmol of MeTFA was produced, which is lower than the 8.26 mmol for [Me4N]2CoCl4 (Entry 10 vs Entry 1 in Table 1). Because the chloride anion is dissociated from [Me4N]2CoCl4 in the course of changing to CoTFA42−, the effect of chloride anions was investigated by adding 2 equiv. of Me4NCl into the [Me4N]2CoTFA4-catalyzed reaction. Entry 11 in Table 1 reveals that the presence of the free chloride anion did not have much effect on the amount of MeTFA; the amount was changed from 5.40 mmol to 5.75 mmol, while CO2 formation noticeably decreased from 7.56 mmol to 3.76 mmol.
M. N. Vargaftik et al. reported that, besides methane oxidation, solvent HTFA oxidation occurs simultaneously in this aerobic Co-catalyzed methane oxidation reaction [32]. We also observed CO2 in the gas phase after methane oxidation in all the reactions shown in Table 1 and found that used Co catalysts also catalyzed HTFA oxidation to generate CO2 and F compounds like HF and CF3Cl. Meanwhile, MeTFA oxidation in the presence of [Me4N]2CoCl4 was negligible (Entry 3 in Table 1).
Table 1 also shows that the catalysts with very low activities for methane oxidation, CoCl2, CoBr2, and CoSO4, also showed marginal CO2 production between 1.50 and 2.49 mmol. The catalysts showing higher activity for MeTFA production also showed increased CO2 formation. However, the catalytic activity on the solvent oxidation was not always proportional to the catalytic activity on the methane oxidation. The catalyst showing the highest activity on MeTFA production, [Me4N]2CoCl4, gave 5.00 mmol of CO2 formation, while Co(NO3)2 and Co(TFA)2, which showed a lower MeTFA amount than [Me4N]2CoCl4, gave increased CO2 formation of 9.18 and 9.86 mmol, respectively.
It was previously reported that the Co(III) complex oxidizes organic aliphatic acid to produce alkyl radicals and CO2 [39,40]. Similarly, the solvent HTFA in the Co catalyst system could be decomposed to CF3 radicals and CO2 as in Equation (1).
Co ( III )   +   CF 3 CO 2 H Co ( II )   +   CF 3   +   CO 2   +   H +
In fact, a large amount of CF3Cl was detected in the [Me4N]2CoCl4-catalyzed reaction in the gas phase of the reaction product when conducted in the absence of methane, suggesting the CF3 radical from HTFA was captured by the chloride in the reaction media (Figure S5 in Supplementary Materials), and when the same reaction proceeded in the presence of methane, CHF3 and CF3Cl were observed together (Figure S6 in Supplementary Materials). The detection of CHF3 could suggest the formation of MeTFA and CF3 radicals linked by the methyl radical formation [41]. However, considering the formed amounts of CHF3 (0.42 mmol) and MeTFA (8.26 mmol), it can be concluded that the major amount of MeTFA was formed by the direct reaction between Co(III) and methane.
On the other hand, in the [Me4N]2CoTFA4-catalyzed reaction, no CHF3 and CF3Cl were observed (Figure S7 in Supplementary Materials), suggesting the CF3 radicals from the HTFA oxidation further degraded to HF and CO2 in the reaction media [42]. The formation of HF could increase the acid concentration on the reaction media, which is known to accelerate the oxidation of organic acid in a Co catalyst system [39,40]. In other words, the chloride in the reaction media could capture CF3 radicals in the form of CHF3 or CF3Cl, thereby suppressing excessive oxidation of HTFA, compared to a chloride-free system. Entry 10 and Entry 11 in Table 1 show that the CO2 formation decreased from 7.56 mmol to 3.76 mmol with the addition of 2 equiv. of Me4NCl to CoTFA4(Me4N)2.
Overall, the higher catalytic activity of [Me4N]2CoCl4 can be attributed to the high solubility of the catalytically active Co species on HTFA and the presence of chloride anions, which appears to suppress solvent oxidation. Besides this, we suspect that the real catalytic species formed from [Me4N]2CoCl4 during the reaction may not be CoTFA42−, which is an isolated species from the [Me4N]2CoCl4-catalyzed reaction. Instead, the partially substituted CoClxTFAy2− species may be the real form in this methane oxidation reaction.
Figure 2 shows that when the reaction was stopped at 30 min, besides the homogeneous pink solution, a sky-blue solid was observed in the bottom of the reactor liner. After 1 h reaction, a transparent solution was attained with a sky-blue solid. This precipitated sky-blue solid appeared to be CoCl2, based on XRD and XPS analyses (Figure 6).
This result indicates that chloride and TFA reversibly interacted with the Co species in the HTFA media to form CoClxTFAy2− but finally precipitated in the form of poorly soluble CoCl2 at the reaction temperature of 180 °C. The reason [Me4N]2CoTFA4 is isolated from the dissolved [Me4N]2CoCl4 in HTFA can possibly be explained as follows: during the HTFA removal process, the chloride could be removed in the form of HCl, which is more volatile than HTFA. On the other hand, no CoF2 was observed in the XRD analysis of the precipitated sky-blue solid, which was assumed to be a deactivated catalyst form [32].

2.2. Effects of Reaction Conditions

The effect of [Me4N]2CoCl4 concentration on the MeTFA formation was investigated. Figure 7a confirms that by increasing [Me4N]2CoCl4 catalyst concentration from 0.2 mmol to 1.0 mmol, the amount of produced MeTFA linearly increased from 1.85 mmol to 8.26 mmol. Further increasing the [Me4N]2CoCl4 concentration to 1.5 mmol resulted in a slight decrease in MeTFA formation to 8.01 mmol, suggesting the probability of MeTFA oxidation at a higher catalyst concentration. The increased amount of dioxygen, from 5.0 bar to 7.5 bar and 10 bar, resulted in a slight increase in MeTFA formation from 8.26 mmol to 9.03 and 9.60 mmol, respectively (Figure 7b), suggesting the effect of O2 amount on this methane oxidation was insignificant. In contrast, the effect of methane pressure was more pronounced than that of dioxygen (Figure 7c). At the lower pressures of 5 and 10 bar, MeTFA formation was 1.52 mmol and 3.28 mmol, respectively. At the methane pressure of 20 bar, it increased to 8.26 mmol; however, a further increase to 30 bar did not result in a proportional increase in MeTFA formation.
The influence of reaction time on the formation of MeTFA was also explored (Figure 7d). By increasing the reaction time from 30 min to 3 h, the formation of MeTFA continuously increased from 5.23 mmol to 9.69 mmol. At 3 h reaction, although 9.69 mmol of MeTFA was produced, the reactor was severely corroded, forming a dark contaminated solution which originated from the SUS reactor (Supplementary Materials, Figure S8). The effect of reaction temperature was also explored at temperatures between 150–220 °C. Figure S9 in Supplementary Materials revealed the formation of MeTFA was the highest at 180 °C. At the temperature of 150 °C, the produced MeTFA was 2.85 mmol, while at the reaction temperatures of 200 and 220 °C the MeTFA amounts were 7.52 and 5.37 mmol, which were lower than the 8.26 mmol produced at 180 °C. This result indicates MeTFA decomposed to a significant degree at higher reaction temperatures above 200 °C.

3. Experimental Section

3.1. General Information

All chemicals were of analytical reagent grade and used without further purification. Cobalt salts (CoCl2, Co(NO3)2.6H2O, CoBr2, CoSO4.7H2O), tetramethyl ammonium chloride, and other chemicals for the synthesis of catalysts were purchased from Sigma Aldrich Chemical Co.(St. Louis, MO, USA), trifluoroacetic acid and trifluoroacetic anhydride were obtained from Alfa Aesar Co.(Haverhill, MA, USA), and high-purity methane containing 1% of Ar and dioxygen O2 gas (99.995 %) were supplied from Shinyang Gas Co.(Gyeonggi-do, Korea).

3.2. Catalyst Synthesis

[Me4N]2CoCl4 catalyst was synthesized by a simple ligand exchange method presented in the previous report [43]. In a typical sample preparation, cobalt dichloride CoCl2 (0.50 g, 3.85 mmol) and tetramethylammonium chloride (CH3)4NCl (0.84 g, 7.70 mmol) were separately dissolved in ethanol. A sky-blue solid was precipitated by mixing the two solutions at room temperature. After being filtered, washed, and dried under vacuum, bis(tetramethylammonium) tetrachlorocobaltate (II) was obtained. Elemental Analysis (Thermo Scientific Flash 2000 CHNS/O, Thermo Fisher Scientific, Cambridge, UK) Cal. C (27.53 wt%), H (6.93 wt%), N (8.03 wt%). Anal. C (27.31 wt%), H (6.84 wt%), N (7.80 wt%), ICP analysis Cal. Co (16.88 wt%), Anal. Co (16.52 wt%).
Co(TFA)2 catalyst was prepared by the metathesis reaction between CoCl2 (0.46 g, 3.51 mmol) and AgTFA (1.55 g, 7.02 mmol) in 30 g of CH3NO2, as described in the literature [37]. The round flask containing CoCl2 and AgTFA was fully covered using aluminum foil and stirred at rt overnight. After the precipitated AgCl was filtered out, Co(TFA)2 was obtained by removing nitromethane under reduced pressure. Elemental Analysis for Co(TFA)2 Cal. C (16.86 wt%), Anal. C (17.05 wt%), ICP analysis Cal. Co (20.68 wt%), Anal. Co (21.30 wt%).
[Me4N]2CoTFA4 was synthesized by the reaction of [Me4N]2CoCl4 (0.670 g, 2.0 mmol) and AgTFA (1.77g, 8.0 mmol) in nitromethane at room temperature overnight [35,43]. After AgCl was filtered out, a violet-colored product could be obtained by evaporation of the solvent under vacuum. Elemental Analysis for [Me4N]2CoTFA4 Cal. C (29.15 wt%), H (3.67 wt%), N (4.25 wt%), Anal. C (29.31 wt%), H (3.58 wt%), N (4.50 wt%), ICP analysis Cal. Co (8.94 wt%), Anal. Co (8.82 wt%).

3.3. Methane Oxidation Reaction

The partial oxidation of methane was conducted using a bomb reactor (SS304, 100 mL) equipped with a glass liner, thermocouple, pressure gauge, and heating jacket. In a typical reaction, a glass liner containing 1.00 mmol of catalyst, 25 g of trifluoroacetic acid, and 5 g of trifluoroacetic anhydride was introduced into the reactor and pressurized with 5 bar of O2 (17.2 mmol) and 20 bar of CH4 (68.6 mmol) at room temperature. The reactor was subsequently heated to 180 °C and stirred (600 rpm) for 1 h.

3.4. Analysis

After the reaction, a liquid product containing MeTFA was analyzed using GC (7890A, Agilent Technologies, Santa Clara, CA, USA). CHCl3 was used as an external reference for the quantitation of MeTFA. For the quantitation of CO2 formed during the oxidation, the gas collected from the reactor was analyzed using GC-MS (HP 6890 GC/5973 MSD, Agilent Technologies, Santa Clara, CA, USA) equipped with a capillary column (Poraplot Q 30 m × 25 um). Ar gas which was included in methane (1%) was used as an internal reference. The cobalt concentration in the liquid phase was determined by ICP-OES (iCAP 700 series-ICP spectrometer, Thermo Fisher Scientific, Waltham, MA, USA) and the X-ray diffraction patterns (XRD) were recorded on a Shimadzu X-ray diffractometer (XRD-6000, Shimadzu, Kyoto, Japan) using nickel-filtered CuKα radiation with 2θ angle from 10 to 80°. FT-IR spectra were obtained by Thermo Nicolet iS10 FTIR (Thermo Fisher Scientific, Waltham, MA, USA) equipped with a Smart Diamond ATR accessory. CNH analysis was conducted using Thermal scientific flash 2000 CHNS/O (Thermo Fisher Scientific, Cambridge, UK). XPS analysis was taken by Nexsa (Thermo Fischer Scientific, Waltham, MA, USA) with a microfocus monochromatic X-ray source (Al-Kα = 1486.6 eV).

4. Conclusions

An [Me4N]2CoCl4-catalyzed aerobic methane oxidation was conducted in HTFA solvent at 180 °C. Because of its ionic structure, it dissolved well in HTFA, even at room temperature, and exhibited the highest catalytic activity among the tested Co catalysts, including CoCl2, CoBr2, CoSO4, Co(NO3)2, Co(TFA)2, and [Me4N]2CoTFA4. Although [Me4N]2CoTFA4 was isolated from the reaction media, the real catalytic species were assumed to be [CoClxTFAy]2−, and it finally turned into CoCl2. It was found that HTFA oxidation to CO2 was also catalyzed by the methane oxidation catalyst and that the presence of chloride in the reaction media appeared to suppress HTFA oxidation to some extent. This could be another reason for the high catalytic activity of [Me4N]2CoCl4 compared to the other chloride-free Co catalysts, such as Co(TFA)2 and CoTFA4(Me4N)2. The production of MeTFA was heavily affected by the reaction condition, including reaction time, catalyst amount, and methane pressure, but the effect of O2 pressure was insignificant. MeTFA was quite stable at 180 °C, but it was further oxidized above 200 °C.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111419/s1, Figure S1. Photos of pressure reactor and reaction system for the methane oxidation; Figure S2. Calibration graph for the quantitation of MeTFA; Figure S3. GC-FID analysis of the liquid product; Figure S4. Calibration curve for CO2; Figure S5. GC-Mass analysis of gas product obtained from the reaction conducted in the absence of methane; Figure S6. GC-Mass analysis of gas product obtained from the [Me4N]2CoCl4 –catalyzed reaction conducted in the presence of methane; Figure S7. GC-Mass analysis of gas product obtained from [Me4N]2CoTFA4 –catalyzed reaction; Figure S8. Photo of the solution after 3 h reaction at 180 °C; Figure S9. Effect of reaction temperature on MeTFA formation; Table S1. A comparison of various catalyst systems for the oxidation of methane to methyl trifluoroacetate. References [14,15,21,24,26,28,32,44,45,46,47,48] are cited in the supplementary materials.

Author Contributions

Conceptualization, H.T.D. and H.L.; Data curation, H.T.D., S.C., J.K. and N.T.T.; Investigation, H.L.; Methodology, J.K. and H.K.; Project administration, H.L.; Supervision, H.L.; Writing—original draft, H.T.D.; Writing—review & editing, H.T.D. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by C1 Gas Refinery Program through the National Research Foundation of Korea (NRF) (NRF-2015M3D3A1A01065435), KIST internal program (Atmospheric Environment Research Program, Project No. 2E31690) and UST Young Scientist Program 2020 (Project No. 2020YS004).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Olah, G.A. Beyond oil and gas: The methanol economy. Angew. Chem. Int. Ed. 2005, 44, 2636–2639. [Google Scholar] [CrossRef] [PubMed]
  2. Simon Araya, S.; Liso, V.; Cui, X.; Li, N.; Zhu, J.; Sahlin, S.L.; Jensen, S.H.; Nielsen, M.P.; Kær, S.K. A Review of The Methanol Economy: The Fuel Cell Route. Energies 2020, 13, 596. [Google Scholar] [CrossRef] [Green Version]
  3. Alvarez-Galvan, M.C.; Mota, N.; Ojeda, M.; Rojas, S.; Navarro, R.M.; Fierro, J.L.G. Direct methane conversion routes to chemicals and fuels. Catal. Today 2011, 171, 15–23. [Google Scholar] [CrossRef]
  4. Meng, X.; Cui, X.; Rajan, N.P.; Yu, L.; Deng, D.; Bao, X. Direct Methane Conversion under Mild Condition by Thermo-, Electro-, or Photocatalysis. Chem 2019, 5, 2296–2325. [Google Scholar] [CrossRef]
  5. Gesser, H.D.; Hunter, N.R. The Direct Conversion of Methane to Methanol by Controlled Oxidation. Chem. Rev. 1985, 85, 235–244. [Google Scholar] [CrossRef]
  6. Stahl, S.S.; Labinger, J.A.; Bercaw, J.E. Homogeneous oxidation of alkanes by electrophilic late transition metals. Angew. Chem. Int. Ed. 1998, 37, 2181–2192. [Google Scholar] [CrossRef]
  7. Ravi, M.; Ranocchiari, M.; van Bokhoven, J.A. The Direct Catalytic Oxidation of Methane to Methanol—A Critical Assessment. Angew. Chem. Int. Ed. Engl. 2017, 56, 16464–16483. [Google Scholar] [CrossRef]
  8. Periana, R.A.; Taube, D.J.; Evitt, E.R.; Löffler, D.G.; Wentrcek, P.R.; Voss, G.; Masuda, T. A Mercury-Catalyzed, High-Yield System for the Oxidation of Methane to Methanol. Science 1993, 259, 340–343. [Google Scholar] [CrossRef]
  9. Periana, R.A.; Taube, D.J.; Gamble, S.; Taube, H.; Satoh, T.; Fujii, H. Platinum catalysts for the high-yield oxidation of methane to a methanol derivative. Science 1998, 280, 560–564. [Google Scholar] [CrossRef]
  10. Zimmermann, T.; Soorholtz, M.; Bilke, M.; Schüth, F. Selective Methane Oxidation Catalyzed by Platinum Salts in Oleum at Turnover Frequencies of Large-Scale Industrial Processes. J. Am. Chem. Soc. 2016, 138, 12395–12400. [Google Scholar] [CrossRef]
  11. Dang, H.T.; Lee, H.W.; Lee, J.; Choo, H.; Hong, S.H.; Cheong, M.; Lee, H. Enhanced Catalytic Activity of (DMSO)2PtCl2 for the Methane Oxidation in the SO3–H2SO4 System. ACS Catal. 2018, 8, 11854–11862. [Google Scholar] [CrossRef]
  12. Lee, H.W.; Dang, H.T.; Kim, H.; Lee, U.; Ha, J.-M.; Jae, J.; Cheong, M.; Lee, H. Pt black catalyzed methane oxidation to methyl bisulfate in H2SO4-SO3. J. Catal. 2019, 374, 230–236. [Google Scholar] [CrossRef]
  13. Meyer, D.; Strassner, T. CH-activation of methane—Synthesis of an intermediate? J. Organomet. Chem. 2015, 784, 84–87. [Google Scholar] [CrossRef]
  14. Ravi, M.; van Bokhoven, J.A. Homogeneous Copper-Catalyzed Conversion of Methane to Methyl Trifluoroacetate in High Yield at Low Pressure. ChemCatChem 2018, 10, 2383–2386. [Google Scholar] [CrossRef]
  15. Muehlhofer, M.; Strassner, T.; Herrmann, W.A. New Catalyst Systems for the Catalytic Conversion of Methane into Methanol. Angew. Chem. Int. Ed. 2002, 41, 1745–1747. [Google Scholar] [CrossRef]
  16. Campbell, A.N.; Stahl, S.S. Overcoming the Oxidant problem Strategies to use O2 as the oxidant in organometallic CH oxidation reactions catalyzed by Pd and Cu. Acc. Chem. Res. 2012, 45, 851–863. [Google Scholar] [CrossRef]
  17. Kalman, S.E.; Munz, D.; Fortman, G.C.; Boaz, N.C.; Groves, J.T.; Gunnoe, T.B. Partial oxidation of light alkanes by periodate and chloride salts. Dalton Trans. 2015, 44, 5294–5298. [Google Scholar] [CrossRef]
  18. Fortman, G.C.; Boaz, N.C.; Munz, D.; Konnick, M.M.; Periana, R.A.; Groves, J.T.; Gunnoe, T.B. Selective monooxidation of light alkanes using chloride and iodate. J. Am. Chem. Soc. 2014, 136, 8393–8401. [Google Scholar] [CrossRef]
  19. Liebov, N.S.; Goldberg, J.M.; Boaz, N.C.; Coutard, N.; Kalman, S.E.; Zhuang, T.; Groves, J.T.; Gunnoe, T.B. Selective Photo-Oxygenation of Light Alkanes Using Iodine Oxides and Chloride. ChemCatChem 2019, 11, 5045–5054. [Google Scholar] [CrossRef]
  20. Schwartz, N.A.; Boaz, N.C.; Kalman, S.E.; Zhuang, T.; Goldberg, J.M.; Fu, R.; Nielsen, R.J.; Goddard, W.A.; Groves, J.T.; Gunnoe, T.B. Mechanism of Hydrocarbon Functionalization by an Iodate/Chloride System: The Role of Ester Protection. ACS Catal. 2018, 8, 3138–3149. [Google Scholar] [CrossRef]
  21. Cheong, S.-H.; Kim, D.; Dang, H.T.; Kim, D.; Seo, B.; Cheong, M.; Hong, S.H.; Lee, H. Methane oxidation to methyl trifluoroacetate by simple anionic palladium catalyst: Comprehensive understanding of K2S2O8-based methane oxidation in CF3CO2H. J. Catal. 2022, 413, 803–811. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Zhang, M.; Han, Z.; Huang, S.; Yuan, D.; Su, W. Atmosphere-Pressure Methane Oxidation to Methyl Trifluoroacetate Enabled by a Porous Organic Polymer-Supported Single-Site Palladium Catalyst. ACS Catal. 2021, 11, 1008–1013. [Google Scholar] [CrossRef]
  23. Munz, D.; Strassner, T. Alkane C-H functionalization and oxidation with molecular oxygen. Inorg. Chem. 2015, 54, 5043–5052. [Google Scholar] [CrossRef] [PubMed]
  24. Strassner, T.; Ahrens, S.; Muehlhofer, M.; Munz, D.; Zeller, A. Cobalt-Catalyzed Oxidation of Methane to Methyl Trifluoroacetate by Dioxygen. Eur. J. Inorg. Chem. 2013, 3659–3663. [Google Scholar] [CrossRef]
  25. An, Z.; Pan, X.; Liu, X.; Han, X.; Bao, X. Combined redox couples for catalytic oxidation of methane by dioxygen at low temperature. J. Am. Chem. Soc. 2006, 128, 16028–16029. [Google Scholar] [CrossRef]
  26. Yuan, J.; Wang, L.; Wang, Y. Direct Oxidation of Methane to a Methanol Derivative Using Molecular Oxygen. Ind. Eng. Chem. Res. 2011, 50, 6513–6516. [Google Scholar] [CrossRef]
  27. Yuan, J.; Liu, L.; Wang, L.; Hao, C. Partial Oxidation of Methane with the Catalysis of Palladium(II) and Molybdovanadophosphoric Acid Using Molecular Oxygen as the Oxidant. Catal. Lett. 2012, 143, 126–129. [Google Scholar] [CrossRef]
  28. Blankenship, A.N.; Ravi, M.; Newton, M.A.; van Bokhoven, J.A. Heterogeneously Catalyzed Aerobic Oxidation of Methane to a Methyl Derivative. Angew. Chem. Int. Ed. 2021, 60, 18138–18143. [Google Scholar] [CrossRef]
  29. Coutard, N.; Goldberg, J.M.; Valle, H.U.; Cao, Y.; Jia, X.; Jeffrey, P.D.; Gunnoe, T.B.; Groves, J.T. Aerobic Partial Oxidation of Alkanes Using Photodriven Iron Catalysis. Inorg. Chem. 2022, 61, 759–766. [Google Scholar] [CrossRef]
  30. Coutard, N.; Musgrave, C.B.; Moon, J.; Liebov, N.S.; Nielsen, R.M.; Goldberg, J.M.; Li, M.; Jia, X.; Lee, S.; Dickie, D.A.; et al. Manganese Catalyzed Partial Oxidation of Light Alkanes. ACS Catal. 2022, 12, 5356–5370. [Google Scholar] [CrossRef]
  31. Tang, R.; Kochi, J.K. Cobalt (III) Trifluoroacetate: An Electron Transfer Oxidant. J. Inorg. Nucl. Chem. 1973, 35, 3845–3856. [Google Scholar] [CrossRef]
  32. Vargaftik, M.N.; Stolarov, I.P.; Moiseev, I.I. Highly Selective Parital Oxidation of Methane to Methyl Trifluoroacetate. Chem. Commun. 1990, 15, 1049–1050. [Google Scholar] [CrossRef]
  33. Wiesner, J.R.; Srivastava, R.C.; Kennard, C.H.L.; Divaira, M.; Lingafelter, E.C. The Crystal Structures of Tetramethylammonium Tetrachloro-cobaltate (II), -nickelate (II) and -zincate (II). Acta Crystallogr. 1966, 23, 565–574. [Google Scholar] [CrossRef]
  34. Stolarov, I.P.; Vargaftik, M.N.; Shishkin, D.I.; Moiseev, I.I. Oxidation of Ethane and Propane with Cobalt (II) Catalyst: Unexpected Formation of 1,2-Diol Esters and C-C Bond Cleavage. J. Chem. Soc. Chem. Commun. 1991, 14, 938–939. [Google Scholar] [CrossRef]
  35. Puri, M.; Verma, R.D. Bis(tetramethylammonium) tetrakis (trifluoroacetato) metallates (II). Indian J. Chem. 1983, 22, 418–419. [Google Scholar]
  36. Bergman Jr, J.G.; Cotton, F.A. The Preparation, Properties, and Struture of Tetraphenylarsonium Tetrakis (trifluoroacetato) cobaltate (II). Inorg. Chem. 1966, 5, 1420–1424. [Google Scholar] [CrossRef]
  37. Baillie, M.J.; Brown, D.H.; Moss, K.C.; Sharp, D.W.A. Anhydrous Metal Trifluoroacetates. J. Chem. Soc. A 1968, 3110–3114. [Google Scholar] [CrossRef]
  38. Agambar, C.A.; Orrell, K.A. Trifuoroacetate Complexes of Cobalt(II), Nickel(II), and Copper(II) with pyridine-type ligands. J. Chem. Soc. 1974, 8, 897–904. [Google Scholar]
  39. Clifford, A.A.; Waters, W.A. Oxidations of organic compounds by cobaltic salts. Part VI. Kinetic and product studies of the oxidation of some carboxylic acids. J. Chem. Soc. 1965, 2796–2804. [Google Scholar] [CrossRef]
  40. Lande, S.S.; Kochi, J.K. Formation and Oxidation of Alkyl Radicals by Cobalt (III) Complexes. J. Am. Chem. Soc. 1968, 90, 5196–5207. [Google Scholar] [CrossRef]
  41. Zargari, N.; Winter, P.; Liang, Y.; Lee, J.H.; Cooksy, A.; Houk, K.N.; Jung, K.W. Unexpected, Latent Radical Reaction of Methane Propagated by Trifluoromethyl Radicals. J. Org. Chem. 2016, 81, 9820–9825. [Google Scholar] [CrossRef] [PubMed]
  42. Hori, H.; Tanako, Y.; Koike, K.; Takeuchi, K.; Einaga, H. Decomposition of Environmentally Persistent Trifluoroacetic Acid to Fluoride Ions by Homogeneous Photocatalyst in Water. Environ. Sci. Technol. 2003, 37, 418–422. [Google Scholar] [CrossRef] [PubMed]
  43. Puri, J.K.; Chhoker, R. Synthesis and Characterization of Metal Tribromoacetates and Their Alkali Metal Derivatives. J. Chem. 2013, 2013, 392310. [Google Scholar] [CrossRef] [Green Version]
  44. Piao, D.-g.; Inoue, K.; Shibasaki, H.; Taniguchi, Y.; Kitamura, T.; Fujiwara, Y. An efficient partial oxidation of methane in trifluoroacetic acid using vanadium-containing heteroplolyacid catalysts. J. Organomet. Chem. 1999, 574, 116–120. [Google Scholar] [CrossRef]
  45. Yamanaka, I.; Soma, M.; Otsuka, K. Oxidation of methane to methanol with oxygen catalysed by Europium trichloride at room temperature. Chem. Commun. 1995, 2235–2236. [Google Scholar] [CrossRef]
  46. Chen, W.; Kocal, J.A.; Brandvold, T.A.; Bricker, M.L.; Bare, S.R.; Broach, R.W.; Greenlay, N.; Popp, K.; Walenga, J.T.; Yang, S.S.; et al. Manganese oxide catalyzed methane partial oxidation in trifluoroacetic acid: Catalysis and kinetic analysis. Catal. Today 2009, 140, 157–161. [Google Scholar] [CrossRef]
  47. Ingrosso, G.; Midollini, N. Palladium(II)- or copper(II)-catalysed solution-phase oxyfunctionalisation of methane and other light alkanes by hydrogen peroxide in trifluoroacetic anhydride. J. Mol. Catal. A Chem 2003, 204–205, 425–431. [Google Scholar] [CrossRef]
  48. Seki, Y.; Min, J.S.; Misono, M.; Mizuno, N. Reaction Mechanism of Oxidation of Methane with Hydrogen Peroxide Catalyzed by 11-Molybdo-1-vanadophosphoric Acid Catalyst Precursor. J. Phys. Chem. B 2000, 104, 5940–5944. [Google Scholar] [CrossRef]
Figure 1. Powder XRD patterns of (a) synthesized [Me4N]2CoCl4 and (b) [Me4N]2CoCl4 reference (JCPDS 01-072-0033).
Figure 1. Powder XRD patterns of (a) synthesized [Me4N]2CoCl4 and (b) [Me4N]2CoCl4 reference (JCPDS 01-072-0033).
Catalysts 12 01419 g001
Scheme 1. Oxidation of methane to methyl trifluoroacetate (MeTFA) by dioxygen.
Scheme 1. Oxidation of methane to methyl trifluoroacetate (MeTFA) by dioxygen.
Catalysts 12 01419 sch001
Figure 2. Photos of solutions of (a) CoCl2 in HTFA at rt, (b) [Me4N]2CoCl4-in HTFA at rt, (c) [Me4N]2CoCl4-catalyzed methane oxidation for 30 min, and (d) for 1 h.
Figure 2. Photos of solutions of (a) CoCl2 in HTFA at rt, (b) [Me4N]2CoCl4-in HTFA at rt, (c) [Me4N]2CoCl4-catalyzed methane oxidation for 30 min, and (d) for 1 h.
Catalysts 12 01419 g002
Figure 3. XPS spectra of (a) [Me4N]2CoCl4 and (b) isolated pink solid from [Me4N]2CoCl4-catalyzed reaction.
Figure 3. XPS spectra of (a) [Me4N]2CoCl4 and (b) isolated pink solid from [Me4N]2CoCl4-catalyzed reaction.
Catalysts 12 01419 g003
Figure 4. C1s and Co2p XPS analyses of CoCl2 reagent (a,e), [Me4N]2CoCl4 (b,f), pink solid isolated from [Me4N]2CoCl4-catalyzed reaction (c,g), and CoTFA4(Me4N)2 synthesized from [Me4N]2CoCl4 and AgTFA (d,h).
Figure 4. C1s and Co2p XPS analyses of CoCl2 reagent (a,e), [Me4N]2CoCl4 (b,f), pink solid isolated from [Me4N]2CoCl4-catalyzed reaction (c,g), and CoTFA4(Me4N)2 synthesized from [Me4N]2CoCl4 and AgTFA (d,h).
Catalysts 12 01419 g004
Figure 5. FT−IR analyses spectra of (a) [Me4N]2CoCl4, (b) isolated solid from [Me4N]2CoCl4-catalyzed reaction, (c) synthesized [Me4N]2CoTFA4 from [Me4N]2CoCl4 and AgTFA, and (d) pink solid isolated from [Me4N]2CoCl4 dissolved in HTFA and (e) Me4NCl.
Figure 5. FT−IR analyses spectra of (a) [Me4N]2CoCl4, (b) isolated solid from [Me4N]2CoCl4-catalyzed reaction, (c) synthesized [Me4N]2CoTFA4 from [Me4N]2CoCl4 and AgTFA, and (d) pink solid isolated from [Me4N]2CoCl4 dissolved in HTFA and (e) Me4NCl.
Catalysts 12 01419 g005
Figure 6. (a) Powder XRD diffractogram and XPS analysis results of (b) Wide scan and (c) Co2p of the precipitated solid collected after [Me4N]2CoCl4-catalyzed methane oxidation at 180 °C for 1 h.
Figure 6. (a) Powder XRD diffractogram and XPS analysis results of (b) Wide scan and (c) Co2p of the precipitated solid collected after [Me4N]2CoCl4-catalyzed methane oxidation at 180 °C for 1 h.
Catalysts 12 01419 g006
Figure 7. Effect of (a) catalyst concentration, (b) O2 pressure, (c) CH4 pressure, and (d) reaction time on the [Me4N]2CoCl4-catalyzed methane oxidation. Conditions: 1 mmol of catalyst, 5 bar of O2, 20 bar of methane, 5 g of TFAA, 25 g of HTFA, 180 °C, 1 h.
Figure 7. Effect of (a) catalyst concentration, (b) O2 pressure, (c) CH4 pressure, and (d) reaction time on the [Me4N]2CoCl4-catalyzed methane oxidation. Conditions: 1 mmol of catalyst, 5 bar of O2, 20 bar of methane, 5 g of TFAA, 25 g of HTFA, 180 °C, 1 h.
Catalysts 12 01419 g007
Table 1. Aerobic methane oxidation using various Co catalysts a.
Table 1. Aerobic methane oxidation using various Co catalysts a.
EntryCatalyst SystemMeTFA (mmol,
molester molmetal−1h−1)
CO2 (mmol)
1[Me4N]2CoCl48.265.00
2 b[Me4N]2CoCl4010.59
3 c[Me4N]2CoCl49.739.62
4CoCl20.331.50
5CoCl2 + 2Me4NCl5.523.82
6Co(NO3)2.6H2O4.389.18
7Co(TFA)24.749.86
8CoBr20.122.49
9CoSO4.7H2O0.101.93
10CoTFA4(Me4N)25.407.56
11CoTFA4(Me4N)2 + 2Me4NCl5.753.76
12--0.25
a Reaction conditions: 1 mmol of catalyst, 5 bar of O2, 20 bar of methane, 5 g of TFAA, 25 g of HTFA, 180 °C, 1 h. b The reaction was carried out in the absence of methane. c 10 mmol of MeTFA was reacted instead of methane.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dang, H.T.; Cheong, S.; Kim, J.; Tran, N.T.; Kim, H.; Lee, H. Tetrachlorocobaltate-Catalyzed Methane Oxidation to Methyl Trifluoroacetate. Catalysts 2022, 12, 1419. https://doi.org/10.3390/catal12111419

AMA Style

Dang HT, Cheong S, Kim J, Tran NT, Kim H, Lee H. Tetrachlorocobaltate-Catalyzed Methane Oxidation to Methyl Trifluoroacetate. Catalysts. 2022; 12(11):1419. https://doi.org/10.3390/catal12111419

Chicago/Turabian Style

Dang, Huyen Tran, Seokhyeon Cheong, Jiyun Kim, Ngoc Tuan Tran, Honggon Kim, and Hyunjoo Lee. 2022. "Tetrachlorocobaltate-Catalyzed Methane Oxidation to Methyl Trifluoroacetate" Catalysts 12, no. 11: 1419. https://doi.org/10.3390/catal12111419

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

Article Metrics

Back to TopTop