Skip to main content
Advanced Science logoLink to Advanced Science
. 2017 Sep 11;4(12):1700379. doi: 10.1002/advs.201700379

Ultrahigh Electrocatalytic Conversion of Methane at Room Temperature

Ming Ma 1,2, Bing Jun Jin 1, Ping Li 1,2, Myung Sun Jung 1, Jin Il Kim 1, Yoonjun Cho 1, Sungsoon Kim 1, Jun Hyuk Moon 3, Jong Hyeok Park 1,
PMCID: PMC5737395  PMID: 29270346

Abstract

Due to the greenhouse effect, enormous efforts are done for carbon dioxide reduction. By contrast, more attention should be paid for the methane oxidation and conversion, which can help the effective utilization of methane without emission. However, methane conversion and utilization under ambient conditions remains a challenge. Here, this study designs a Co3O4/ZrO2 nanocomposite for the electrochemical oxidation of methane gas using a carbonate electrolyte at room temperature. Co3O4 activated the highly efficient oxidation of methane under mild electric energy with the help of carbonate as an oxidant, which is delivered by ZrO2. Based on the experimental results, acetaldehyde is the key intermediate product. Subsequent nucleophilic addition and free radical addition reactions accounted for the generation of 2‐propanol and 1‐propanol, respectively. Surprisingly, this work achieves a production efficiency of over 60% in the conversion of methane to produce these long‐term stable products. The as‐proposed regional electrochemical methane oxidation provides a new pathway for the synthesis of higher alcohols with high production efficiencies under ambient conditions.

Keywords: 1‐propanol, 2‐propanol, electrochemical, methane oxidation, ZrO2/Co3O4

1. Introduction

Methane (CH4) is the primary component of natural gas, which constitutes 21.4% of the total primary energy sources in the world,1 and is widely used as an important fuel in both industrial chemical process and human daily life. Compared with other fossil fuels, such as oil and coal, the combustion of natural gas provides lower carbon dioxide (CO2) emissions,2 making natural gas a suitable alternative transitional energy source until carbon‐free energy sources are sufficiently mature to be deployed.3 However, the emission of CH4 gas has long been ignored and regarded as a trivial matter,4 even though its effect as a greenhouse gas is over 30 times more potent than that of CO2.5 In particular, global warming and the exploitation of shale gas aggravate this emission. Recently, more attention has been given to the negative impact of CH4 emissions due to increasing environmental pollution and climate change.6, 7, 8 Hence, efforts have focused on the conversion of atmospheric CH4 into an equimolar amount of CO2 through thermocatalysis or photocatalysis. However, conventional CH4 conversion processes still suffer from various drawbacks, including the use of precious metal catalysts, high reaction temperatures that are far from ambient conditions and extremely low conversion efficiencies.9, 10, 11 In this regard, the oxidation and conversion of CH4 to liquid alcohols, such as methanol, ethanol, and propanol, is much more economical and energy‐efficient. Among liquid alcohols, higher alcohols with high energy densities have wide applications in fabricating commodity chemicals.12

Methanol, as a product of CH4 conversion, has been widely investigated. Currently, the syngas reaction is the main route for the industrial production of methanol, whereby syngas is produced through CH4 steam reforming. The two reactions for CH4 conversion to methanol are given in Equations (1) and (2), 13, 14

CH4(g)+H2O(g)NiCO(g)+3H2(g)ΔH298=49.3kcalmol1 (1)
CO(g)+2H2(g)Cu/ZnO/Al2O3CH3OH(g)ΔH298=21.7kcalmol1 (2)

From the equations, the initial reaction in syngas reforming requires more energy than that released from the second reaction, illustrating that additional energy input is necessary for the conversion of CH4 to methanol. Some scientists have employed a class of bacteria known as methanotrophs with methane mono‐oxygenases to convert CH4 into methanol under ambient conditions using oxygen.15 The methods mentioned above require complex processes, extra energy consumption, or enzyme cultivation and critical control of the conditions.13, 16 In fact, the direct oxidation of CH4 by oxygen gas (O2) is always accompanied by substantial overoxidation, which is both kinetically and thermodynamically favorable.13 The conversion of CH4 to methanol with O2 is triggered by energy input and proceeds spontaneously with a substantial release of energy, which makes it exergonic for the further oxidation of methanol to formaldehyde, formic acid, carbon monoxide (CO), and CO2. Controlling CH4 oxidation to obtain methanol is extremely difficult. Thus, compared with the route of inhibiting CH4 overoxidation to obtain methanol, the control of further oxidation processes under mild conditions may enable the production of more useful higher alcohols, such as propanol. Electrochemical oxidation in aqueous electrolyte has been demonstrated to be a suitable method for the conversion of CH4 or other volatile organic compounds (VOCs) at low temperature using simple reaction instruments.17, 18, 19

The C—H bonds in CH4 have a high dissociation energy of 104 kcal mol−1, which makes CH4 extremely inert. When O2 is used as an oxidizing agent for CH4 oxidation, the reaction of triplet O2 with singlet CH4 to form singlet methanol is a spin‐forbidden process.13 Thus, protons in CH4 are not expected to be readily abstracted by O2 under mild conditions, such as low temperature. Therefore, other oxidizing agents are necessary to replace O2. In conventional alkaline electrochemical systems, which are always performed at room temperature, hydroxide (HO) generally functions as the oxidizing agent. However, hydroxide has previously been demonstrated to have negligible activity for abstracting protons from CH4 under mild conditions.20 By contrast, carbonate (CO3 2−) oxidizes species by donating a charged oxygen atom accompanied by CO2 release, resulting in a large enthalpy of reaction and favorable oxidation kinetics.21, 22, 23 Therefore, CO3 2− may be an attractive alternative to OH for alkaline electrochemical reactions. In addition, to realize the donation of an oxidizing agent from CO3 2−, zirconia (ZrO2) should be employed to facilitate CO3 2− adsorption because of its surface Lewis acid sites and electron‐accepting capabilities.24, 25 Based on the above results, Mustain's group selected nickel oxide (NiO) as the catalyst and prepared NiO/ZrO2 for the electrochemical oxidation of CH4 with CO3 2− as the oxidizing agent.26 However, from their results, NiO did not exhibit satisfactory selectivity for CH4 oxidation, and the reaction mechanism was also unclear. Thus, more efficient catalysts with high activity and selectivity should be employed.

Among metal oxide catalysts, cobalt oxide (Co3O4) has been demonstrated to be one of the most efficient catalysts for the oxidation of VOCs.27 In addition, Co3O4 has a strong surface adsorption capacity for formaldehyde.28 According to theoretical observations of the reduction of CO2 on transition metal surfaces,29 formaldehyde tends to be oxidized to CO. However, when adsorbed on the surface of Co3O4, formaldehyde, converted from methanol, may be more active for further additional reactions to higher alcohols, illustrating the regional selectivity of Co3O4 for CH4 oxidation. Thus, we designed a Co3O4/ZrO2 composite for the electrochemical oxidation of CH4, using CO3 2− as the oxidizing agent source, that resulted in high selectivity for the production of 1‐propoanl and 2‐propanol with over 60% production efficiency. The reaction process is shown in Scheme 1 .

Scheme 1.

Scheme 1

The reaction process of electrochemical oxidation of methane gas.

2. Results and Discussion

2.1. Fabrication and Characterization of the Co3O4/ZrO2 Nanocomposite

Coprecipitation and hydrothermal methods were employed to fabricate the Co3O4/ZrO2 nanocomposite (for details, see the Experimental Section). Scanning electron microscopy (SEM) was used to observe the morphology of different samples fabricated using different ZrO2/Co3O4 ratios, denoted 1–2 ZrO2/Co3O4, 1–4 ZrO2/Co3O4, and 1–6 ZrO2/Co3O4, as shown in Figure 1 . In all samples, oval‐shaped ZrO2 nanoparticles with uniform sizes were formed and were adsorbed on the surface of the Co3O4 plates. Upon increasing the amount of Co3O4, the particle size of Co3O4 gradually increased until it reached bulk (Figure 1; and Figure S1, Supporting Information), which may affect the catalytic properties of the ZrO2/Co3O4 nanocomposite. Pure Co3O4 powder was also prepared for comparison (Figure S2, Supporting Information) and showed large particles that were more than 10 µm in size. The size of the Co3O4 plates can be controlled through coprecipitation with ZrO2. The elemental ratios of Zr/Co for all samples were obtained through EDS (energy‐dispersive X‐ray spectroscopy) measurement (Table S1 and Figure S3, Supporting Information). The Zr/Co ratio decreased upon increasing the amount of Co precursor. In the 1–6 ZrO2/Co3O4 sample, big Co3O4 plates were surrounded by small ZrO2 particles, which may have caused the sudden decrease in the Zr/Co ratio. The EDS measurement was focused on the surface of the samples, which would be affected much by the surface state of the materials. Thus, the ratios showed different value when compared to the stoichiometric ratio of elements calculated from the reactants described in the Experimental Section. In order to obtain the stoichiometric ratio of Zr/Co, ICP‐OES (inductively coupled plasma optical emission spectrometry) measurement was conducted, shown in Table S1 (Supporting Information). The 1–4 ZrO2/Co3O4 sample exhibited relatively small particle sizes and a suitable amount of Co3O4 and showed the best electrochemical catalytic performance for CH4 oxidation, which will be discussed next.

Figure 1.

Figure 1

Morphologies of the ZrO2/Co3O4 nanocomposites with different ratios. SEM images of the a,b) 1–2 ZrO2/Co3O4, c,d) 1–4 ZrO2/Co3O4, and e,f) 1–6 ZrO2/Co3O4 samples. The scale bars in (a)–(f) are 1 µm. g) HR‐TEM image and h) TEM image with elemental mappings of the 1–4 ZrO2/Co3O4 sample. The inserts are fast‐Fourier transformation patterns.

For further examination of ZrO2 and Co3O4, 1–4 ZrO2/Co3O4 was examined by transmission electron microscopy (TEM) (Figure 1g,h). In the TEM image (Figure 1h), the Co3O4 plate can be easily distinguished from the surrounding ZrO2 nanoparticles. From the high‐resolution TEM (HR‐TEM) image (Figure 1g), the lattice constant of ZrO2 was found to be 0.252 nm, corresponding to the (002) facet, and that of Co3O4 was 0.238 nm, corresponding to the (022) facet. In addition, elemental mapping was performed for Co, Zr, and O atoms (Figure 1h). From the elemental distribution, the Co3O4 plates and ZrO2 nanoparticles were confirmed to have different structures. From the SEM and TEM images, well physical connection between ZrO2 and Co3O4 could be demonstrated, which may lead to synergistic effects in the electrochemical oxidation of CH4. In addition, from the X‐ray diffraction (XRD) and X‐ray photoelectron spectroscopy (XPS) spectra (Figure 2 ), peaks of each sample showed almost same position without shifting, illustrating no obvious chemical bondings can be observed between ZrO2 and Co3O4.

Figure 2.

Figure 2

Characteristics of the ZrO2/Co3O4 nanocomposite. XRD spectra of a) the 1–2 ZrO2/Co3O4, 1–4 ZrO2/Co3O4, and 1–6 ZrO2/Co3O4 samples and b) pure Co3O4. c) Co 2p XPS signals of pure Co3O4 and ZrO2/Co3O4 samples with different ratios of 1–2, 1–4, and 1–6. d) Zr 3d XPS signals of ZrO2/Co3O4 samples with different ratios of 1–2, 1–4, and 1–6. e) Deconvoluted O 1s XPS signals of pure Co3O4 and ZrO2/Co3O4 samples with different ratios of 1–2, 1–4, and 1–6.

The crystalline structures of the ZrO2/Co3O4 composites with different ratios were analyzed by powder XRD using Cu Kα radiation and then compared with that of pure Co3O4 (Figure 2a,b). The diffraction peaks of ZrO2 corresponded to the monoclinic phase (JCPDS No. 37‐1484), and the peaks of Co3O4 were indexed to the cubic structure (JCPDS No. 42‐1467). In the XRD patterns, the typical (001), (100), (011), (−111), and (022) planes of ZrO2 were observed at ≈17.5°, 24.2°, 24.6°, 28.3°, and 50.3°. All related peaks gradually decreased in intensity as the amount of Co3O4 increased. The typical (111), (311), and (440) planes of Co3O4 were observed at ≈19.0°, 36.9°, and 65.2°, which also showed the same change in peak intensity with an increase in ZrO2.

The abovementioned data explained the microscopic and crystalline structures of the ZrO2/Co3O4 nanocomposite. Therefore, we next set out to determine the surface state, which is responsible for the unique regional selectivity of CH4 oxidation, by XPS, as shown in Figure 2c–e. XPS spectra of the Co 2p (Figure 2c), Zr 3d (Figure 2d), and O 1s (Figure 2e) core levels, along with a survey scan (Figure S4a, Supporting Information), were obtained for the samples with different ZrO2/Co3O4 ratios. The binding energies of the Co and Zr signals did not greatly shift upon changing the component ratio. However, the intensity of the Co signals clearly changed as the amount of Co decreased. The Zr signals also changed, but not as much as the Co signals, which may be due to ZrO2 being on top of the Co3O4 surface. Additionally, the O 1s spectroscopic signals were affected. The O 1s peaks at ≈530 and 532 eV are ascribed to lattice oxygen and nonuniform surface sites, respectively.30 The peak at ≈532 eV is related to defect sites on the surface, such as chemisorbed or dissociated oxygen or hydroxyl species.31 The surface of Co3O4 is known to be readily covered with a monolayer of negatively charged chemisorbed oxygen.28 In this case, the peak at ≈532 eV in the O 1s spectrum of pure Co3O4, shown in Figure 2e, can be clearly deconvoluted to demonstrate the surface adsorption ability of Co3O4. In addition, after coprecipitation with ZrO2, the O 1s signals of all ZrO2/Co3O4 composites showed enhanced peaks at ≈532 eV, which can be observed more clearly in the overlapping plots in Figure S4b (Supporting Information). This result shows the surface electron accepting capabilities of ZrO2, as reported previously,24, 25 indicating its strong adsorption of CO3 2− electrolyte during the electrochemical oxidation of methane.

2.2. Electrochemical Performance for CH4 Oxidation

To measure the electrochemical performance of CH4 oxidation, a glassy carbon disc was employed to load the catalyst, forming the working electrode. Details of the preparation and measurement process are provided in the Experimental Section. As mentioned above, NiO is not an efficient catalyst for CH4 oxidation. Thus, we prepared a ZrO2/NiO composite and compared its catalytic ability with that of ZrO2/Co3O4, as shown in Figure S5 (Supporting Information). In the measurement of the current–voltage (JV) curves and Nyquist plots (for impedance analysis), argon (Ar) saturation was used to analyze the water oxidation and CH4 saturation was used to analyze the CH4 oxidation competing with water oxidation. Additional anodic activity under CH4 saturation indicated that both NiO and Co3O4 have CH4 oxidizing activity. However, CH4 oxidation performed better on the surface of Co3O4 than NiO. The same results were observed via electrochemical impedance spectroscopy (EIS) (Figure S5b,d, Supporting Information), which illustrated the superior activity of Co3O4 for CH4 electrochemical oxidation.

Figure 3 shows the linear sweep voltammetry (LSV) curves of the ZrO2/Co3O4 samples with different ratios in CH4‐saturated carbonate electrolyte. The 1–4 ZrO2/Co3O4 sample showed the highest electrochemical current density for CH4 oxidation, which agrees with the sample morphology results after adjusting the component ratio, as 1–4 ZrO2/Co3O4 had a relatively suitable size and amount of Co3O4. The current density of the 1–2 ZrO2/Co3O4 sample was relatively low even after increasing the potential to a high value, which illustrated that a low amount of Co3O4 provided poor catalytic activity in CH4 oxidation. By contrast, ZrO2 showed extremely weak oxidizing activity under both H2O and CH4 saturation due to its high bandgap of more than 5 eV.32 The 1–6 ZrO2/Co3O4 sample showed the same JV curve trend as the 1–4 ZrO2/Co3O4 sample but a lower current density value, illustrating the weaker surface adsorption ability toward the carbonate electrolyte due to the lower amount of ZrO2. In addition, JV curves of the optimized sample in Ar‐ and CH4‐saturated electrolyte were recorded to demonstrate its ability for CH4 oxidation, as shown in Figure S6a (Supporting Information). In addition, pure Co3O4 powder was also prepared to examine the electrochemical oxidation of CH4 (Figure S6b, Supporting Information). Unfortunately, the Co3O4 sample without ZrO2, which aids the adsorption of oxygen donors, showed no additional anodic activity in CH4‐saturated electrolyte and even worse performance, which confirmed that the outstanding electrochemical oxidation of CH4 resulted from the synergistic effects of ZrO2 and Co3O4, i.e., carbonate adsorption and CH4 oxidation, respectively.

Figure 3.

Figure 3

Electrochemical performance of CH4 oxidation. a) JV curves and b) magnified curves of the ZrO2/Co3O4 samples with different ratios of 1–2, 1–4, and 1–6.

2.3. CH4 Conversion Measurement and Product Analysis

In the measurement of the CH4 conversion, the catalyst was loaded on carbon paper to form the anode, carbonate was the electrolyte and platinum foil was the counter electrode. The preparation and measurement processes are shown in the schematic diagram in Figure S7 (Supporting Information), and further details are provided in the Experimental Section. CH4 conversion using the Co3O4 catalyst was performed to obtain more stable and useful higher alcohols. Therefore, a sealed reactor instrument is more suitable for the conversion and for tracking the CH4 oxidation progress in real time. To determine the optimal potential for long reaction times, the difference in the current density of the optimized sample in Ar‐ and CH4‐saturated electrolyte (see Figure S6a, Supporting Information) was calculated, and the curves are provided in Figure S8 (Supporting Information). Based on the curves, 2.0 V was selected as the suitable potential for CH4 electrochemical oxidation to obtain a relatively high current density and low competition with water oxidation. The products were collected after reaction for 3, 6, and 12 h with vigorous stirring. It curves for 12 h measurement was shown in Figure S9 (Supporting Information). To identify the products, proton nuclear magnetic resonance (1H‐NMR) spectroscopy was performed. Figure 4 a shows the 1H‐NMR spectrum of the products obtained after 12 h of reaction. The main products were 1‐propanol and 2‐propanol.33, 34 However, by‐products, such as methanol, ethanol, acetaldehyde, and acetone, were also observed. The typical 1H‐NMR peak of methanol is located at ≈3.3–3.5 ppm,34 which may overlap with that of 1‐propanol. The 1H‐NMR peaks of ethanol appear at the same positions as the 2‐propanol peaks.34 However, the subpeak numbers of ethanol and 2‐propanol are different; thus, the product can be identified as 2‐propanol. However, small ethanol peaks with weak peak intensity may be obscured by the 2‐propanol peaks. The small peak at ≈2.2 ppm can be ascribed to acetaldehyde and acetone.33 Compared with the main products (1‐propanol and 2‐propanol), all the by‐products were observed in negligible quantities. In order to eliminate other influence factors, control experiment was conducted at the same condition for 12 h long‐term reaction except the presence of CH4. The 1H‐NMR result of the products from control experiment was shown in Figure S10a (Supporting Information). In addition, the pure carbonate electrolyte before reaction was also detected with 1H‐NMR spectrum for comparison (Figure S10b, Supporting Information). To confirm the amount of CH4 consumed and products generated, gas chromatography (GC) and GC‐mass spectrometry (MS) were performed. The amount of CH4 remaining in the reactor after 3, 6, and 12 h of reaction is shown in Figure 4b. In addition, the reference line of the amount of CH4, measured by GC, is shown in Figure S11 (Supporting Information). CH4 gas was mostly consumed, and the amount decreased gradually with the reaction time. After 12 h of reaction, almost 40% of the CH4 gas was converted. Meanwhile, the amount of various products measured by the GC‐MS system is shown in Table 1 .

Figure 4.

Figure 4

Product analysis and production efficiency. a) 1H‐NMR spectrum of the products after 12 h. b) The amount of CH4 remaining after electrochemical oxidation. c) Production efficiencies of the products of 1‐propanol, 2‐propanol, and acetaldehyde at different reaction times.

Table 1.

Concentrations of various products after the electrochemical oxidation of CH4 for 3, 6, and 12 h

Time [h] Methanol [μg mL−1] Formaldehyde [μg m L−1] Ethanol [μg mL−1] Acetaldehyde [μg mL−1] 1‐Propanol [μg mL−1] 2‐Propanol [μg mL−1] Acetone [μg mL−1]
3 29.95 0.88 0 261.50 0.50 19.53 0
6 33.15 1.11 0.49 170.34 56.51 101.74 3.82
12 33.71 1.10 35.21 153.42 1336.12 1315.56 14.16

As described in Table 1, seven products were detected: methanol, formaldehyde, ethanol, acetaldehyde, 1‐propanol, 2‐propanol, and acetone. The amount of methanol and formaldehyde (products containing one carbon atom) did not greatly change with reaction time, illustrating a balance between generation and consumption. Thus, methanol and formaldehyde are the primary products of CH4 oxidization. As detailed in a previous study, formaldehyde should be the methanol oxidation product.23 After comparing the amount of ethanol and acetaldehyde (products containing two carbon atoms), acetaldehyde can be confirmed as the main product from the addition reaction of CH4 and formaldehyde. Moreover, the amount of acetaldehyde decreased with the reaction time, which illustrates that acetaldehyde plays a pivotal role in the production of 1‐propanol and 2‐propanol, indicating that 1‐propanol and 2‐propanol were converted from acetaldehyde. Then, after 12 h of reaction, 1‐propanol, and 2‐propanol were the main stable products of CH4 oxidation, which agrees with the 1H‐NMR results. The conversion efficiencies for acetaldehyde, 1‐propanol and 2‐propanol were calculated and are shown in Figure 4c. After 12 h of reaction, the main products, 1‐propanol and 2‐propanol, showed total production efficiency of over 60%.

2.4. Reaction Mechanism Analysis

As observed in the results in Table 1, acetaldehyde was the key product. The reactions involved in the formation of acetaldehyde are shown below, according to previous investigations35

CH4oxidantCH3OHoxidantHCHO (3)
CH4+CH3OHoxidantCH3CH2OHoxidantCH3CHO (4)
CH4+HCHOoxidantCH3CHO (5)
CH3OH+HCHOdehydrationCH3CHO (6)

In the primary reaction involved in CH4 oxidation, CH4 was oxidized by the oxidant (carbonate in this work) to form CH3OH, which was subsequently oxidized to HCHO. The production mechanism has been investigated in detail, and there are several ways to achieve the reaction.36 After that, several reactions can occur to generate acetaldehyde by employing the reactants CH4, methanol, and formaldehyde. Thus, the generation and accumulation of acetaldehyde is rapid and large, which is in accordance with the results in Table 1. For comprehensive understanding of the reaction processes, theoretical potentials of several oxidation reactions related to the methane conversion were listed in Table S2 (Supporting Information).

The production of 1‐propanol and 2‐propanol from acetaldehyde is the most important reaction step in CH4 conversion. The mechanism involves a type of addition reaction, as shown in Figure 5 . The formation of 2‐propanol is common and has been reported in previous work.23 As shown in Figure 5a, the methyl group in CH4 acts as a nucleophilic reagent and attacks the carbonyl carbon in acetaldehyde. Then, a nucleophilic addition reaction occurs to form 2‐propanol as one of the main products in this CH4 conversion reaction. However, considering the addition reaction mechanism, the formation of 1‐propanol from acetaldehyde and CH4 is impossible. Therefore, we considered all the reaction conditions to determine a possible route for 1‐propanol production and found that the free radical addition reaction is suitable (Figure 5b1–b3). First, in route b1, a methyl radical is generated from CH4 with the participation of Co3O4 and carbonate. A carbonate radical is generated through anodic oxidation with the help of Co3O4 due to the relatively low generation energy compared with that of the hydroxyl radical,37 which can be obtained from electrochemical oxidation processes.38 The carbonate radical acted as an intermediate to generate a methyl radical through reaction with CH4. At the same time, in route b2, the as‐produced acetaldehyde equilibrates between its isomers, acetaldehyde, and vinyl alcohol. The vinyl alcohol configuration has a higher energy of 45 kJ mol−1 than the acetaldehyde form but is reachable in the presence of carbonate.39, 40 In the normal electrophilic addition to alkenes, the products follow the Markovnikov rule,41 illustrating 2‐propanol as the main product when CH4 reacts with vinyl alcohol. However, when the addition reaction is conducted through the free radical route, the products follow the anti‐Markovnikov rule, thereby producing 1‐propanol as the main product, as shown in route b3. When a methyl radical attacks carbon 1, the resulting 2‐propanol radical (free electron on carbon 2) is not the most stable state. However, when the methyl radical attacks carbon 2, the resulting 1‐propanol radical (free electron on carbon 1) is more stable than the 2‐propabol radical, illustrating 1‐propanol as the main product. 2‐Propanol can be directly converted from acetaldehyde and CH4 through a nucleophilic addition reaction, leading to more 2‐propanol being produced than 1‐propanol at short oxidation times. However, after long reaction times, the amount of 1‐propanol exceeds that of 2‐propanol, even though 2‐propanol is more thermodynamically stable, illustrating the unique regional selectivity of 1‐propanol production through radical addition with the participation of a Co3O4 catalyst and carbonate electrolyte. In summary, after comprehensive analysis with the above content, the complete reaction pathways for electrochemical oxidation of methane with ZrO2/Co3O4 anode and carbonate electrolyte were proposed, shown in Figure S12 (Supporting Information). The reaction process may help other researchers to have an overall understanding of the one carbon related reactions.

Figure 5.

Figure 5

Reaction mechanism analysis. a) Nucleophilic addition reaction of methane and acetaldehyde to form 2‐propanol. b1–b3) Free radical addition reaction of methane and acetaldehyde to form 1‐propoanl.

3. Conclusion

In summary, we designed a ZrO2/Co3O4 nanocomposite that aids in the regional selective oxidation of CH4 to 1‐propanol and 2‐propanol via an electrochemical method. We expected stable products, such as higher alcohols, to be formed due to the strong surface adsorption ability of Co3O4 and the participation of carbonate, which has a soft oxidizing ability, delivered by ZrO2. After tracking the products for different reaction times, acetaldehyde was found to be the key intermediate. To convert acetaldehyde to 1‐propanol with the participation of CH4, a free radical addition reaction was conducted by an electrochemical reaction, with Co3O4 as the catalyst and carbonate as the electrolyte. Finally, as a result of competition between reactions, both 1‐propanol and 2‐propanol were the main products. This electrochemical partial oxidation of CH4 may aid in the synthesis of other oxygenates and long‐chain hydrocarbons.

4. Experimental Section

Fabrication of Electrocatalyst Materials: All reagents were used as received without further treatment. The ZrO2/Co3O4 nanocomposite was synthesized using precipitation and a hydrothermal method. For the 1–2 ZrO2/Co3O4 sample, 0.1611 g ZrOCl2·8H2O (99.0%, Junsei, Japan), 0.291 g Co(NO3)2·6H2O (98%, Aldrich, US), and 9.6 g NaOH (96%, Samchun, Korea) were dissolved in 40 mL deionized (DI) water with vigorous stirring for 30 min. Then, the solution was transferred to a 60 mL autoclave container and heated at 180 °C for 24 h. After that, the powder was collected by centrifuge and washed with DI water 3 times. Finally, the 1–2 ZrO2/Co3O4 sample was obtained after thermal annealing at 500 °C for 3 h. For the ZrO2/Co3O4 nanocomposites with different ratios, the amount of Co(NO3)2·6H2O was adjusted to 0.582 g for 1–4 ZrO2/Co3O4 and 0.873 g for 1–6 ZrO2/Co3O4, with no changes in the other conditions. For comparison, pure Co3O4 and ZrO2/NiO samples were prepared by the same method as 1–4 ZrO2/Co3O4, without the addition of ZrOCl2·8H2O and with the addition of 0.582 g Ni(NO3)2·6H2O (97%, Aldrich, US) instead of Co(NO3)2·6H2O, respectively.

Electrochemical Test: LSV and EIS tests for the comparison of ZrO2/Co3O4 and ZrO2/NiO samples were conducted in a three‐electrode system using a potentiostat (CH Instrument, CHI 660) with a glassy carbon electrode as the working electrode, Ag/AgCl as the reference electrode, a Pt foil as the counter electrode and 0.5 m Na2CO3 solution as the electrolyte. Meanwhile, LSV tests for ZrO2/Co3O4 samples with different component ratio were conducted in a two‐electrode system without the utilization of Ag/AgCl reference electrode. All working electrodes were prepared by dispersing the powder samples in DI water in a concentration of 3 mg mL−1 with vigorous stirring for 30 min and then dropping 20 µL of the dispersed solution on the surface of the glassy carbon electrode (area = 0.07 cm2), followed by drying at room temperature. Next, 10 µL of a 5% Nafion 117 solution (Aldrich) was deposited on the surface of the glassy carbon electrode to cover the sample films, followed by drying at room temperature. Ultrahigh‐purity argon gas (Ar, 99.999%) and methane gas (CH4, 99.999%) were used. Before each electrochemical test, the electrolyte was bubbled with Ar or CH4 for 1 h to prepare the Ar‐ or CH4‐saturated electrolyte.

CH4 Conversion Test: The long‐term electrochemical oxidation of CH4 was conducted in a two‐electrode system with a closed reaction instrument, employing carbon paper (Alfa) as the working electrode, Pt foil as the counter electrode and 30 mL 0.5 m Na2CO3 solution (pH around 12.0 before reaction and about 11.9 after 12 h reaction) as the electrolyte. The working electrode was prepared by dispersing the powder sample in DI water in a concentration of 3 mg mL−1 with vigorous stirring for 30 min and then dropping 5.7 mL of the dispersed solution on the surface of the carbon paper (area = 20 cm2), followed by drying at room temperature. Next, 3 mL of a 5% Nafion 117 solution was deposited to cover the sample film on the carbon paper, followed by drying at room temperature. Before the electrochemical reaction, the electrolyte was bubbled with CH4 for 1.5 h to remove the oxygen and fill the space in the reaction instrument. In this case, after the consumption of saturated CH4 in aqueous solution, the gas phase CH4 could dissolve in the electrolyte continually, guaranteeing the adequate reactant. Electrochemical oxidation was conducted at 2.0 V versus Pt for 3, 6, or 12 h.

Characterization and Products Analysis: Morphology analyses of the samples were carried out using field‐emission scanning electron microscopy (FESEM, JSM‐7000F, Japan) and a JEOL JEM‐2100F (Japan) electron microscope. EDS spectra and ICP‐OES measurements were employed for the elements ratio detection. XRD measurements were conducted using a Siemens diffractometer D500/5000 in a Bragg–Brentano geometry. XPS data were obtained from a K‐alpha instrument (Thermo Scientific Inc., UK). 1H‐NMR was conducted using an Avance III HD 400 FT‐NMR instrument (Bruker Biospin), where the sample was prepared by mixing 0.4 mL of the product solution with 0.2 mL D2O. The amount of methane was determined using a 7890B GC instrument (Agilent Technologies). The product was examined using a 7890B‐5977A GC‐MS instrument (Agilent Technologies).

Conflict of Interest

The authors declare no conflict of interest.

Supporting information

Supplementary

Acknowledgements

This work was supported by the NRF of Korea Grant funded by the Ministry of Science, ICT, and Future Planning (Nos. 2016R1A2A1A05005216, 2015M1A2A2074663, and 2016M3D3A1A01913117 (C1 Gas Refinery Program)).

Ma M., Jin B. J., Li P., Jung M. S., Kim J. I., Cho Y., Kim S., Moon J. H., Park J. H., Adv. Sci. 2017, 4, 1700379 https://doi.org/10.1002/advs.408

References

  • 1. Key World Energy Statistics 2015. International Energy Agency, Paris. [Google Scholar]
  • 2. Carbon Dioxide Emissions Coefficients by Fuel. U.S. Energy Information Administration, 2016. [Google Scholar]
  • 3. Kerr R. A., Science 2010, 328, 1624. [DOI] [PubMed] [Google Scholar]
  • 4. Brandt A. R., Heath G. A., Kort E. A., Sullivan F. O., Petron G., Jordaan S. M., Tans P., Wilcox J., Gopstein A. M., Arent D., Wofsy S., Brown N. J., Bradley R., Stucky G. D., Eardley D., Harriss R., Science 2014, 343, 733. [DOI] [PubMed] [Google Scholar]
  • 5. Shindell D. T., Faluvegi G., Koch D. M., Schmidt G. A., Unger N., Bauer S. E., Science 2009, 326, 716. [DOI] [PubMed] [Google Scholar]
  • 6. Farrauto R. J., Science 2012, 337, 659. [DOI] [PubMed] [Google Scholar]
  • 7. Alvarez R. A., Pacala S. W., Winebrake J. J., Chameides W. L., Hamburg S. P., Proc. Natl. Acad. Sci. USA 2012, 109, 6435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Chen X., Li Y., Pan X., Cortie D., Huang X., Yi Z., Nat. Commun. 2016, 7, 12273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Cargnello M., Delgado Jaen J. J., Hernandez Garrido J. C., Bakhmutsky K., Montini T., Calvino Gamez J. J., Gorte R. J., Fornasiero P., Science 2012, 337, 713. [DOI] [PubMed] [Google Scholar]
  • 10. Tao F. F., Shan J., Nguyen L., Wang Z., Zhang S., Zhang L., Wu Z., Huang W., Zeng S., Hu P., Nat. Commun. 2015, 6, 7798. [DOI] [PubMed] [Google Scholar]
  • 11. Yuliati L., Yoshida H., Chem. Soc. Rev. 2008, 37, 1592. [DOI] [PubMed] [Google Scholar]
  • 12. He Z., Qian Q., Ma J., Meng Q., Zhou H., Song J., Liu Z., Han B., Angew. Chem., Int. Ed. 2016, 55, 737. [DOI] [PubMed] [Google Scholar]
  • 13. Wang V. C.‐C., Maji S., Chen P. P.‐Y., Lee H. K., Yu S. S.‐F., Chan S. I., Chem. Rev. 2017, 117, 8574. [DOI] [PubMed] [Google Scholar]
  • 14. Tang P., Zhu Q., Wu Z., Ma D., Energy Environ. Sci. 2014, 7, 2580. [Google Scholar]
  • 15. Nazaries L., Murrell J. C., Millard P., Baggs L., Singh B. K., Environ. Microbiol. 2013, 15, 2395. [DOI] [PubMed] [Google Scholar]
  • 16. Olah G. A., Goeppert A., Czaun M., Prakash G. K. S., J. Am. Chem. Soc. 2013, 135, 648. [DOI] [PubMed] [Google Scholar]
  • 17. Petrushina I. M., Bandur V. A., Bjerrum N. J., Cappeln F., Qingfeng L., J. Electrochem. Soc. 2002, 149, D143. [Google Scholar]
  • 18. Eng D., Stoukides M., Catal. Rev. Sci. Eng. 1991, 33, 375. [Google Scholar]
  • 19. Promoppatum P., Viswanathan V., ACS Sustainable Chem. Eng. 2016, 4, 1736. [Google Scholar]
  • 20. Amenomiya Y., Birss V., Goledzinowski M., Galuszka J., Sanger A. R., Catal. Rev. Sci. Eng. 1990, 32, 163. [Google Scholar]
  • 21. Spinner N., Vega J. A., Mustain W. E., Catal. Sci. Technol. 2012, 2, 19. [Google Scholar]
  • 22. Spinner N., Mustain W. E., Electrochim. Acta 2011, 56, 5656. [Google Scholar]
  • 23. Spinner N., Mustain W. E., J. Electrochem. Soc. 2012, 159, E187. [Google Scholar]
  • 24. Ahn H., Marks T. J., J. Am. Chem. Soc. 1998, 120, 13533. [Google Scholar]
  • 25. Samaranch B., Piscina P., Glet G., Houalla M., Gelin P., Homs N., Chem. Mater. 2007, 19, 1445. [Google Scholar]
  • 26. Spinner N., Mustain W. E., J. Electrochem. Soc. 2013, 160, F1275. [Google Scholar]
  • 27. Zhu Z., Lu G., Zhang Z., Guo Y., Guo Y., Wang Y., ACS Catal. 2013, 3, 1154. [Google Scholar]
  • 28. Kim J. Y., Choi N., Park H. J., Kim J., Lee D., Song H., J. Phys. Chem. C 2014, 118, 25994. [Google Scholar]
  • 29. Peterson A., Abild‐Pedersen F., Studt F., Rossmeisl J., Norskov J., Energy Environ. Sci. 2010, 3, 1311. [Google Scholar]
  • 30. Li T., He J., Pena B., Berlinguette C. P., Angew. Chem., Int. Ed. 2016, 55, 1769. [DOI] [PubMed] [Google Scholar]
  • 31. Huang M., Zhang Y., Li F., Wang Z., Alamusi, N. Hu , Wen Z., Liu Q., Sci. Rep. 2014, 4, 4518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Schattka J., Shchukin D., Jia J., Antonietti M., Caruso R., Chem. Mater. 2002, 14, 5103. [Google Scholar]
  • 33. Carey F. A., Organic Chemistry, 4th Ed., McGraw‐Hill Higher Education, Boston, MA, USA 2000, Ch. 13. [Google Scholar]
  • 34.Biological Magnetic Resonance Data Bank, The Board of Regents of the University of Wisconsin System.
  • 35. Bar‐Nahum I., Khenkin A. M., Neumann R., J. Am. Chem. Soc. 2004, 126, 10236. [DOI] [PubMed] [Google Scholar]
  • 36. Iwasita T., Electrochim. Acta 2002, 47, 3663. [Google Scholar]
  • 37. Medinas D. B., Cerchiaro G., Trindade D. F., Augusto O., IUBMB Life 2007, 59, 255. [DOI] [PubMed] [Google Scholar]
  • 38. Kesselman J. M., Weres O., Lewis N. S., Hoffmann M. R., J. Phys. Chem. B 1997, 101, 2637. [Google Scholar]
  • 39. Bouma W. J., Radom L., Rodwell W. R., Theor. Chim. Acta 1980, 56, 149. [Google Scholar]
  • 40. Karton A., Chem. Phys. Lett. 2014, 592, 330. [Google Scholar]
  • 41. Jones G., J. Chem. Educ. 1961, 38, 297. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary


Articles from Advanced Science are provided here courtesy of Wiley

RESOURCES