Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2020 Aug 6;8(32):12151–12160. doi: 10.1021/acssuschemeng.0c03606

Dehydrogenative Coupling of Methanol for the Gas-Phase, One-Step Synthesis of Dimethoxymethane over Supported Copper Catalysts

Anh The To , Trenton J Wilke , Eric Nelson , Connor P Nash , Andrew Bartling , Evan C Wegener , Kinga A Unocic §, Susan E Habas , Thomas D Foust †,*, Daniel A Ruddy †,*
PMCID: PMC10906941  PMID: 38435970

Abstract

graphic file with name sc0c03606_0007.jpg

Oxymethylene dimethyl ethers (OMEs), CH3-(OCH2)n-OCH3, n = 1–5, possess attractive low-soot diesel fuel properties. Methanol is a key precursor in the production of OMEs, providing an opportunity to incorporate renewable carbon sources via gasification and methanol synthesis. The costly production of anhydrous formaldehyde in the typical process limits this option. In contrast, the direct production of OMEs via a dehydrogenative coupling (DHC) reaction, where formaldehyde is produced and consumed in a single reactor, may address this limitation. We report the gas-phase DHC reaction of methanol to dimethoxymethane (DMM), the simplest OME, with n = 1, over bifunctional metal–acid catalysts based on Cu. A Cu-zirconia-alumina (Cu/ZrAlO) catalyst achieved 40% of the DMM equilibrium-limited yield under remarkably mild conditions (200 °C, 1.7 atm). The performance of the Cu/ZrAlO catalyst was attributed to metallic Cu nanoparticles that enable dehydrogenation and a distribution of acid strengths on the ZrAlO support, which reduced the selectivity to dimethyl ether compared to a that obtained with a Cu/Al2O3 catalyst. The DMM formation rate of 6.1 h–1 compares favorably against well-studied oxidative DHC approaches over non-noble, mixed-metal oxide catalysts. The results reported here set the foundation for further development of the DHC route to OME production, rather than oxidative approaches.

Keywords: Oxymethylene dimethyl ethers, OMEs, Dimethoxymethane, Methanol, Dehydrogenative coupling, Supported copper catalysts

Short abstract

Bifunctional metal−acid catalysts based on Cu set the foundation for the single-step conversion of renewable methanol to low-soot diesel fuel.

Introduction

Diesel fuel for compression ignition (CI) engines is essential to the transportation sector, especially in mid- and heavy-duty applications, where it is advantageous over gasoline for spark ignition (SI) engines due to higher torque output and higher fuel efficiency of CI engines over SI engines.1,2 However, particulate matter emissions (colloquially referred to as soot) are a significant issue for CI engines. Oxygenated fuel compounds, such as dimethyl ether (DME), have demonstrated suitable diesel fuel properties and reduced sooting, leading to an ASTM International specification.3 However, handling gaseous DME and implementing the necessary engine modifications complicate its widespread adoption into the current infrastructure. On the other hand, oxymethylene dimethyl ethers (OMEs)—oligomeric structures of the formula CH3-(OCH2)n-OCH3, with n = 1–5—are room-temperature liquids with low vapor pressures that have received increased attention for diesel fuel applications. In addition to their excellent cetane numbers, exceeding that of conventional diesel, OMEs have demonstrated significant soot-reducing performance both as a standalone fuel and when blended into hydrocarbon diesel (e.g., yield sooting index (YSI) of 7–11 for dimethoxymethane (DMM) compared to 215 for conventional diesel fuel).1,49

The most common production route to OMEs is through a multi-step synthesis involving the acid-catalyzed acetalization reaction of methanol with formaldehyde.4,10,11 Anhydrous formaldehyde is preferred in this scheme to achieve high equilibrium conversion (i.e., water shifts the equilibrium away from OME products), but this is a costly step within the typical industrial oxidative dehydrogenation of methanol, where one equivalent of water is produced with each formaldehyde molecule.12,13 The use of methanol as the key precursor for OME production provides an attractive opportunity to utilize renewable and waste carbon sources (e.g., biomass, biogas, municipal solid waste), since gasification and methanol synthesis are established technologies.14,15 However, the costly production of formaldehyde, including carbon loss to DME, CO, and CO2 and the drying step, hinders the use of these alternative resources. Thus, routes that circumvent the carbon loss and drying cost associated with anhydrous formaldehyde production are needed to realize sustainable and cost-effective production of OMEs.

To explore new routes for OME production, the reaction of methanol to DMM, the simplest OME having n = 1, serves as a model reaction for catalyst development. The direct production of DMM from methanol over bifunctional catalysts, mostly via an oxidative coupling reaction, was recently reviewed.2 Re- and Ru-based catalysts, heteropolyacids, and various mixed oxide catalysts (e.g., V2O5–TiO2) have demonstrated activity for this transformation through the oxidative dehydrogenation of methanol to formaldehyde with subsequent acetalization of formaldehyde with methanol to yield DMM (Scheme 1a).1619 The reaction employs molecular oxygen as the oxidant, formaldehyde is generated in situ but not isolated, and water is produced as a byproduct. Sun et al.2 used DMM formation rate (i.e., molar flow rate of DMM produced [molC/h]/mol of metal on catalyst [mol]) to compare catalytic performance across various reports using different catalysts under different reaction conditions, and this analysis demonstrated that the majority of studies with non-noble, mixed-metal oxide catalysts had DMM formation rates in the range of 0.4–6.6 h–1. The most active catalysts were either supported noble metals (e.g., RuO2/carbon nanotube) or commercial iron-molybdate catalysts under methanol-rich feed conditions, having DMM formation rates of 23–26 h–1. However, the redox activity of Ru catalysts results in the formation of methyl formate (MF) with selectivity in the range of 35–70%. Hence, production of DMM using Ru catalysts would require costly and energy-intensive separation.2

Scheme 1. Reaction of Methanol to DMM in a Single Step via (a) Oxidative Dehydrogenation and (b) Non-oxidative Dehydrogenation.

Scheme 1

Integrated processes to produce hydrocarbon fuels from renewable biomass are often limited in hydrogen content by the feedstock composition and require hydrogen to generate the final product (i.e., biomass molecular formula of CnH1.5nO0.66n and fuel molecular formula of CnH2n+2).15,20 The oxygenated OME fuel product discussed here limits the hydrogen requirement versus hydrocarbons, but the overall H2 demand remains an important cost factor within a conceptual market-responsive biorefinery, where a variety of fuels and chemicals are produced to meet market demand. For example, imported H2 costs can range from less than $2/kg for relatively inexpensive H2 from steam methane reforming of natural gas to more than $4/kg for renewably sourced H2.21,22 Consequently, if H2 is not imported for the processes, the ability to recycle H2 within the biorefinery would provide a unique advantage that could reduce overall production costs. In contrast to oxidative dehydrogenation routes where hydrogen is removed as water (Scheme 1a), a dehydrogenative coupling (DHC) route to OMEs offers this H2-recycle option. As shown in Scheme 1b, one equivalent of water is still produced from the acetalization of methanol and formaldehyde, and gaseous H2 is produced as a byproduct. However, DHC approaches have been much less studied, and are limited to just a few reports for the liquid-phase conversion and no reports for the gas-phase reaction.2 This lack of investigation into the gas-phase reaction may be due to the difference in reaction thermodynamics (Scheme 1) or the postulated incompatibility of traditional methanol dehydrogenation at high temperatures (500–800 °C) with the acetalization reaction at low temperatures (<300 °C).2

We envisioned that this DHC reaction could be promoted over bifunctional metal–acid catalysts at low temperatures, similar to the oxidative pathways, where methanol dehydrogenation occurs at metallic sites with subsequent acetalization at acidic sites. For this initial investigation, Cu was chosen for the metallic functionality due to its known dehydrogenation activity and ability to be supported on a variety of acidic supports.13,2326 In addition, Cu catalysts do not suffer from the Mo migration/sublimation issue with iron-molybdate oxidative dehydrogenation catalysts.2 Here we report the gas-phase DHC reaction of methanol to DMM over bifunctional metal–acid catalysts under the remarkably mild conditions of 200 °C and 1.7 atm. A Cu/ZrAlO catalyst demonstrated the highest activity for DMM synthesis among the four catalysts investigated, exhibiting methanol conversion of 25% with DMM selectivity of 12%. This Cu/ZrAlO catalyst was approximately 3-fold more active than Cu/Al2O3 (the second most active catalyst), achieving 40% of the DMM equilibrium-limited single-pass yield (3.0%). This greater DMM yield is attributed to decreased DME formation at the weaker acid sites of Cu/ZrAlO, as supported by pyridine-adsorption DRIFTS analysis. The DMM formation rate of 6.1 h–1 was achieved over Cu/ZrAlO, comparable to high-performing non-noble metal oxide catalysts in the direct DMM synthesis via methanol oxidation.2 This report sets the foundation for catalyst development and reaction engineering to further decrease DME formation and increase DMM selectivity via the DHC reaction.

Results and Discussions

Synthesis and Characterization of Supported Cu Catalysts

The following support materials were used to provide acidity for the acetalization step: commercial acidic supports SiO2–Al2O3 (SiAlO) and Al2O3; a non-commercial ZrOx–Al2O3 (ZrAlO) mixed oxide, comparable to a reported catalyst that exhibited high activity for OME production from methanol and formaldehyde;27 and a commercial ZrO2 for comparison to the mixed oxide. Preliminary experiments employing Cu on SiO2 for comparison to the SiAlO support demonstrated no activity, attributed to low acidity of SiO2, and were not considered for further study. Copper was supported via incipient wetness impregnation, targeting 3 wt% Cu, and elemental analysis confirmed loadings of 2.8–3.1 wt%. The Cu loadings, surface areas, acid site densities, and Brønsted:Lewis (B:L) acid site ratios are presented in Table 1. Additional physisorption data are included in Figure S1.

Table 1. Cu Loading, Surface Areas, Total Acid Site Density, and Brønsted:Lewis (B:L) Acid Site Ratios for Supported Cu Catalysts.

catalyst Cu loading (wt%) surface area (m2/g) acid site density (μmol/gcat) B:L ratio (mol/mol)
Cu/Al2O3 3.1 120 650 0.46
Cu/SiAlO 3.1 300 1540 0.40
Cu/ZrO2 2.8 120 640 0
Cu/ZrAlO 3.0 110 690 0.39

Total acid site density was measured using temperature-programmed desorption of ammonia (Figure S2). The greater total acid site density exhibited by the Cu/SiAlO material (1540 μmol/g) is likely due to its higher surface area (300 m2/g vs 120 m2/g), whereas the other catalysts had 2-fold lower acid site densities (640–690 μmol/g). The molar ratio of B:L acid sites was measured using pyridine-adsorption DRIFTS (py-DRIFTS) after reduction with H2. The Cu/ZrO2 catalyst did not contain Brønsted sites (100% Lewis sites), and the other catalysts were predominantly Lewis-acidic, having B:L mol ratios of ca. 0.4, corresponding to ca. 70% of the total sites being Lewis acid sites. The peak position of chemisorbed pyridine near 1450 cm–1 is indicative of Lewis acid strength, where higher values correspond to stronger acid strength and vice versa.2831 Relating changes to the 1450 cm–1 peak position with Lewis acid strength is analogous to relating peak positions in the 1600–1650 cm–1 range to Lewis acid strength, as recently demonstrated.31 The peak positions trended from strongest to weakest, Cu/SiAlO (1452 cm–1) ∼ Cu/Al2O3 (1452 cm–1) > Cu/ZrAlO (1447 cm–1) > Cu/ZrO2 (1443 cm–1) (Figure 1). The peak position observed for the Cu/ZrAlO material is indicative of a distribution of acid strengths associated with stronger Al2O3-based acid sites and weaker ZrOx-based acid sites.

Figure 1.

Figure 1

Pyridine-DRIFTS spectra for the supported Cu catalysts. The dashed lines from left to right highlight the peak centers at 1452 (Cu/Al2O3, Cu/SiAlO), 1447 (Cu/ZrAlO), and 1443 (Cu/ZrO2) cm–1.

A variety of techniques were employed to characterize the Cu speciation on these materials. XRD patterns of the oxidized catalysts did not exhibit reflections attributed to crystalline Cu oxides (Figure S3). These data suggest amorphous CuOx species or small nanoparticles (<3 nm), such that broad, low-intensity reflections from the low loading of Cu species could not be detected, especially in the presence of intense reflections due to the crystalline oxide supports, Al2O3 and ZrO2. Previous reports have established correlations between Cu speciation and the observed reduction temperature during temperature-programmed reduction (TPR) with H2. The TPR profiles of the Cu materials investigated here are presented in Figure S4. Cu/SiAlO exhibited two reduction peaks similar to that observed previously with an acidic silica–alumina support,32 where the low temperature peak (at ca. 300 °C) represented reduction of oxo-cation-like [Cu-O-Cu]2+ species, and the high temperature peak (at ca. 460 °C) corresponded to small and well-dispersed CuOx species being reduced to Cu0. Cu/Al2O3 exhibited a reduction peak at ca. 230 °C, similar to a previous report that attributed this to the reduction of CuOx clusters to Cu0, and a broad reduction event up to 500 °C attributed to copper-aluminate-like species.33,34 Cu species on the reducible ZrO2 support reduced at a lower temperature of ca. 185 °C with a small shoulder at ca. 175 °C. Similar low-temperature reduction events were attributed to well-dispersed CuOx species being reduced to Cu0.35,36 Addition of reducible ZrOx species to Al2O3 at a low loading (4.79 wt% Zr) slightly lowered the reduction temperature of CuOx species, exemplified by a peak shift to ca. 220 °C on the Cu/ZrAlO catalyst comparing to ca. 230 °C on Cu/Al2O3. In addition, there was a smaller contribution from copper-aluminate-like species, indicated by the decrease of the broad reduction event near 500 °C. Based on these TPR data, in situ reduction of the catalysts on the XRD stage at 300 °C with flowing 5% H2 was investigated, but no crystalline metallic Cu or Cu-oxide species were observed (Figure S3). Again, this is attributed to small particle sizes and the low Cu loading on these materials. When the Cu/SiAlO material was reduced at 500 °C, characteristic reflections for large metallic Cu particles were observed, consistent with TPR data of a high-temperature reduction event. Further, large Cu particles were observed in STEM images for Cu/SiAlO catalyst reduced at 450 °C as discussed below (Figure S5).

DRUV–vis–NIR spectroscopy was employed to investigate Cu speciation before and after reduction, since characteristic electronic absorption spectra are well-known for Cu oxides versus metallic Cu nanoparticles.3740 All of the oxidized catalysts exhibited a broad absorption in their DRUV–vis–NIR spectra centered near 800 nm and extending into the NIR, and a high-energy absorption in the UV region (Figure 2). The band at 800 nm has been attributed to both d-d transitions of octahedral Cu2+ ions and to a blue-shifted (i.e., shorter wavelength) absorption for nano-CuO versus the bulk CuO transition at ca. 870 nm, and the high-energy absorption was assigned to oxygen-to-metal charge-transfer bands in Cu oxides and clusters.37,39,41 Consistent with the TPR data, these features suggest Cu-oxo clusters and/or CuO nanoparticles prior to reduction. After in situ reduction at 300 °C, the spectra for all materials except Cu/SiAlO were dominated by an absorption centered between 565 and 615 nm (Figure 2B–D). This feature is characteristic for the local surface plasmon resonance (LSPR) of Cu nanoparticles.38,40 For Cu/SiAlO, this feature was observed after reduction at 450 °C, consistent with the TPR data (Figure 2A). The position of the LSPR is not strongly affected by Cu nanoparticle size, but it is known to be sensitive to the local electronic environment of the Cu nanoparticles.38,42 The LSPR peak was observed at 565–570 nm on the irreducible oxides, Cu/SiAlO and Cu/Al2O3, but was red-shifted to lower energy (610 nm) on the reducible oxide, Cu/ZrO2. The Cu/ZrAlO material exhibited both of these LSPR peak positions (Figure 2D), suggesting the presence of Cu nanoparticles on alumina-rich and zirconia-rich regions of the support.

Figure 2.

Figure 2

DRUV–vis–NIR spectra after oxidation at 450 °C in air and reduction at 300 °C in H2 for (A) Cu/SiAlO also reduced at 450 °C, (B) Cu/Al2O3, (C) Cu/ZrO2, and (D) Cu/ZrAlO.

X-ray absorption spectroscopy (XAS) was used to further characterize the Cu species. The small pre-edge peak observed at 8977.6 eV in the Cu K-edge XANES spectrum of the oxidized samples (Figure S6) indicates the presence of Cu2+ species prior to reduction, consistent with the assignment by DRUV–vis–NIR spectroscopy. The EXAFS analyses of the samples (Figure 3) show a single first-shell peak at ca. 1.50 Å (phase uncorrected distance) which is characteristic of Cu–O scattering. The fitted coordination numbers and bond distances (ca. 4 Cu–O bonds at 1.94 Å, Table S1) are similar to those of CuO;43 however, the lack of a strong second-shell Cu–Cu scattering peak (i.e., Cu–O–Cu) suggests the Cu species are highly dispersed or form small clusters, which is in agreement with the TPR and XRD results. Following reduction at 300 °C the EXAFS spectra for Cu/Al2O3, Cu/ZrO2, and Cu/ZrAlO exhibit contributions from both Cu–O (1.50 Å) and Cu–Cu (2.15 Å) scattering. The Cu–Cu bond distances of the three samples were determined to be 2.52 Å (Table S1), slightly shorter than bulk FCC Cu and indicative of the formation of metallic nanoparticles.44 The metal scattering feature was absent from the spectrum of Cu/SiAlO following reduction at 300 °C, but appeared after treatment at 450 °C in agreement with the TPR and DRUV–vis–NIR spectroscopy results.

Figure 3.

Figure 3

Magnitude of the Fourier transform of the Cu K edge k2-weighted EXAFS after oxidation at 150 °C in air and reduction at 300 °C in H2 for (A) Cu/SiAlO also reduced at 450 °C, (B) Cu/Al2O3, (C) Cu/ZrO2, and (D) Cu/ZrAlO.

High-angle annular dark field (HAADF) STEM imaging with energy-dispersive X-ray spectroscopy (EDS) was used to investigate the distribution of Cu in the reduced and passivated catalysts. Clustering of Cu was not readily observed in the reduced and passivated Cu/Al2O3, Cu/ZrO2, and Cu/ZrAlO catalysts. However, the Al-containing materials were sensitive to the electron beam and therefore were not suitable for EDS analysis at the required beam current, and the high contrast of Zr limited the ability to distinguish between oxidized Cu in the passivated samples and the Zr-containing supports by HAADF-STEM. The absence of distinct Cu clustering in these catalysts suggests that the metallic Cu particles observed by DRUV–vis–NIR spectroscopy and XAS are small, which is in agreement with the absence of crystalline metallic Cu reflections in the XRD patterns. STEM analysis of the Cu/SiAlO catalyst, on the other hand, indicated Cu clustering after reduction at both 300 and 450 °C and passivation (Figure S5). The absence of Cu–Cu metal scattering in the EXAFS spectrum and lack of LSPR feature in the DRUV–vis–NIR spectrum of the Cu/SiAlO catalyst reduced at 300 °C suggests that the clustered Cu species remained oxidized (e.g., Cu2O-like species) until reduction to metallic Cu at 450 °C. The combined TPR, DRUV–vis–NIR, XAS, and STEM characterization data indicate that metallic Cu nanoparticles formed on the Cu/Al2O3, Cu/ZrO2, and Cu/ZrAlO materials after reduction at 300 °C, and on the Cu/SiAlO after reduction at 450 °C. In combination with the acid site characterization, this suite of materials represents bifunctional metal–acid catalysts with varying acid site strength for exploration in the methanol conversion chemistry.

Catalytic Testing of the DHC Reaction

Catalytic performance in the DHC reaction of methanol to DMM was evaluated at 1.7 atm, with varying temperature (175–225 °C) and WHSV (2.5–9.5 h–1). There were no effects from internal or external mass transfer limitations (calculations provided in the Supporting Information). Under these reaction conditions, the catalysts were stable and did not show considerable deactivation as measured by methanol conversion, product selectivity, or product yield over the course of 22–32 h time-on-stream (Figures S7 and S8). The results for reaction at 200 °C and WHSV of 5 h–1 are presented in Table 2 and compared with the results from the Cu-free support materials at the same conditions. Data is reported as the average ± standard deviation of 10–12 data points during approximately 6 h time-on-stream (Table 2). Selectivity is reported as C-selectivity. Methanol conversions were below 15% except for the most active Cu/ZrAlO catalyst, which exhibited a conversion of 24.7%. For Al2O3, ZrO2, and ZrAlO materials, conversion increased when Cu was added, serving as the first indication of additional reaction pathways due to the addition of metal species.

Table 2. Catalytic Performance of the Unmodified Supports and Supported Cu Catalysts in the Conversion of Methanol to DMMa.

catalyst X (%) SDME (%) SMF (%) SDMM (%) SHCs (%) SCOx (%) H2 productivity (mg/gcat/h)
Al2O3 7.1 ± 0.1 98.3 ± 0.3 0.7 ± 0.1 0.0 ± 0.0 0.8 ± 0.1 0.2 ± 0.0 0.5
Cu/Al2O3 14.0 ± 0.2 72.3 ± 0.8 1.3 ± 0.1 7.0 ± 0.2 0.2 ± 0.0 19.3 ± 1.2 25.0
SiAlO 3.7 ± 0.5 78.1 ± 2.8 7.1 ± 1.2 0.3 ± 0.1 9.8 ± 0.8 4.7 ± 0.8 1.9
Cu/SiAlO 3.0 ± 0.1 85.7 ± 3.9 1.8 ± 0.6 3.6 ± 0.8 3.3 ± 0.7 5.6 ± 1.9 4.4
ZrO2 2.5 ± 0.1 2.1 ± 0.1 8.1 ± 0.1 3.2 ± 0.2 10.4 ± 0.6 76.2 ± 0.7 16.8
Cu/ZrO2 10.7 ± 0.2 0.5 ± 0.1 63.5 ± 1.5 0.7 ± 0.1 4.6 ± 0.1 30.7 ± 1.5 45.8
ZrAlO 9.4 ± 0.4 63.6 ± 3.0 13.5 ± 2.1 0.3 ± 0.1 3.6 ± 0.2 19.0 ± 1.0 17.8
Cu/ZrAlO 24.7 ± 0.2 47.4 ± 0.3 4.4 ± 0.1 12.0 ± 0.3 0.8 ± 0.1 35.4 ± 0.2 84.3
a

Cu catalysts were reduced with H2 at 300 °C prior to reaction (450 °C for Cu/SiAlO). Reaction conditions: 200 °C, 1.7 atm, methanol WHSV = 5 h–1. X = conversion; Si = C-selectivity of product i. DME = dimethyl ether, MF = methyl formate, DMM = dimethoxymethane, HCs = hydrocarbons, COx = CO and CO2.

The major observed products were DME, MF, DMM, hydrocarbons (HCs, predominantly methane), CO and CO2 (combined as COx), and H2. Formaldehyde was only observed in trace amount (less than 0.1% C-selectivity). A proposed reaction network based on the known chemistry of methanol and formaldehyde to OMEs is presented in Scheme 2,2 and consideration of this reaction network aids the interpretation of the observed selectivity. Acid-catalyzed dehydration of methanol to DME, which is known to occur at both Lewis and Brønsted acid sites, was the major pathway observed over the irreducible, predominantly Lewis-acidic supports, where DME C-selectivity was 98.3% and 78.1% over Al2O3 and SiAlO catalysts, respectively. The ZrO2 support favored methanol decomposition to COx (76% C-selectivity), which is often attributed to basic sites on ZrO2,45,46 and a low level of background dehydrogenation activity on this reducible support was observed (8.1% C-selectivity to MF). Dehydration was not observed at the comparatively weaker Lewis acid sites on ZrO2 than on Al2O3 and SiAlO catalysts. Consistent with the py-DRIFTS characterization indicating a distribution of Lewis acid strength, the ZrAlO catalyst exhibited intermediate reactivity between Al2O3 and ZrO2, demonstrating significantly lower C-selectivity to DME (66.7%) and COx (19.8%), while selectivity to the dehydrogenation product (i.e., MF) increased (13.5%).

Scheme 2. Proposed Reaction Network for DHC Reaction of Methanol to DMM over Bifunctional Metal–Acid Catalysts.

Scheme 2

Metal or acid sites for each step are shown in italics.2

With the addition of Cu, dehydrogenation and DHC were observed over Cu/Al2O3, Cu/ZrO2, and Cu/ZrAlO catalysts, evidenced by increased selectivity to MF or DMM and enhanced production of H2. In contrast, Cu/SiAlO exhibited slightly reduced activity compared to the support and minimal DHC activity. This is tentatively attributed to lower dehydrogenation activity of larger Cu particles observed by XRD and STEM for the Cu/SiAlO catalyst, rather than the small nano-Cu on the other catalysts. MF is the product of formaldehyde dimerization (i.e., Tischenko reaction),4749 rather than participating in the DHC reaction. It is worth noting that hydrogenation of MF would enable it to participate in DMM formation, and the Cu catalysts reported here may perform this reaction with H2 generated via dehydrogenation. However, specific experiments to explore MF hydrogenation were not the focus of this report. The formation of DMM demonstrates the cooperativity between metal and acid sites, enabling the DHC chemistry outlined in Scheme 2. For Cu/Al2O3, the DMM C-selectivity was 7.0%, but the Al2O3-catalyzed production of DME remained the dominant pathway (72.3% C-selectivity) and methanol decomposition to COx was also observed (19.3% C-selectivity). Cu/ZrO2 significantly increased methanol conversion to MF versus the Cu-free support (63.5% versus 8.1% C-selectivity on the support). This catalyst demonstrated Cu-based dehydrogenation activity for the formation of formaldehyde with subsequent dimerization to form MF. However, the lack of DMM production is attributed to the weak Lewis acid sites on ZrO2 that were also inactive for dehydration. The Cu/ZrAlO catalyst, having a distribution of Lewis acid strengths compared to Al2O3 and ZrO2, exhibited the highest methanol conversion (24.7%) with a 12.0% C-selectivity to DMM. We attribute the decreased DME C-selectivity versus Cu/Al2O3 (47.4 vs 72.6%) to the weaker Lewis acid strength but note that there was an increase in COx selectivity (35.4 vs 19.3%) attributed to the ZrOx species.

Despite the carbon-loss to COx, the single-pass DMM carbon-yield tripled from 1.0% over Cu/Al2O3 to 3.0% over Cu/ZrAlO. Gas-phase dehydrogenation of methanol to MF over supported Cu catalysts has been reported4753 but no effort has focused on the synthesis of DMM via gas-phase DHC of methanol. A few studies have investigated liquid-phase methanol dehydrogenation to DMM and although DMM selectivity in the liquid-phase products was high (i.e., close to 100%), the overall yield was either very low (below 0.5%)54 or not reported.55 Thus, the single-pass DMM yield achieved in this report (3.0%) greatly exceeds previous reports for the direct DMM synthesis from methanol via the DHC approach, and importantly, this was demonstrated at remarkably lower temperature (200 °C) than typical methanol dehydrogenation (500–800 °C).

The production of DMM through the DHC reaction is expected to be sensitive to reaction temperature due to thermodynamic equilibrium limitations. Methanol dehydrogenation is only spontaneous above 450 °C,56 but the coupling of formaldehyde with methanol is preferred at lower temperatures—below 300 °C for DMM2,19,57 and below 120 °C for OMEs5759 due to equilibrium considerations. Increased reaction temperature favors the formation of formaldehyde,56 the key intermediate for DMM formation, and therefore the effect of reaction temperature on DMM production was investigated experimentally (Table S2). When the reaction temperature was increased from 200 to 225 °C, competing reactions from methanol dehydration to DME and decomposition to COx limited DMM formation. DME was the favored product over Cu on acidic catalysts (i.e., Al2O3, SiAlO and ZrAlO), and dehydrogenation activity dropped significantly, evidenced by the decrease in MF and DMM selectivity and yield (<0.5% yield). The Cu/ZrO2 catalyst favored MF and COx products (13.6 and 3.1% yield, respectively, DMM yield of 0.1%). For all catalysts, decreasing the reaction temperature from 200 to 175 °C favored dehydrogenation and DHC products over dehydration, exhibiting increased selectivity to MF and DMM and decreased selectivity to DME (Table S2). However, overall activity was also reduced significantly, resulting in low single-pass DMM yields less than 1.0%. Figure S9 presents the equilibrium carbon yield for methanol decomposition (CO), dehydrogenation (HCHO, MF), and the DHC (DMM) reaction pathways. Due to the thermodynamic limitations, the equilibrium single-pass C-yield of DMM at the present condition (200 °C, 0.85 atm of methanol) is 7.6%. Thus, the DMM C-yield over Cu/ZrAlO reported here reaches 40% of the equilibrium limit. Despite the modest equilibrium-limited yield, similar thermodynamically challenging processes remain industrially relevant (e.g., ammonia production, alkane dehydrogenation). From an economic perspective, these types of processes rely on high-value products produced in sufficiently high volumes to leverage economies of scale. On the production side, these processes typically rely on high product selectivity, efficient product separation with reactant recycle, and low residence times (i.e., rapid reaction kinetics). If demand for DMM and/or OMEs reaches that of high-volume fuel blendstocks, these same process parameters will be important in the development of the DHC reaction. From a catalyst development standpoint, reaching the equilibrium-limited yield can be addressed through tailoring the bifunctional metal–acid active sites identified here to further favor acetalization, as the marginal improvement in the DMM equilibrium yield with temperature is hindered by greater DME and COx equilibrium yields at higher temperatures.

The relationship between product selectivity and methanol conversion is presented in Figure 4 for Cu/ZrAlO and Cu/Al2O3 catalysts, our two best performing catalysts. Different conversions were achieved by varying WHSV from 5 to 9.5 h–1 for Cu/ZrAlO and from 2.5 to 5 h–1 for Cu/Al2O3. In general, selectivity to dehydrogenation and DHC products (i.e., MF and DMM, respectively) decreased as conversion increased, with a corresponding increase in DME or COx selectivity for the reaction over Cu/ZrAlO or Cu/Al2O3 catalysts, respectively. For Cu/Al2O3, high DMM selectivity was only achieved at low conversion (e.g., 12% selectivity at ca. 10% conversion) and decreased rapidly to 2.5% as conversion increased to 18%. In contrast, DMM selectivity was more consistent over Cu/ZrAlO, where it only decreased slightly with increasing conversion, from 17% at 18% conversion to 12% at 25% conversion. At similar conversions (18%), Cu/ZrAlO exhibited markedly greater selectivity to DMM than Cu/Al2O3 (16.7% vs 2.5%). It is important to note that for both catalysts, undesired products (e.g., DME and COx) still had a significant contribution to product slates, emphasizing an area for further catalyst development.

Figure 4.

Figure 4

Product carbon selectivity as a function of methanol conversion over (A) Cu/ZrAlO and (B) Cu/Al2O3 catalysts. Reaction temperature was 200 °C, and reaction pressure was 1.7 atm. Lines between data points serve as a guide to the eye to highlight the trend.

Rather than comparing catalyst performance using turnover frequency for the multi-step, multi-site (i.e., metal and acid sites) conversion of methanol to DMM, Sun et al.2 compared catalysts by calculating the DMM formation rate. This metric was developed for the oxidative dehydrogenation coupling route, since this report is the first for the gas-phase DHC approach. Catalyst performance varied over a wide range of 0.4–6.6 h–1 for non-noble, mixed-metal oxide catalysts (e.g., V2O5/TiO2). The DMM formation rate observed over Cu/ZrAlO in this study was 6.1 h–1, which is in the range of high performing non-noble, mixed-metal oxide catalysts. The highest DMM formation rate (23.5 h-1) achieved for non-noble, mixed-metal oxide catalysts was over a commercial iron-molybdate catalyst, which is a well-known catalyst for formaldehyde production via methanol oxidation.60 Similar ranges of DMM formation rate were also reported for supported noble metal catalysts, such as Re (3–19 h–1) or Ru (0–26 h–1), but the activity varied widely depending on the support and synthesis procedure. There are significant concerns that limit applicability of these noble metal and commercial iron molybdate (FeMo) catalysts: (i) sublimation at high reaction temperature (240–300 °C) for Re and FeMo catalysts, and (ii) formation of byproduct MF (35–70% selectivity) due to the redox ability of Ru catalysts, which requires costly and energy intensive separation.2 Thus, the non-noble metal, supported Cu catalysts reported here, which generate DMM and hydrogen via the DHC pathway, do not contain volatile species and demonstrate lower MF selectivity, addressing these shortcomings and representing a promising foundation for continued catalyst and process development.

Finally, stability of the Cu species on the Cu/Al2O3 and Cu/ZrAlO catalysts was investigated using in situ EXAFS analysis (Figure S10). The spectra collected at 200 °C under flowing methanol exhibit a prominent Cu–Cu scattering peak at ca. 2.20 Å (phase uncorrected distance). At initial time-on-stream, the fitted coordination numbers of Cu/Al2O3 and Cu/ZrAlO are 9.4 and 8.2, respectively (Tables S3 and S4). As described above for the reduced Cu materials, these coordination number values are consistent with the presence of small Cu nanoparticles on the oxide supports. At longer times under flowing methanol, the intensity of the Cu–Cu scattering peak slightly increased, giving an increase in the average coordination number of less than 0.5 after 6 h for both samples, indicating minimal particle growth or sintering over this time. Although the in situ conditions used for XAS measurements were not able to precisely match those of the performance tests described above, the results suggest that the Cu nanoparticles are stable under these mild DHC reaction conditions. Further, these XAS results are consistent with the lack of observed catalyst deactivation during the course of reaction under varying conditions as presented above (Figures S7 and S8).

Conclusions

Conversion of methanol to DMM serves as a model reaction for catalyst development in the single-step production of OMEs. In contrast to common oxidative dehydrogenation routes, where H2 is removed as water, a DHC route to OMEs generates H2, which can be advantageous for other processes in a biorefinery. In this report, Cu supported on commercial acidic non-reducible oxides (SiAlO and Al2O3) and acidic reducible oxides (ZrO2, ZrAlO) provides metal–acid bifunctionality to facilitate the gas-phase DHC reaction of methanol to DMM under remarkably mild conditions. The Cu/ZrAlO catalyst demonstrated the highest activity among the investigated catalysts, achieving a methanol conversion of 24.7% and a resulting DMM single-pass yield of 3.0%, which is 40% of the equilibrium limit under these conditions and greatly exceeds the yields of ca. 0.5% in liquid-phase chemistry. Considering the reaction network, the greater DMM yield is attributed to decreased DME formation over the lower strength acid sites of the Cu/ZrAlO catalyst. In contrast, the stronger acid sites of Cu/Al2O3 favored DME formation, and the weaker acid sites of Cu/ZrO2 were not active for acetalization or DME formation. These results also indicate that metallic Cu nanoparticles are effective for the dehydrogenation of methanol under these mild conditions, providing an advantage for methanol–formaldehyde coupling in a single reactor versus typical high-temperature dehydrogenation catalysts (i.e., 500–800 °C). A DMM formation rate of 6.1 h–1 was achieved over Cu/ZrAlO, comparable to high-performing non-noble metal-oxide catalysts for direct DMM synthesis via traditional methanol oxidation. This report sets the foundation for catalyst development to further increase DMM selectivity via tailored bifunctional metal–acid active sites that favor acetalization and disfavor decomposition, dehydration, and dimerization.

Acknowledgments

This work was authored in part by the National Renewable Energy Laboratory, operated by Alliance for Sustainable Energy, LLC, in part by Argonne National Laboratory, operated by The University of Chicago, and in part by Oak Ridge National Laboratory, operated by UT-Battelle, LLC, for the U.S. Department of Energy (DOE) under Contract Nos. DE-AC36-08GO28308, DE-AC02-06CH11357, and DE-AC05-00OR22725, respectively. A portion of this research was conducted as part of the Co-Optimization of Fuels & Engines (Co-Optima) project sponsored by the U.S. Department of Energy, Office of Energy Efficiency and Renewable Energy, Bioenergy Technologies and Vehicle Technologies Offices. Co-Optima is a collaborative project of several national laboratories initiated to simultaneously accelerate the introduction of affordable, scalable, and sustainable biofuels and high-efficiency, low-emission vehicle engines. Part of this research was conducted in collaboration with the Chemical Catalysis for Bioenergy (ChemCatBio) Consortium, a member of the Energy Materials Network (EMN). This research used resources of the Advanced Photon Source, a U.S. DOE Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract DE-AC02-06CH11357. MRCAT operations were supported by the DOE and the MRCAT member institutions. The microscopy was performed as part of a user project at the Center for Nanophase Materials Sciences (CNMS), which is a U.S. DOE Office of Science User Facility. The views expressed in this article do not necessarily represent the views of the DOE or the U.S. Government. The U.S. Government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssuschemeng.0c03606.

  • Experimental details; mass transfer limitation calculations; N2 physisorption data; plots of NH3-TPD, XRD, TPR, STEM, and XAS characterization data; catalytic data at varying temperatures and space velocities; plot of equilibrium-limited C-yield; and EXAFS fitting parameters with and without flowing methanol (PDF)

The authors declare no competing financial interest.

Supplementary Material

sc0c03606_si_001.pdf (1.3MB, pdf)

References

  1. Liu H.; Wang Z.; Wang J.; He X.; Zheng Y.; Tang Q.; Wang J. Performance, Combustion and Emission Characteristics of a Diesel Engine Fueled with Polyoxymethylene Dimethyl Ethers (PODE3–4)/ Diesel Blends. Energy 2015, 88, 793–800. 10.1016/j.energy.2015.05.088. [DOI] [Google Scholar]
  2. Sun R.; Delidovich I.; Palkovits R. Dimethoxymethane as a Cleaner Synthetic Fuel: Synthetic Methods, Catalysts, and Reaction Mechanism. ACS Catal. 2019, 9, 1298–1318. 10.1021/acscatal.8b04441. [DOI] [Google Scholar]
  3. ASTM International . Standard Specification for Dimethyl Ether for Fuel Purposes; ASTM D7901-14b; West Conshohocken, PA, 2014. 10.1520/D7901-14B. [DOI]
  4. Burger J.; Siegert M.; Ströfer E.; Hasse H. Poly(Oxymethylene) Dimethyl Ethers as Components of Tailored Diesel Fuel: Properties, Synthesis and Purification Concepts. Fuel 2010, 89 (11), 3315–3319. 10.1016/j.fuel.2010.05.014. [DOI] [Google Scholar]
  5. Lumpp B.; Rothe D.; Pastotter C.; Lammermann R.; Jacob E. Oxymethylene Ethers as Diesel Fuel Additives of The Future. MTZ Worldwide eMag. 2011, 72 (3), 34–38. 10.1365/s38313-011-0027-z. [DOI] [Google Scholar]
  6. Härtl M.; Seidenspinner P.; Jacob E.; Wachtmeister G. Oxygenate Screening on a Heavy-Duty Diesel Engine and Emission Characteristics of Highly Oxygenated Oxymethylene Ether Fuel OME1. Fuel 2015, 153, 328–335. 10.1016/j.fuel.2015.03.012. [DOI] [Google Scholar]
  7. Feiling A.; Munz M.; Beidl C. Potential of the Synthetic Fuel OME1b for the Soot-free Diesel Engine. ATZextra Worldwide 2016, 21 (11), 16–21. 10.1007/s40111-015-0516-1. [DOI] [Google Scholar]
  8. Omari A.; Heuser B.; Pischinger S. Potential of Oxymethylenether-Diesel Blends for Ultra-Low Emission Engines. Fuel 2017, 209, 232–237. 10.1016/j.fuel.2017.07.107. [DOI] [Google Scholar]
  9. Pélerin G. K. D.; Härtl M.; Wachtmeister G.. Recent Results of The Sootless Diesel Fuel Oxymethylene Ether. In Internationaler Motorenkongress 2017 Wiesbaden, June 2017; Liebl C., Beidl C., Eds.; Springer Vieweg, Wiesbaden, 2017; pp 439–456. 10.1007/978-3-658-17109-4_28. [DOI]
  10. Hagen G. P.; Spangler M. J.. Preparation of Polyoxymethylene Dimethyl Ethers by Catalytic Conversion of Dimethyl Ether with Formaldehyde Formed by Dehydrogenation of Dimethyl Ether. US 6160186 A, 2000.
  11. Burger J.; Ströfer E.; Hasse H. Production Process for Diesel Fuel Components Poly(Oxymethylene) Dimethyl Ethers from Methane-Based Products by Hierarchical Optimization with Varying Model Depth. Chem. Eng. Res. Des. 2013, 91 (12), 2648–2662. 10.1016/j.cherd.2013.05.023. [DOI] [Google Scholar]
  12. Drunsel J. O.; Renner M.; Hasse H. Experimental Study and Model of Reaction Kinetics of Heterogeneously Catalyzed Methylal Synthesis. Chem. Eng. Res. Des. 2012, 90 (5), 696–703. 10.1016/j.cherd.2011.09.014. [DOI] [Google Scholar]
  13. Lausche A. C.; Hummelshøj J. S.; Abild-Pedersen F.; Studt F.; Nørskov J. K. Application of a New Informatics Tool in Heterogeneous Catalysis: Analysis of Methanol Dehydrogenation on Transition Metal Catalysts for The Production of Anhydrous Formaldehyde. J. Catal. 2012, 291, 133–137. 10.1016/j.jcat.2012.04.017. [DOI] [Google Scholar]
  14. Tan E. C. D.; Talmadge M.; Dutta A.; Hensley J.; Schaidle J.; Biddy M.; Humbird D.; Snowden-Swan L. J.; Ross J.; Sexton D.; Yap R.; Lukas J.. Process Design and Economics for the Conversion of Lignocellulosic Biomass to Hydrocarbons via Indirect Liquefaction. Thermochemical Research Pathway to High-Octane Gasoline Blendstock Through Methanol/Dimethyl Ether Intermediates; NREL/TP-5100-62402; National Renewable Energy Lab. (NREL), Golden, CO, 2015. 10.2172/1215006 [DOI]
  15. Tan E. C.; Talmadge M.; Dutta A.; Hensley J.; Snowden-Swan L. J.; Humbird D.; Schaidle J.; Biddy M. Conceptual Process Design and Economics for the Production of High-Octane Gasoline Blendstock via Indirect Liquefaction of Biomass through Methanol/Dimethyl Ether Intermediates. Biofuels, Bioprod. Biorefin. 2016, 10 (1), 17–35. 10.1002/bbb.1611. [DOI] [Google Scholar]
  16. Liu H.; Iglesia E. Selective One-Step Synthesis of Dimethoxymethane via Methanol or Dimethyl Ether Oxidation on H3+nVnMo12-nPO40 Keggin Structures. J. Phys. Chem. B 2003, 107 (39), 10840–10847. 10.1021/jp0301554. [DOI] [Google Scholar]
  17. Zhang Y.; Drake I. J.; Briggs D. N.; Bell A. T. Synthesis of Dimethyl Carbonate and Dimethoxy Methane over Cu-ZSM-5. J. Catal. 2006, 244 (2), 219–229. 10.1016/j.jcat.2006.09.002. [DOI] [Google Scholar]
  18. Chen S.; Meng Y.; Zhao Y.; Ma X.; Gong J. Selective Oxidation of Methanol to Dimethoxymethane over Mesoporous Al-P-V-O Catalysts. AIChE J. 2013, 59 (7), 2587–2593. 10.1002/aic.14033. [DOI] [Google Scholar]
  19. 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 (4), 958–970. 10.1039/C5CY01858G. [DOI] [Google Scholar]
  20. Dutta A.; Schaidle J. A.; Humbird D.; Baddour F. G.; Sahir A. Conceptual Process Design and Techno-Economic Assessment of Ex Situ Catalytic Fast Pyrolysis of Biomass: A Fixed Bed Reactor Implementation Scenario for Future Feasibility. Top. Catal. 2016, 59 (1), 2–18. 10.1007/s11244-015-0500-z. [DOI] [Google Scholar]
  21. Dillich S.; Ramsden T.; Melaina M.. Hydrogen Production Cost Using Low-Cost Natural Gas. DOE Hydrogen and Fuel Cells Program Record #12024; Sept 19, 2012.
  22. James B. D; DeSantis D. A.; Saur G.. Final Report: Hydrogen Production Pathways Cost Analysis (2013–2016); National Renewable Energy Lab. (NREL), Golden, CO, Argonne National Lab. (ANL), Argonne, IL, 2016. 10.2172/1346418 [DOI]
  23. Sato A. G.; Volanti D. P.; de Freitas I. C.; Longo E.; Bueno J. M. C. Site-selective Ethanol Conversion over Supported Copper Catalysts. Catal. Commun. 2012, 26, 122–126. 10.1016/j.catcom.2012.05.008. [DOI] [Google Scholar]
  24. Wang Q. N.; Shi L.; Lu A. H. Highly Selective Copper Catalyst Supported on Mesoporous Carbon for the Dehydrogenation of Ethanol to Acetaldehyde. ChemCatChem 2015, 7 (18), 2846–2852. 10.1002/cctc.201500501|. [DOI] [Google Scholar]
  25. Hernández W. Y.; De Vlieger K.; Van Der Voort P.; Verberckmoes A. Ni-Cu Hydrotalcite-Derived Mixed Oxides as Highly Selective and Stable Catalysts for the Synthesis of β-Branched Bioalcohols by the Guerbet Reaction. ChemSusChem 2016, 9 (22), 3196–3205. 10.1002/cssc.201601042. [DOI] [PubMed] [Google Scholar]
  26. Ponomareva E. A.; Krasnikova I. V.; Egorova E. V.; Mishakov I. V.; Vedyagin A. A. Ethanol Dehydrogenation over Copper Supported on Carbon Macrofibers. Mendeleev Commun. 2017, 27 (2), 210–212. 10.1016/j.mencom.2017.03.035. [DOI] [Google Scholar]
  27. Zhang J.; Fang D.; Liu D. Evaluation of Zr–Alumina in Production of Polyoxymethylene Dimethyl Ethers from Methanol and Formaldehyde: Performance Tests and Kinetic Investigations. Ind. Eng. Chem. Res. 2014, 53 (35), 13589–13597. 10.1021/ie501231a. [DOI] [Google Scholar]
  28. Parry E. P. An Infrared Study of Pyridine Adsorbed on Acidic Solids. Characterization of Surface Acidity. J. Catal. 1963, 2 (5), 371–379. 10.1016/0021-9517(63)90102-7. [DOI] [Google Scholar]
  29. Zaki M. I.; Hasan M. A.; Pasupulety L. Surface Reactions of Acetone on Al2O3, TiO2, ZrO2, and CeO2: IR Spectroscopic Assessment of Impacts of the Surface Acid-Base Properties. Langmuir 2001, 17 (3), 768–774. 10.1021/la000976p. [DOI] [Google Scholar]
  30. Crépeau G.; Montouillout V.; Vimont A.; Mariey L.; Cseri T.; Maugé F. Nature, Structure and Strength of the Acidic Sites of Amorphous Silica Alumina: An IR and NMR Study. J. Phys. Chem. B 2006, 110 (31), 15172–15185. 10.1021/jp062252d. [DOI] [PubMed] [Google Scholar]
  31. Bailleul S.; Yarulina I.; Hoffman A. E. J.; Dokania A.; Abou-Hamad E.; Chowdhury A. D.; Pieters G.; Hajek J.; De Wispelaere K.; Waroquier M.; Gascon J.; Van Speybroeck V. A Supramolecular View on the Cooperative Role of Brønsted and Lewis Acid Sites in Zeolites for Methanol Conversion. J. Am. Chem. Soc. 2019, 141 (37), 14823–14842. 10.1021/jacs.9b07484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Bennici S.; Gervasini A.; Ravasio N.; Zaccheria F. Optimization of Tailoring of CuOx Species of Silica Alumina Supported Catalysts for the Selective Catalytic Reduction of NOx. J. Phys. Chem. B 2003, 107 (22), 5168–5176. 10.1021/jp022064x. [DOI] [Google Scholar]
  33. Yan J. Y.; Lei G. D.; Sachtler W. M. H.; Kung H. H. Deactivation of Cu/ZSM-5 Catalysts for Lean NOxReduction: Characterization of Changes of Cu State and Zeolite Support. J. Catal. 1996, 161 (1), 43–54. 10.1006/jcat.1996.0160. [DOI] [Google Scholar]
  34. Hoang D. L.; Dang T. T. H.; Engeldinger J.; Schneider M.; Radnik J.; Richter M.; Martin A. TPR Investigations on the Reducibility of Cu Supported on Al2O3, Zeolite Y and SAPO-5. J. Solid State Chem. 2011, 184 (8), 1915–1923. 10.1016/j.jssc.2011.05.042. [DOI] [Google Scholar]
  35. Shimokawabe M.; Asakawa H.; Takezawa N. Characterization of Copper/Zirconia Catalysts Prepared by an Impregnation Method. Appl. Catal. 1990, 59 (1), 45–58. 10.1016/S0166-9834(00)82186-7. [DOI] [Google Scholar]
  36. Chary K. V. R.; Sagar G. V.; Srikanth C. S.; Rao V. V. Characterization and Catalytic Functionalities of Copper Oxide Catalysts Supported on Zirconia. J. Phys. Chem. B 2007, 111 (3), 543–550. 10.1021/jp063335x. [DOI] [PubMed] [Google Scholar]
  37. Boccuzzi F.; Coluccia S.; Martra G.; Ravasio N. Cu/SiO2 and Cu/SiO2-TiO2 Catalysts: I. TEM, DR UV-Vis-NIR, and FTIR Characterization. J. Catal. 1999, 184 (2), 316–326. 10.1006/jcat.1999.2428. [DOI] [Google Scholar]
  38. Hung L. I.; Tsung C. K.; Huang W.; Yang P. Room-Temperature Formation of Hollow Cu2O Nanoparticles. Adv. Mater. 2010, 22 (17), 1910–1914. 10.1002/adma.200903947. [DOI] [PubMed] [Google Scholar]
  39. Giordanino F.; Vennestrøm P. N. R.; Lundegaard L. F.; Stappen F. N.; Mossin S.; Beato P.; Bordiga S.; Lamberti C. Characterization of Cu-exchanged SSZ-13: a Comparative FTIR, UV-Vis, and EPR Study with Cu-ZSM-5 and Cu-β with Similar Si/Al and Cu/Al Ratios. Dalton Trans. 2013, 42 (35), 12741–12761. 10.1039/c3dt50732g. [DOI] [PubMed] [Google Scholar]
  40. Marimuthu A.; Zhang J.; Linic S. Tuning Selectivity in Propylene Epoxidation by Plasmon Mediated Photo-Switching of Cu Oxidation State. Science 2013, 339 (6127), 1590–1593. 10.1126/science.1231631. [DOI] [PubMed] [Google Scholar]
  41. Pakharukova V. P.; Moroz E. M.; Kriventsov V. V.; Larina T. V.; Boronin A. I.; Dolgikh L. Y.; Strizhak P. E. Structure and State of Copper Oxide Species Supported on Yttria-Stabilized Zirconia. J. Phys. Chem. C 2009, 113 (51), 21368–21375. 10.1021/jp907652n. [DOI] [Google Scholar]
  42. Pedersen D. B.; Wang S. Surface Plasmon Resonance Spectra of 2.8 ± 0.5 nm Diameter Copper Nanoparticles in Both Near and Far Fields. J. Phys. Chem. C 2007, 111, 17493–17499. 10.1021/jp075076x. [DOI] [Google Scholar]
  43. Niggli P. Die Kristallstruktur einiger Oxyde I. Z. Kristallogr. - Cryst. Mater. 1922, 57, 253–299. 10.1524/zkri.1922.57.1.253. [DOI] [Google Scholar]
  44. Otte H. M. Lattice Parameter Determinations with an X-Ray Spectrogoniometer by the Debye-Scherrer Method and the Effect of Specimen Condition. J. Appl. Phys. 1961, 32 (8), 1536–1546. 10.1063/1.1728392. [DOI] [Google Scholar]
  45. Tatibouët J. M. Methanol Oxidation as a Catalytic Surface Probe. Appl. Catal., A 1997, 148 (2), 213–252. 10.1016/S0926-860X(96)00236-0. [DOI] [Google Scholar]
  46. Viinikainen T.; Rönkkönen H.; Bradshaw H.; Stephenson H.; Airaksinen S.; Reinikainen M.; Simell P.; Krause O. Acidic and Basic Surface Sites of Zirconia-Based Biomass Gasification Gas Clean-Up Catalysts. Appl. Catal., A 2009, 362 (1), 169–177. 10.1016/j.apcata.2009.04.037. [DOI] [Google Scholar]
  47. Tonner S. P.; Trimm D. L.; Wainwright M. S.; Cant N. W. Dehydrogenation of Methanol to Methyl Formate over Copper Catalysts. Ind. Eng. Chem. Prod. Res. Dev. 1984, 23 (3), 384–388. 10.1021/i300015a012. [DOI] [Google Scholar]
  48. Cant N. W.; Tonner S. P.; Trimm D. L.; Wainwright M. S. Isotopic Labeling Studies of the Mechanism of Dehydrogenation of Methanol to Methyl Formate over Copper-Based Catalysts. J. Catal. 1985, 91 (2), 197–207. 10.1016/0021-9517(85)90334-3. [DOI] [Google Scholar]
  49. Rodriguez-Ramos I.; Guerrero-Ruiz A.; Rojas J. L. G.; Fierro M. L. Dehydrogenation of Methanol to Methyl Formate over Copper-Containing Perovskite-Type Oxides. Appl. Catal. 1991, 68 (1), 217–228. 10.1016/S0166-9834(00)84104-4. [DOI] [Google Scholar]
  50. Ai M. Dehydrogenation of Methanol to Methyl Formate over Copper-Based Catalysts. Appl. Catal. 1984, 11 (2), 259–270. 10.1016/S0166-9834(00)81884-9. [DOI] [Google Scholar]
  51. Guerrero-Ruiz A.; Rodriguez-Ramos I.; Fierro J. L. G. Dehydrogenation of Methanol to Methyl Formate over Supported Copper Catalysts. Appl. Catal. 1991, 72 (1), 119–137. 10.1016/0166-9834(91)85033-R. [DOI] [Google Scholar]
  52. Guerreiro E. D.; Gorriz O. F.; Larsen G.; Arrúa L. A. Cu/SiO2 Catalysts for Methanol to Methyl Formate Dehydrogenation: A Comparative Study using Different Preparation Techniques. Appl. Catal., A 2000, 204 (1), 33–48. 10.1016/S0926-860X(00)00507-X. [DOI] [Google Scholar]
  53. Minyukova T. P.; Simentsova I. I.; Khasin A. V.; Shtertser N. V.; Baronskaya N. A.; Khassin A. A.; Yurieva T. M. Dehydrogenation of Methanol over Copper-Containing Catalysts. Appl. Catal., A 2002, 237 (1), 171–180. 10.1016/S0926-860X(02)00328-9. [DOI] [Google Scholar]
  54. Yamakawa T.; Ohnishi T.; Shinoda S. Methanol Dehydrogenation in the Liquid Phase with Cu-Based Solid Catalysts. Catal. Lett. 1994, 23 (3–4), 395–401. 10.1007/BF00811374. [DOI] [Google Scholar]
  55. Wu L.; Li B.; Zhao C. Direct Synthesis of Hydrogen and Dimethoxylmethane from Methanol on Copper/Silica Catalysts with Optimal Cu+/Cu0 Sites. ChemCatChem 2018, 10 (5), 1140–1147. 10.1002/cctc.201701416. [DOI] [Google Scholar]
  56. Usachev N. Y.; Krukovsky I.; Kanaev S. A. The Nonoxidative Methanol Dehydrogenation to Formaldehyde: A review. Pet. Chem. 2004, 44, 379–394. [Google Scholar]
  57. Lei Y. H.; Sun Q.; Chen Z. X.; Shen J. Y. Theoretical Calculations on the Thermodynamics for the Synthesis Reactions of Polyoxymethylene Dimethyl Ethers. Acta Chim. Sinica 2009, 67 (8), 767–772. [Google Scholar]
  58. Schmitz N.; Burger J.; Hasse H. Reaction Kinetics of the Formation of Poly(oxymethylene) Dimethyl Ethers from Formaldehyde and Methanol in Aqueous Solutions. Ind. Eng. Chem. Res. 2015, 54 (50), 12553–12560. 10.1021/acs.iecr.5b04046. [DOI] [Google Scholar]
  59. Schmitz N.; Homberg F.; Berje J.; Burger J.; Hasse H. Chemical Equilibrium of the Synthesis of Poly(oxymethylene) Dimethyl Ethers from Formaldehyde and Methanol in Aqueous Solutions. Ind. Eng. Chem. Res. 2015, 54 (25), 6409–6417. 10.1021/acs.iecr.5b01148. [DOI] [Google Scholar]
  60. Gornay J.; Sécordel X.; Tesquet G.; de Ménorval B.; Cristol S.; Fongarland P.; Capron M.; Duhamel L.; Payen E.; Dubois J.-L.; Dumeignil F. Direct Conversion of Methanol Into 1,1-Dimethoxymethane: Remarkably High Productivity over an Femo Catalyst Placed Under Unusual Conditions. Green Chem. 2010, 12 (10), 1722–1725. 10.1039/c0gc00194e. [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

sc0c03606_si_001.pdf (1.3MB, pdf)

Articles from ACS Sustainable Chemistry & Engineering are provided here courtesy of American Chemical Society

RESOURCES