SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over...

10
CHEMISTRY 2017 © The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). MnTiO 3 -driven low-temperature oxidative coupling of methane over TiO 2 -doped Mn 2 O 3 -Na 2 WO 4 /SiO 2 catalyst Pengwei Wang, 1 * Guofeng Zhao, 2 * Yu Wang, 1 Yong Lu 1,2Oxidative coupling of methane (OCM) is a promising method for the direct conversion of methane to ethene and ethane (C 2 products). Among the catalysts reported previously, Mn 2 O 3 -Na 2 WO 4 /SiO 2 showed the highest conversion and selectivity, but only at 800° to 900°C, which represents a substantial challenge for commercial- ization. We report a TiO 2 -doped Mn 2 O 3 -Na 2 WO 4 /SiO 2 catalyst by using Ti-MWW zeolite as TiO 2 dopant as well as SiO 2 support, enabling OCM with 26% conversion and 76% C 2 -C 3 selectivity at 720°C because of MnTiO 3 for- mation. MnTiO 3 triggers the low-temperature Mn 2+ Mn 3+ cycle for O 2 activation while working synergistically with Na 2 WO 4 to selectively convert methane to C 2 -C 3 . We also prepared a practical Mn 2 O 3 -TiO 2 -Na 2 WO 4 /SiO 2 catalyst in a ball mill. This catalyst can be transformed in situ into MnTiO 3 -Na 2 WO 4 /SiO 2 , yielding 22% conver- sion and 62% selectivity at 650°C. Our results will stimulate attempts to understand more fully the chemistry of MnTiO 3 -governed low-temperature activity, which might lead to commercial exploitation of a low-temperature OCM process. INTRODUCTION Methane is a clean and cheap hydrocarbon resource that is abundantly available in natural gas, shale gas, and gas hydrates (1, 2). It can also be sustainably produced from renewable biomass, offering much greater availability than crude oil. In this context, the depletion of global crude oil has stimulated intense efforts on converting methane into high-value chemicals and transportable fuels, which are traditionally derived by pe- trochemical routes (2, 3). In particular, light olefins, the key building blocks in modern chemical industry, need an urgent shift from crude oil to methane. To date, the industrial-scale conversion from methane to olefins uses an indirect route, where methane is transformed into CO and H 2 above 700°C, followed by conversion to methanol and then to olefins (46). Recently, a bifunctional catalyst (ZnCrO x combined with mesoporous SAPO zeolite) (7) and cobalt carbide nanoprisms (8) have been reported to be capable of directly converting CO and H 2 into ole- fins with surprising selectivity under mild conditions. Despite these advances, the strong CH bonds in methane make this an energy- demanding approach, and such an indirect pathway has a low atom uti- lization efficiency. One developing technology is the direct conversion of methane into olefins and aromatics in the absence of molecular ox- ygen (O 2 ) using suitable catalysts. Zeolite-supported Mo catalysts (Mo/ zeolites) have been intensively studied (9), and recently, a single-atom iron catalyst embedded in a silica matrix (Fe©SiO 2 ) was reported to have promising methane conversion and light olefin selectivity (10). However, commercial prospects for these processes may be hampered by rapid catalyst deactivation (for Mo/zeolites) and ultrahigh reaction temperature (up to 1100°C for Fe©SiO 2 ). In principle, methane can also be directly converted into C 2 hydro- carbons [ethene (C 2 H 4 ) and ethane (C 2 H 6 )] in the presence of O 2 in the so-called oxidative coupling of methane (OCM) (11). The pioneering work by Keller and Bhasin in 1982 initiated a worldwide research effort to explore this process (12). It has been recognized that the OCM pro- cess includes both a heterogeneous catalytic step, which involves the ac- tivation of O 2 and CH 4 on the catalyst surface to generate methyl (CH 3 ) radicals (13), and a homogeneous gas-phase step, involving the coupling of CH 3 radicals to C 2 H 6 followed by dehydrogenation to C 2 H 4 (3, 11, 14). However, introduction of O 2 is double-edged: Although it saves energy to activate CH 4 , overoxidation is unavoidable. Hundreds of catalysts have been examined since 1982, for C 2 selectivity and ability to suppress overoxidation. One early representative catalyst is lithium-doped magnesia (Li/MgO), where [Li + O ] centers are formed to effectively generate CH 3 from CH 4 , but rapidly deactivates due to Li loss (15). Another typical class of catalysts is based on lanthanide oxide in both pure and promoted forms (16, 17), whose surface oxygen vacancies are responsible for generating reactive oxygen, but with relatively low C 2 selectivity. Of these catalysts, Mn 2 O 3 -Na 2 WO 4 /SiO 2 first reported by Fang et al. (18), is considered optimal in terms of its hundreds of hours of stability and high-temperature productivity (that is, 60 to 80% C 2 se- lectivity and 20 to 30% methane conversion at 800° to 900°C) (19). Since its discovery, this catalyst has been extensively studied with re- spect to preparation/modification, catalytic mechanism, and microki- netic modeling (1923). Despite these intensive efforts, no catalysts have been successfully applied in a commercial process because of the high reaction temperature (24). It is widely accepted that the OCM reaction is usually initiated on the solid catalyst surface by reacting CH 4 with surface oxygen species to form methyl radicals and continues in the gas phase (19). Activating O 2 into desirable species on the catalyst surface is a pivotal step that governs the CH 4 activation to CH 3 and subsequent oxidative dehydrogenation of C 2 H 6 . Recently, Cui et al. provided a highly ordered CaO film doped with Mo 2+ ions and suggested that the formed superoxide anions (O 2 ) attribute to CH 4 activation (25). Schwach et al. showed that polycrystal- line MgO codoped with Fe and Au can effectively activate O 2 to peroxy (O 2 2) species, greatly enhancing the C 2 formation with good stability (26). Titanium-containing oxides were also used with the aim to activate O 2 into desirable active species, but the C 2 yield seemed to be improved only to a limited extent (27, 28). An important question is whether the OCM reaction temperature is dominated by the activation of O 2 . If so, lowering the O 2 activation temperature by catalyst modification will be the key to improving the low-temperature activity of OCM catalysts such as Mn 2 O 3 -Na 2 WO 4 /SiO 2 . To test this idea, we explored a TiO 2 - doped Mn 2 O 3 -Na 2 WO 4 /SiO 2 catalyst for the OCM reaction. MnTiO 3 1 Shanghai Key Laboratory of Green Chemistry and Chemical Processes, East China Normal University, Shanghai 200062, China. 2 School of Chemistry and Molecular Engineering, East China Normal University, Shanghai 200062, China. *These authors contributed equally to this work. Corresponding author. Email: [email protected] (G.Z.); [email protected]. edu.cn (Y.L.) SCIENCE ADVANCES | RESEARCH ARTICLE Wang et al., Sci. Adv. 2017; 3 : e1603180 9 June 2017 1 of 9 on November 10, 2020 http://advances.sciencemag.org/ Downloaded from

Transcript of SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over...

Page 1: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

CHEM ISTRY

1Shanghai Key Laboratory of Green Chemistry and Chemical Processes, East ChinaNormal University, Shanghai 200062, China. 2School of Chemistry and MolecularEngineering, East China Normal University, Shanghai 200062, China.*These authors contributed equally to this work.†Corresponding author. Email: [email protected] (G.Z.); [email protected] (Y.L.)

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

2017 © The Authors,

some rights reserved;

exclusive licensee

American Association

for the Advancement

of Science. Distributed

under a Creative

Commons Attribution

NonCommercial

License 4.0 (CC BY-NC).

Dow

n

MnTiO3-driven low-temperature oxidative coupling ofmethane over TiO2-doped Mn2O3-Na2WO4/SiO2 catalystPengwei Wang,1* Guofeng Zhao,2*† Yu Wang,1 Yong Lu1,2†

Oxidative coupling of methane (OCM) is a promising method for the direct conversion of methane to etheneand ethane (C2 products). Among the catalysts reported previously, Mn2O3-Na2WO4/SiO2 showed the highestconversion and selectivity, but only at 800° to 900°C, which represents a substantial challenge for commercial-ization. We report a TiO2-doped Mn2O3-Na2WO4/SiO2 catalyst by using Ti-MWW zeolite as TiO2 dopant as well asSiO2 support, enabling OCM with 26% conversion and 76% C2-C3 selectivity at 720°C because of MnTiO3 for-mation. MnTiO3 triggers the low-temperature Mn2+↔Mn3+ cycle for O2 activation while working synergisticallywith Na2WO4 to selectively convert methane to C2-C3. We also prepared a practical Mn2O3-TiO2-Na2WO4/SiO2

catalyst in a ball mill. This catalyst can be transformed in situ into MnTiO3-Na2WO4/SiO2, yielding 22% conver-sion and 62% selectivity at 650°C. Our results will stimulate attempts to understand more fully the chemistry ofMnTiO3-governed low-temperature activity, which might lead to commercial exploitation of a low-temperatureOCM process.

loa

on N

ovember 10, 2020

http://advances.sciencemag.org/

ded from

INTRODUCTIONMethane is a clean and cheap hydrocarbon resource that is abundantlyavailable in natural gas, shale gas, and gas hydrates (1, 2). It can also besustainably produced from renewable biomass, offering much greateravailability than crude oil. In this context, the depletion of global crudeoil has stimulated intense efforts on convertingmethane into high-valuechemicals and transportable fuels, which are traditionally derived by pe-trochemical routes (2, 3). In particular, light olefins, the key buildingblocks inmodern chemical industry, need an urgent shift from crude oilto methane. To date, the industrial-scale conversion from methane toolefins uses an indirect route, where methane is transformed into COand H2 above 700°C, followed by conversion to methanol and then toolefins (4–6). Recently, a bifunctional catalyst (ZnCrOx combined withmesoporous SAPO zeolite) (7) and cobalt carbide nanoprisms (8) havebeen reported to be capable of directly converting CO and H2 into ole-fins with surprising selectivity under mild conditions. Despite theseadvances, the strong C–H bonds in methane make this an energy-demanding approach, and such an indirect pathway has a low atom uti-lization efficiency. One developing technology is the direct conversionof methane into olefins and aromatics in the absence of molecular ox-ygen (O2) using suitable catalysts. Zeolite-supported Mo catalysts (Mo/zeolites) have been intensively studied (9), and recently, a single-atomiron catalyst embedded in a silica matrix (Fe©SiO2) was reported tohave promising methane conversion and light olefin selectivity (10).However, commercial prospects for these processes may be hamperedby rapid catalyst deactivation (for Mo/zeolites) and ultrahigh reactiontemperature (up to 1100°C for Fe©SiO2).

In principle, methane can also be directly converted into C2 hydro-carbons [ethene (C2H4) and ethane (C2H6)] in the presence of O2 in theso-called oxidative coupling of methane (OCM) (11). The pioneeringwork by Keller and Bhasin in 1982 initiated a worldwide research effortto explore this process (12). It has been recognized that the OCM pro-cess includes both a heterogeneous catalytic step, which involves the ac-

tivation ofO2 andCH4 on the catalyst surface to generatemethyl (CH3⋅)radicals (13), and ahomogeneous gas-phase step, involving the couplingofCH3⋅ radicals toC2H6 followedbydehydrogenation toC2H4 (3, 11,14).However, introduction of O2 is double-edged: Although it saves energyto activate CH4, overoxidation is unavoidable. Hundreds of catalystshave been examined since 1982, for C2 selectivity and ability to suppressoveroxidation. One early representative catalyst is lithium-dopedmagnesia (Li/MgO), where [Li+O−] centers are formed to effectivelygenerate CH3⋅ from CH4, but rapidly deactivates due to Li loss (15).Another typical class of catalysts is based on lanthanide oxide in bothpure and promoted forms (16, 17), whose surface oxygen vacancies areresponsible for generating reactive oxygen, but with relatively low C2

selectivity. Of these catalysts, Mn2O3-Na2WO4/SiO2 first reported byFang et al. (18), is considered optimal in terms of its hundreds of hoursof stability and high-temperature productivity (that is, 60 to 80% C2 se-lectivity and 20 to 30% methane conversion at 800° to 900°C) (19).Since its discovery, this catalyst has been extensively studied with re-spect to preparation/modification, catalytic mechanism, and microki-netic modeling (19–23). Despite these intensive efforts, no catalystshave been successfully applied in a commercial process because of thehigh reaction temperature (24).

It is widely accepted that theOCMreaction is usually initiated on thesolid catalyst surface by reactingCH4with surface oxygen species to formmethyl radicals and continues in the gas phase (19). Activating O2 intodesirable species on the catalyst surface is a pivotal step that governs theCH4 activation to CH3⋅ and subsequent oxidative dehydrogenation ofC2H6. Recently, Cui et al. provided a highly ordered CaO film dopedwith Mo2+ ions and suggested that the formed superoxide anions (O2

−)attribute to CH4 activation (25). Schwach et al. showed that polycrystal-line MgO codoped with Fe and Au can effectively activate O2 to peroxy(O2

2−) species, greatly enhancing the C2 formation with good stability(26). Titanium-containing oxideswere also usedwith the aim to activateO2 into desirable active species, but the C2 yield seemed to be improvedonly to a limited extent (27, 28). An important question is whether theOCM reaction temperature is dominated by the activation of O2. If so,lowering the O2 activation temperature by catalyst modification will bethe key to improving the low-temperature activity of OCM catalystssuch as Mn2O3-Na2WO4/SiO2. To test this idea, we explored a TiO2-doped Mn2O3-Na2WO4/SiO2 catalyst for the OCM reaction. MnTiO3

1 of 9

Page 2: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

was formed under the reaction conditions and showed a marked en-hancement in activity and selectivity at low reaction temperature(that is, catalyst bed temperature): 26% CH4 conversion with 76% C2

plus C3 [C2-C3; propene (C3H6) and propane (C3H8)] selectivity at720°C or 22% conversion with 62% C2-C3 selectivity at 650°C, usinga typical feed of 50% CH4 in air at a total gas hourly space velocity(GHSV) of 8000 ml gcat.

−1 hour−1. Moreover, this catalyst is stable forat least 500 hours on stream for the OCM process. Most notably, thein situ–formed MnTiO3 triggers the low-temperature Mn2+↔Mn3+

chemical cycle of O2 activation, thereby leading to improvement in theOCM activity at low temperatures.

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

RESULTS AND DISCUSSIONLow-temperature active and selective TiO2-dopedMn2O3-Na2WO4/SiO2 catalystWe initially prepared a series of catalysts by supporting 2.7 weight %(wt %) Mn2O3 and 5.0 wt % Na2WO4 on various Ti-containing mate-rials [includingTi-MWWandTS-1 zeoliteswith Si:Timolar ratio of 40:1,pure anatase-TiO2 (a-TiO2), Ti2O3, and perovskite CaTiO3, which alldelivered very poor catalytic performance for the OCM reaction (fig.S1)], and the as-prepared catalysts were referred to as Mn2O3-Na2WO4/Ti-MWW (TS-1, a-TiO2, Ti2O3, and CaTiO3). For comparison, a ref-erence catalyst Mn2O3-Na2WO4/SiO2 (2.7 wt % Mn2O3 and 5.0 wt %Na2WO4)was prepared strictly according to thewidely applied and rec-ommended procedures (18, 19). The detailed preparation informationis given in the SupplementaryMaterials. All the catalysts were examinedin the OCM reaction initially from 800°C (commonly used for theMn2O3-Na2WO4/SiO2 catalyst), cooling down to 760°C using a feedof 50% CH4 in air at a total GHSV of 8000 ml gcat.

−1 hour−1 to quicklyscreen for a catalyst that might afford acceptable low-temperature cat-alytic activity and selectivity. The reference catalyst Mn2O3-Na2WO4/SiO2 delivered a performance (Fig. 1A) similar to performances re-ported at 800°C (18–23, 29, 30), indicating its effectiveness in screeningthese catalysts. On the basis of this reference catalyst, we can separateexperimental trials into two groups: Ti-MWW– and TS-1–supportedcatalysts with higher performance, and pure TiOx- and CaTiO3-supported counterpartswith lower performance (fig. S1). Consequently,the Ti-MWW–, TS-1–, and SiO2-supported catalysts were furtherexamined with temperature further decreased from 760° to 720°C;the results are shown in Fig. 1A. The Mn2O3-Na2WO4/Ti-MWWshowed the best OCM performance, delivering 26% CH4 conversionand 76% C2-C3 selectivity at 720°C; another zeolite-supported catalystMn2O3-Na2WO4/TS-1 delivered slightly lower activity than theMn2O3-Na2WO4/Ti-MWW catalyst. In contrast, Mn2O3-Na2WO4/SiO2, which is generally considered to be the state-of-the-art catalyst,yielded much lower CH4 conversion and C2-C3 selectivity of only 7.2and 9.2%, respectively, at 720°C. Thus, the unprecedented low-temperature OCM activity observed on the Ti-MWW– and TS-1–produced catalysts is attributed to the Ti-doping effect.

In situ MnTiO3 formation–dependent low-temperatureactivity improvementTo optimize the observed Ti-doping effect and exploit itmore rationallyin futurework, it is imperative to explore the origin of the phenomenon,which we accomplished by structural and chemical characterizations ofthe Mn2O3-Na2WO4/Ti-MWW (Si:Ti of 40:1), Mn2O3-Na2WO4/TS-1 (Si:Ti of 40:1), and Mn2O3-Na2WO4/SiO2 catalysts. Notably, theabovementioned three catalysts (that is, first reacting at 800°C for

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

2 hours and then at 720°C for 0.5 hour; see detailed treatment historyin table S1) were exclusively denoted as Cat-Ti-MWW for Mn2O3-Na2WO4/Ti-MWW, Cat-TS-1 for Mn2O3-Na2WO4/TS-1, andCat-SiO2 for Mn2O3-Na2WO4/SiO2. First, the elemental analyses andlow-temperature N2-sorption measurements of these three catalysts re-vealed almost identical content ofMn,W, andNa as well as similar spe-cific surface areas (1.2 to 1.4m2 g−1) (table S2 and fig. S2), indicating thatthe low-temperature activity difference should be correlated with theirdifferent phase or chemical properties rather than element content andsurface area. Subsequently, the phase properties of the Cat-Ti-MWW,Cat-TS-1, and Cat-SiO2 were probed by x-ray diffraction (XRD). TheXRD patterns are shown in Fig. 1B, and themagnified 2q part from 31°to 37° is shown in Fig. 1C. For the reference catalyst Cat-SiO2, the phaseof the main a-cristobalite as well as of Na2WO4 andMn2O3, which iscommonly reported, was detected (18–23); the Cat-Ti-MWW andCat-TS-1 also hada-cristobalite andNa2WO4phases. Therewas a phasetransformation fromMn2O3 to a new compound,MnTiO3, with almostfull transformation for the Cat-Ti-MWW and partial for the Cat-TS-1(Fig. 1C). This observation, combinedwith the higher activity/selectivityover the Cat-Ti-MWW and Cat-TS-1, indicates an improvementeffect similar to that of MnTiO3 on the CH4 low-temperature conver-sion. As noted above, after undergoing OCM reaction at 800°C beforeperforming at 720°C, MnTiO3 was in situ–formed over these three cat-alysts. Conversely, MnTiO3 was formed only for the Ti-MWW–supported catalyst but not for the TS-1–supported catalyst when directlyrunning at 720°C, associatedwith aCH4 conversion of 25% and aC2-C3

selectivity of 73% for theTi-MWW–supported catalyst but a conversionof only 4% and a C2-C3 selectivity of 16% for the TS-1–supported cat-alyst (fig. S3). Once MnTiO3 was formed for the TS-1–supported cata-lyst after reaction at 800°C, a high CH4 conversion of 23% could beobtained with C2-C3 selectivity increased to 69% as the reaction tem-perature was decreased to 720°C. The abovementioned results demon-strate that the appearance of MnTiO3 is responsible for the low-temperature CH4 conversion (fig. S3).

To confirm the as-observedMnTiO3-governed low-temperature ac-tivity improvement, we synthesized a series of Ti-MWW zeolites withvarying Ti:Si ratio from 1:80 to 1:40, as well as a full-Si zeolite as areference (fig. S4A); all the Ti-MWW–supportedMn2O3-Na2WO4 cat-alysts were directly tested at 720°C.With increase of the Ti:Si ratio in thezeolites, the catalysts showed a monotonously increasing CH4 conver-sion and a synchronous MnTiO3 intensity (Fig. 1, D to F). Moreover,TS-1 zeolites with varying Ti:Si ratios were also synthesized (fig. S4B),and their supported catalysts were also tested in the OCM reaction.Running all these TS-1–supported catalysts directly at 720°C did notlead to MnTiO3 and was associated with a very low CH4 conversionof <5% and a C2-C3 selectivity of <20%. However, MnTiO3 was formedafter the reaction at 800°C and could be sustained when reaction tem-perature was decreased to 720°C, andCH4 conversionwas also progres-sively increased at 720°C along with the corresponding amount ofMnTiO3 (fig. S4, C to E). These results provide solid proof for the criticalrole played by MnTiO3 in enhancing catalyst performance, although theslight differences, which likely result from the structural rigidity differencebetween Ti-MWW and TS-1 zeolites, were not explored further.

Despite the clear proof fromXRDcharacterization, the questionnat-urally arises as to the surface composition and structure of the Cat-Ti-MWWandCat-TS-1 compared with the Cat-SiO2 because the catalyticreaction is a surface-dependent process. To answer this question, wefurther probed the surface states of these three catalysts by the surface-sensitive techniques of x-ray photoelectron spectroscopy (XPS)

2 of 9

Page 3: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

and Raman spectroscopy as good complements for XRD. The XPSresults show almost identical surface content of W, Mn, and Na forthese catalysts (table S3), indicating that their activity/selectivity shouldbe controlled by other factors rather than the surface element contents.However, it should be noted that the XPS spectrum of Mn inMn2O3 isidentical to that in MnTiO3 (fig. S5A); therefore, the Mn2O3-to-MnTiO3 evolution cannot be distinguished by XPS. Moreover, thebinding energies of W, Mn, and Na species on the Cat-Ti-MWWand Cat-TS-1 shifted very slightly compared with the Cat-SiO2 surface(fig. S5, B to F), indicating that theMnTiO3 formation did notmarkedlymodify the original electronic states of these species on the catalyst surface.

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

Another surface-sensitive technique of Raman spectroscopy wasused because it is sensitive to the local structure of oxides especially withpoor crystallinity (31, 32). For the Cat-Ti-MWW, Cat-TS-1, and Cat-SiO2,Mn2O3 andNa2WO4, which are generally considered to be crucialfor the traditional SiO2-supported catalyst in the OCM reaction (19),and MnTiO3, which is paramount for low-temperature activity im-provement as evidenced by the above experiments, were analyzed in par-ticular. For the Cat-SiO2 catalyst, we observed signals of Mn2O3 andNa2WO4 aswell asa-cristobalite (Fig. 1G and fig. S6) that were identicalto those reported previously (33). On the Cat-Ti-MWW and Cat-TS-1,in contrast, we detected no signals of Mn2O3, but found strong signals

Fig. 1. The MnTiO3-governed OCM performance of the TiO2-doped Mn2O3-Na2WO4/SiO2 catalysts. (A) CH4 conversion and C2-C3 selectivity over the Ti-MWW–,TS-1–, and SiO2-supported Mn2O3-Na2WO4 catalysts. (B) XRD patterns of the Mn2O3-Na2WO4/Ti-MWW, Mn2O3-Na2WO4/TS-1, and Mn2O3-Na2WO4/SiO2 catalysts, with (C) themagnified part of 2q from 31° to 37°. a.u., arbitrary units; a-Crist., a-cristobalite. (D) CH4 conversion and C2-C3 selectivity and (E) XRD patterns as well as (F) the magnified part of2q from 31° to 37° of the Mn2O3-Na2WO4/Ti-MWW catalysts with different Si:Ti molar ratio in Ti-MWW zeolites after directly running at 720°C. (G and H) Raman spectra of theTi-MWW–, TS-1–, and SiO2-supported Mn2O3-Na2WO4 catalysts. Reaction conditions: GHSV of 8000 ml gcat.

−1 hour−1 of a feed of 50% CH4 in air. C3 selectivity was 3 to 5%, 2 to3%, and 0 to 2% for all catalysts at a C2-C3 total selectivity above 60%, between 40 and 60%, and below 40%, respectively.

3 of 9

Page 4: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

of MnTiO3, and clear signals of Na2WO4 and a-cristobalite (Fig. 1Gand fig. S6). These Raman results also showed a clear transformationfromMn2O3 toMnTiO3 on theCat-Ti-MWWandCat-TS-1, which con-firmed again the critical role of MnTiO3 in improving low-temperatureactivity. Quite different temperature-dependent surface evolutionwas also captured by Raman for the Ti-MWW– and TS-1–supportedcatalysts (Fig. 1H). When directly running at 720°C for 0.5 hour, Mnspecies in the Ti-MWW–supported catalyst were almost fully trans-formed into MnTiO3 associated with a CH4 conversion of 25%, whereasthe TS-1–supported catalyst displayed a spectrum identical to that ofTiO2with a conversion of 4%. As expected, after the TS-1–supported cat-alyst underwent theOCMreaction at 800°C for 2 hours in advance, dom-inant MnTiO3 was detected, associated with a sharp increase of CH4 to23% at 720°C (fig. S3). The Raman-indicated correspondence of thehigh CH4 conversion to the appearance of surfaceMnTiO3 corroboratedthe similar improvement effect of MnTiO3 on the low-temperatureCH4 conversion once again.

Synergistic catalysis of MnTiO3 with Na2WO4

Our probes of phase and surface structures consistently showed the crit-ical role ofMnTiO3 in enhancing the low-temperatureOCMperformanceof the TiO2-doped Mn2O3-Na2WO4/SiO2 catalyst; even more notably,CH4 conversion and C2-C3 selectivity were progressively enhanced withincreasing amount ofMnTiO3 (Fig. 1, D to F, and fig. S4), suggesting thatMnTiO3 contains the active sites for the OCM reaction. To verify thispoint, we prepared and examined an MnTiO3/SiO2 catalyst in theOCM reaction. At 800°C, this catalyst offered slightly lower CH4 con-version (22% versus 26%) and much lower C2-C3 selectivity (47% ver-sus 75%) than the Mn2O3-Na2WO4/SiO2 catalyst, but similar at 760°C(17% versus 14%; 34% versus 38%) and much higher at 720°C (15%versus 7.2%; 27% versus 9.2%) (fig. S7). Nevertheless, both the activityand the selectivity of the MnTiO3/SiO2 catalyst was much lower thanthe results of the Mn2O3-Na2WO4/Ti-MWW with MnTiO3-Na2WO4

combination (26 to 28% conversion and 76 to 79% C2-C3 selectivity at720° to 800°C; Fig. 1A).Moreover, two other catalysts,Mn2O3/SiO2 andNa2WO4/SiO2, were prepared and tested, delivering much lower CH4

conversion (8 to 14% forMn2O3/SiO2 and 5 to 10% for Na2WO4/SiO2)and C2-C3 selectivity (14 to 18% for Mn2O3/SiO2 and 28 to 33% forNa2WO4/SiO2) at 720° to 800°C (fig. S7). It is thus rational to infer thatthe coexistence of Na2WO4 and MnTiO3 is paramount for the low-temperature OCM reactivity improvement. Subsequently, the questionarises whether theMnTiO3-Na2WO4 combination affects their physicalstructure [such as solely improving dispersion to enhance CH4 conver-sion (34)] or the intrinsic catalytic performance. Therefore, a kineticstudy was carried out over the Mn2O3-Na2WO4/SiO2 and Mn2O3-Na2WO4/Ti-MWW catalysts, and the apparent activation energies (Ea)were calculated with the results as shown in fig. S8. Mn2O3-Na2WO4/Ti-MWW provided a much lower Ea (80 to 110 kJ/mol) than Mn2O3-Na2WO4/SiO2 (180 to 210 kJ/mol) (see the Supplementary Materials forthe calculation details), indicating that there are different active sites in thetwo catalysts and that the MnTiO3-Na2WO4 combination markedlypromotes the catalyst intrinsic performance.

MnTiO3-triggered Mn2+↔Mn3+ chemical cycle—The natureof low-temperature OCM catalysisAs mentioned above, it has been recognized that the OCM processobeys themechanism of heterogeneous-homogeneous radical reactions(3, 11, 14), and the heterogeneous surface activation of O2 and CH4 isthe pivotal prerequisite for the OCM process. Therefore, we inferred

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

that the Mn2O3-Na2WO4/Ti-MWW catalyst can activate O2 andCH4more effectively than the other catalysts. To confirm this conjecture,we designed and conducted a three-step experiment over the Mn2O3-Na2WO4/Ti-MWW and Mn2O3-Na2WO4/SiO2 catalysts: reducing inCH4 stream for 0.5 hour and then switching to an O2 stream, followedby switching toCH4 at the desired temperature for some time (see detailedtreatment history in table S1). The evolution of phase structures and sur-face states of these as-resulting catalysts was characterized by XRD andRaman techniques, with results as shown in Fig. 2. After reducing inCH4 at 800°C, Mn species for the Mn2O3-Na2WO4/Ti-MWW catalystexisted in the form of the MnTiO3 phase; then, full oxidation of MnTiO3

to form Mn2O3 took only 1 min in the O2 stream, and the subsequentreversal of Mn2O3 into MnTiO3 took 10 min in the CH4 stream at800°C (Fig. 2A). Differently, theMn2O3-Na2WO4/SiO2 catalyst afterreducing in the CH4 stream at 800°C displayed clear MnWO4 diffrac-tions. Thereafter, full oxidation of MnWO4 to formMn2O3 took 3 minin theO2 stream,whereas the reversal ofMn2O3 intoMnWO4 took15minin the CH4 stream at 800°C (Fig. 2A). At 760°C, the evolution of theMnTiO3↔Mn2O3 cycle slowed down slightly for theMn2O3-Na2WO4/Ti-MWW catalyst, but that of the MnWO4↔Mn2O3 cycle sloweddown greatly for the Mn2O3-Na2WO4/SiO2 catalyst (Fig. 2B). The ox-idation of MnTiO3 to form Mn2O3 took 2 min, and the reversal ofMn2O3 to MnTiO3 took 15 min. However, the transition of MnWO4

to Mn2O3 took a much longer time of 15 min, and that from Mn2O3

into MnWO3 also took a much longer time of 45 min. At 720°C, theMnTiO3↔Mn2O3 cycle could easily take place despite a longer time:MnTiO3 to Mn2O3 took 4 min and Mn2O3 to MnTiO3 took 30 min(Fig. 2C). Most notably, MnWO4 to Mn2O3 took 30 min, but Mn2O3

could not be fully reversed into MnWO4 even after 1.5 hours in a CH4

stream at 720°C (Fig. 2C). Obviously, the Mn2O3-Na2WO4/Ti-MWWcatalyst delivered a much easier Mn2+↔Mn3+ cycle especially at 720°C(Fig. 2, D to F) because of the participation of MnTiO3. Combining theOCM results of these two catalysts, we conclude that the more facileMn2+↔Mn3+ cycle of the Mn2O3-Na2WO4/Ti-MWW catalyst matchesneatlywith itsmuchhigher performance at 720°C. In addition, theRamanresults of the Mn2O3-Na2WO4/Ti-MWW and Mn2O3-Na2WO4/SiO2

catalysts (Fig. 2, G to I) corroborated that the Mn2O3-Na2WO4/Ti-MWWcatalystofferedamucheasierMn2+↔Mn3+cycle, implying amuch higheractivity to activate O2 (via one half-cycle of Mn2+→Mn3+) and to seizethe reducing species (for the other half-cycle of Mn3+→Mn2+) than theMn2O3-Na2WO4/SiO2 catalyst at low temperature.

Conflicting reports on the structure of the active sites of theMn2O3-Na2WO4/SiO2 catalyst are found in literature, and three typical descrip-tions have beenwidely proposed (19). The unitary site of Na−O−Mnwassuggested because of the similarity between the catalytic performance ofMn2O3-Na2WO4/MgO andNa2MnO4/MgO. In contrast, a redoxmech-anism involving a W6+/W4+ couple with W−O−Si bonds was suggested,and the gas-phaseO2 was proposed to be involved in electron transfer. Inparticular, a two-metal sitemodel was proposed, whereO2 is activated onanMn3+ site to take charge of catalyst activity andCH4 is activated on theneighboringW6+ site to increase catalyst selectivity, with oxygen spilloverfromMn2O3 to theNa2WO4 to accomplish the reactive cycle. Our resultsindicate that theMn2+↔Mn3+ cycle is confined toMn2O3 andMnWO4

(Fig. 2, A to C); therefore, we prefer theMn−Wcombined sitemodel fortheMn2O3-Na2WO4/SiO2 catalyst in our present OCM reaction.More-over, the evolution rate of the Mn2+↔Mn3+ cycle in Mn2O3-MnWO4

was much lower than in MnTiO3-Na2WO4, especially at low tempera-tures (Fig. 2); therefore, we hypothesized thatMnTiO3 plays a dominantrole in enhancing low-temperature activity and Na2WO4 is key to

4 of 9

Page 5: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

improving C2-C3 selectivity. To further confirm this conclusion, weprepared four catalysts, including MnWO4-Mn2O3, MnTiO3-Mn2O3,MnWO4-Mn2O3-Na2WO4, and MnTiO3-Mn2O3-Na2WO4, bygrinding. As expected, the catalysts containing MnTiO3 delivered amuch higher CH4 conversion at 720° to 760°C than the counterpartswith absence of MnTiO3, and the catalysts with Na2WO4 delivered amuch higher C2-C3 selectivity at 720° to 800°C (fig. S9). For example,the MnTiO3-Mn2O3 catalyst delivered a CH4 conversion of 16%,whereas Mn2O3-MnWO4 delivered only 4% at 720°C, but selectivitieswere very low for both catalysts (42% forMnWO4-Mn2O3 and 29% forMnTiO3-Mn2O3). As we added Na2WO4 into MnTiO3-Mn2O3, C2-C3

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

selectivity was sharply increased from29 to 73%,whereas CH4 conversionwas slightly increased from 16 to 22%. These experimental results clearlydemonstrate that MnTiO3 was mainly responsible for low-temperatureO2 and CH4 activation and Na2WO4 for increased selectivity, indicatingthat their combination (Na2WO4-MnTiO3) resulted in a much higherconversion rate and selectivity. This observation could be tentativelyattributed to the synergetic interactions between Mn species (Mn2O3

or MnTiO3) and Na2WO4 [see temperature-programmed reduction(TPR) results in fig. S10 and the corresponding discussion in the Sup-plementaryMaterials], in which the detailed spectroscopic/microscopicinvestigations are particularly desirable.

Fig. 2. The temperature-dependent evolution of phase structures and surface states for the Ti-MWW– and SiO2-supported Mn2O3-Na2WO4 catalysts in O2 orCH4 at 720° to 800°C. (A to C) XRD patterns. The Mn3+ fractions evolving at (D) 800°C, (E) 760°C, and (F) 720°C. (G to I) Raman spectra. The Ti-MWW–supported catalystreduced in CH4 at 800°C for 0.5 hour exhibits dominant MnTiO3 signal. After subsequent oxidation in O2 stream for 1 min at 800°C, 2 min at 760°C, and 4 min at 720°C,the MnTiO3 signal disappears, whereas the TiO2 appears, and is reformed in CH4 stream for 9 min at 800°C, 15 min at 760°C, and 30 min at 720°C. The SiO2-supportedcatalyst reduced in CH4 at 800°C for 0.5 hour exhibits a dominant MnWO4 signal. After subsequent oxidation in O2 stream for 3 min at 800°C, 15 min at 760°C, and30 min at 720°C, the MnWO4 signal disappears, whereas the Mn2O3 appears, and can be reformed in CH4 stream for 15 min at 800°C, 35 min at 760°C, and 60 min at720°C. The detailed calculations of the Mn3+ fractions are given in the Supplementary Materials.

5 of 9

Page 6: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

On the basis of the evolution behavior from Mn2+ to Mn3+ (asso-ciated with MnTiO3 or MnWO4 to Mn2O3 for O2 activation) of theMn2O3-Na2WO4/Ti-MWW and Mn2O3-Na2WO4/SiO2 catalysts inthe O2 stream at 720° to 800°C (Fig. 2, D to F), the transition rate fromMn2+ toMn3+was calculated and shown in Fig. 3A (see the Supplemen-tary Materials for the calculation details). At 800°C, the transition ratefromMn2+ to Mn3+ was 98%/min for the Mn2O3-Na2WO4/Ti-MWWcatalyst and 53%/min for theMn2O3-Na2WO4/SiO2 catalyst, associatedwith similar CH4 conversion (26 to 28%) and C2-C3 selectivity (76 to79%). At 760°C, the rate fromMn2+ to Mn3+ for the Mn2O3-Na2WO4/Ti-MWW was 37%/min with a CH4 conversion of 27% and a C2-C3

selectivity of 78%, whereas the rate for Mn2O3-Na2WO4/SiO2 was de-creased to only 7%/min with a low CH4 conversion of 14% and a poorC2-C3 selectivity of 38%. At 720°C, most notably, the rate fromMn2+ toMn3+ for theMn2O3-Na2WO4/Ti-MWWwas still retained at 13%/minwith a high CH4 conversion of 26% and a high C2-C3 selectivity of 76%,but a very slow rate of only 4%/min was obtained for the Mn2O3-Na2WO4/SiO2 with only 7.2% CH4 conversion and 9.2% C2-C3 se-lectivity. These results show that [Mn2O3-Na2WO4]–based catalystperformance was strongly dependent on the Mn2+-to-Mn3+ rateindependent of temperature, and a transition rate of 10%/min seemedcritical to increased OCM conversion/selectivity. As previously shownfor the Mn2O3-Na2WO4/SiO2 catalyst, O2 is activated on the Mn2+ siteand CH4 on theW

6+ site, and the active oxygen species play a key role inactivating CH4 after spillover to theW

6+ site (19). Accordingly, it is safeto say that the Mn2O3-Na2WO4/SiO2 catalyst has an adequate O2 acti-vation rateonly through theMn2O3↔MnWO4 chemical cycle at 800°Cand therefore delivers acceptably highOCMperformance only at 800°C(Fig. 3, A and B). For the Mn2O3-Na2WO4/Ti-MWW catalyst, in con-trast, because of the MnTiO3-triggered low-temperature Mn2+↔Mn3+

chemical cycle, a highO2 activation rate could be obtained at lower tem-perature (for example, 720°C); as a result, this catalyst yielded betterlow-temperature OCM performance (Fig. 3, A and B).

A practical MnTiO3-Na2WO4/SiO2 catalystInspired by the obtained insight into the MnTiO3-enhanced low-temperature catalysis for the OCM process, we successfully prepareda catalyst by a solvent-free method, simply grinding commercialMn2O3, TiO2, Na2WO4, and SiO2 gel in a high-speed ball mill. Byvarying the loadings of Mn2O3, TiO2, and Na2WO4, the mixed systemyielded a highly effective OCM catalyst in the range of 8 to 15 wt % forNa2WO4, and 6 to 28 wt % forMn2O3 plus TiO2 (with a stoichiometricratio of Mn2O3 to TiO2 to be fully transformed intoMnTiO3), with the

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

SiO2 gel making up the balance (fig. S11). For example, the 6.2Mn2O3-6.3TiO2-10Na2WO4/SiO2 catalyst (6.2 wt%Mn2O3, 6.3wt%TiO2, and10wt%Na2WO4)was first activated in the reaction stream at 800°C for2 hours and delivered an interesting CH4 conversion of 24% with 73%C2-C3 selectivity at 720°C (comparable to the conversion/selectivity forthe Mn2O3-Na2WO4/Ti-MWW catalyst at 720°C in Fig. 1A) and amore interesting 22% CH4 conversion with 62% C2-C3 selectivity evenat 650°C (Fig. 4A). After the reaction from 800° to 650°C, TiO2 andMn2O3 were in situ–transformed into MnTiO3 (Fig. 4, B and C) withthe corresponding loading of 11.8 wt %. The catalytically initiatedreaction over this 11.8MnTiO3-10Na2WO4/SiO2 catalyst (that is,derived from 6.2Mn2O3-6.3TiO2-10Na2WO4/SiO2 after undergoingOCM reaction at 800° to 650°C) compared very favorably with otherreportedOCM catalysts, as summarized in table S4. Among them, anorderedmesoporous SBA-15–supportedMn2O3-Na2WO4 catalyst (34)has been recently reported to yield better results than the reference cat-alyst Mn2O3-Na2WO4/SiO2 at 750°C: CH4 conversion, 14% versus 7%;C2-C3 selectivity, 63% versus 50%. By comparison, our 11.8MnTiO3-10Na2WO4/SiO2 catalyst delivered a much better low-temperatureperformance than these reported catalysts for the OCM process (tableS4). Besides the enhanced low-temperature activity and selectivity, such11.8MnTiO3-10Na2WO4/SiO2 catalyst provided promising chemical/mechanical stability during the OCM reaction. With a total GHSV of8000 ml gcat.

−1 hour−1, a scale-up experiment using 10 ml of particu-late catalyst (20 to 40meshes) was carried out for the stability test, andno deactivation was observed during a 500-hour run at 720°C with afeed gas of 50% methane in air (Fig. 4D). CH4 conversion remainedat 22 to 25% throughout the entire 500-hour testing, whereas C2-C3

selectivity was retained at 68 to 73% with an ethylene/ethane ratio of~1.9. Similar behavior was also seen at 800°C, with well-retained CH4

conversion at 24 to 28% andC2-C3 selectivity at 73 to 77% for a 400-hourrun (fig. S12). Moreover, our 6.2Mn2O3-6.3TiO2-10Na2WO4/SiO2 cata-lyst showed a marked low-temperature reaction ignition property that isan important consideration for the practical OCM process. The OCMreaction could be directly started over this catalyst at a low reaction tem-perature (that is, catalyst bed temperature) of 650°C, offering a highconversionof 17%but a lowC2-C3 selectivity of only 47%.As the reactiontemperaturewas increased from650° to 720°C, a highC2-C3 selectivity of73% could be obtained with an improved conversion of 25% (fig. S13A)because of the facilitation of in situ formation of MnTiO3 at that hightemperature point. The MnTiO3 phase was detected on the sample afterdirect reaction at 720°C for only 1 hour and remained almost unchangedwith prolonged time on stream. In contrast, only a small amount of

Fig. 3. The Mn2+-to-Mn3+ transition rate and proposed catalytic recycle for OCM process. (A) Temperature-dependent transition rate of Mn2+ to Mn3+ correlatedto CH4 conversion. (B) The Mn2O3-Na2WO4/Ti-MWW and Mn2O3-Na2WO4/SiO2 catalysts and the proposed catalytic cycles of Mn2O3-Na2WO4 and MnTiO3-Na2WO4

combinations. Detailed calculations of the transition rate are given in the Supplementary Materials.

6 of 9

Page 7: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

MnTiO3 was formed after a direct reaction at 650°C for 1 hour; however,its content increased with prolonged time on stream, after 8 hours be-coming comparable in results to those of a direct reaction at 720°C for1 hour (fig. S13B). Not surprisingly, at 650°C, CH4 conversion andC2-C3

selectivity gradually increased from 17 and 47% at the beginning to 21and 60% after 8 hours on the reaction stream in association with clearMnTiO3 phase formation while remaining almost stable along with fur-ther prolonged time on stream to 12 hours (fig. S13B).Moreover, becauseof the similarity of MnTiO3 phase intensity between samples after directreaction at 720°C for 1 hour and those after direct reaction at 650°C for8 hours or longer (fig. S13B), the former catalyst sample, not surprising-ly, delivered results comparable to those of the latter one, as the reactiontemperature was reduced from 720° to 650°C (fig. S13).

CONCLUSIONSOur results established a TiO2-doped Mn2O3-Na2WO4/SiO2 catalystsystem, which provides an enhanced low-temperature activity andselectivity in combination with promising stability for the OCM pro-cess. MnTiO3 is in situ–generated in the reaction stream and triggersthe low-temperature MnTiO3↔Mn2O3 chemical cycle for O2 activa-tion, thereby leading to a marked improvement of the low-temperatureactivity/selectivity. Guided by these findings, we discovered anMnTiO3-Na2WO4/SiO2 catalyst obtained by the ball milling methodusing Mn2O3, TiO2, Na2WO4, and SiO2 gel as starting materials, ofwhich the reaction temperature can be further lowered to 650°C withacceptable CH4 conversion (~22%) and C2-C3 selectivity (~62%). Wesuggest that the reaction temperature might be decreased further if theMn2+↔Mn3+ chemical cycle could be accelerated at much lower tem-peratures for the [Mn2O3-Na2WO4]–based catalysts. To accomplishthis goal, discovery of new Mn2+↔Mn3+ chemical cycles and a deep

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

understanding of the chemistry of the MnTiO3 (or analog)–governedlow-temperature OCM activity/selectivity at the atomic level are partic-ularly desirable. In addition, yield or selectivity of C2-C3 products willneed to be further improved, perhaps in an appropriate reactor conceptsuch as a membrane and/or structured reactor.

MATERIALS AND METHODSCatalyst preparationTheMn2O3-Na2WO4/support catalystswithMn2O3 loading of 2.7wt%[Mn(NO3)2 aqueous solution as precursor] and Na2WO4 loading of5 wt % (Na2WO4·2H2O as precursor) for the OCM reaction wereobtained by the incipient wetness impregnation (IWI) method (18, 19).The supports, including Ti-MWW and TS-1 zeolites (Si:Ti molar ratioof 40:1 to 80:1), amorphous SiO2 gel (~73 mm), pure a-TiO2 (~44 mm,99.8%), Ti2O3 (~44 mm, 99.8%), andCaTiO3 (~44 mm, 99%+), were im-pregnated with an aqueous solution of Mn(NO3)2 [analytical reagentgrade (AR), Sinopharm Chemical Reagent Co.] and Na2WO4·2H2O(AR, Sinopharm Chemical Reagent Co.) containing appropriate con-centrations at room temperature, followed by constant stirring for5 hours at room temperature and for 1 hour at 180°C. The resultingslurry was dried at 100°C overnight and then calcined in air at 550°Cfor 2 hours. The Ti-MWW and TS-1 zeolites (Si:Ti molar ratio of40:1 to 80:1) were synthesized according to the reported methods(35, 36). The amorphous SiO2 gel and a-TiO2 were purchased fromAladdin Industrial Corporation. TheTi2O3 andCaTiO3were purchasedfrom Alfa Aesar Co. Ltd.

The MnTiO3/SiO2, Mn2O3/SiO2, and Na2WO4/SiO2 catalysts werealso prepared by the IWImethod. TheMn,Na, andWcontents of thesethree catalysts were consistent with the abovementioned loadings.Taking the MnTiO3/SiO2 catalyst as an example, the amorphous

Fig. 4. The OCM performance and XRD/Raman results of the 6.2Mn2O3-6.3TiO2-10Na2WO4/SiO2 catalyst. (A) Temperature-dependent CH4 conversion and C2-C3selectivity. (B) XRD pattern and (C) Raman spectrum of this catalyst. (D) CH4 conversion and C2-C3 selectivity along with the time on stream at 720°C. Reactionconditions: GHSV of 8000 ml gcat.

−1 hour−1 of a feed of 50% CH4 in air. The C3 selectivity was 3 to 5%, 2 to 3%, and 0 to 2% for all catalysts at a C2-C3 total selectivityabove 60%, between 40 and 60%, and below 40%, respectively.

7 of 9

Page 8: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

SiO2 gel was mixed with a-TiO2 (Si:Ti molar ratio of 40:1) and impreg-nated with an aqueous solution of Mn(NO3)2 (AR, Sinopharm Chem-ical Reagent Co.). The solution system was stirred for 5 hours at roomtemperature and for 1 hour at 180°C. The obtained slurry was dried at100°C overnight and then calcined in air at 550°C for 2 hours.

TheMnWO4-Mn2O3 (5:4weight ratio),MnTiO3-Mn2O3 (1:1weightratio), MnWO4-Mn2O3-Na2WO4 (5:4:1 weight ratio), and MnTiO3-Mn2O3-Na2WO4 (5:5:1 weight ratio) catalysts were prepared bygrinding in a high-energy planetary ball mill (37). These four catalystswere milled for 2 hours to obtain a homogeneous mixture. The massratio of the balls to particleswas 10:1, and the rotation speedwas 320 rpm.

The practical catalyst Mn2O3-TiO2-Na2WO4/SiO2 was alsoprepared by grinding (38) in a high-energy planetary ball mill for2 hours with the same mass ratio of balls to particles and rotationspeed as for the other catalysts. By varying the contents of Mn2O3,TiO2, and Na2WO4, the mixed system led to a highly effective OCMcatalyst in the range of 8 to 15 wt % loading for Na2WO4 and 6 to28wt % loading forMn2O3 plus TiO2 (with properMn/Ti ratio to formMnTiO3), with amorphous SiO2 gel making up the balance.

CharacterizationThe catalysts were characterized by scanning electron microscopy(SEM; Hitachi S-4800) equipped with an energy-dispersive x-ray fluo-rescence spectrometer (EDX; Oxford), inductively coupled plasmaatomic emission spectrometry (ICP Thermo IRIS Intrepid II XSP),and XPS [ESCALAB 250Xi spectrometer, Al Ka, adventitious C 1s line(284.8 eV) as the reference]. The specific surface area was determinedusing standard Brunauer-Emmett-Teller theory on the N2 adsorptionisothermmeasured at−196°C on aQuantachromeAutosorb-3B instru-ment. The pore size distribution was determined using the Barrett-Joyner-Halenda (BJH) method. TPR with hydrogen (H2-TPR) wasperformed on a Quantachrome ChemBET 3000 chemisorption appa-ratus with a thermal conductivity detector (TCD). In each H2-TPRexperiment, the sample (100 mg) purged by He at 300°C for 1 hourin advancewas heated from20° to 1000°C in a gasmixture of 5%H2 inN2 (30 ml min−1) at a rate of 10°C min−1. XRD was performed on aRigaku Ultima IV diffractometer with Cu Ka radiation (35 kV and25 mA). The Raman measurements were carried out using a Ramanspectrometer (Renishaw inVia) with a 532-nm semiconductor laser asexcitation, equipped with a charge-coupled device camera enablingmicroanalysis on a sample point. The scanning range was set from80 to 2000 cm−1.

Reactivity testsThe OCM reaction was performed in a fixed bed quartz tube reactor,400 mm of straight cylindrical tubing with an internal diameter of16 mm, under atmospheric pressure. The catalyst bed was placed be-tween quartz wool plugs in the reactor. For the Ti-MWW–supportedcatalyst, 0.25 g of catalyst was loaded in the reactor with the catalyst bedthickness of approximately 10 mm; however, the density of every othercatalyst is much higher than that of the Ti-MWW–supported catalyst,and to get the identical catalyst bed thickness of about 10 mm, everyother catalyst of 1.5 g was loaded in the reactor. The reactants, CH4

(99.99%) and O2 (99.999%) with dilution of N2 (99.99%), were cofedinto the reactor using calibrated mass flow controllers. The CH4:O2:N2

molar ratio of 5:1:4 imitated the contents of a coal bed gas (50 volume%CH4 in air) with a GHSV of 8000 ml gcat.

−1 hour−1. The reaction tem-perature (that is, catalyst bed temperature) was monitored by a thermo-couple placed in the middle of the catalyst bed and was 720°, 740°, 760°,

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

780°, and 800°C. The effluent gas was analyzed with an online gaschromatograph equipped with a TCD using a 60-mDM-Plot Msieve5A column (for the separation of N2, O2, CO, and CH4) and a 30-mDM-Plot Q capillary column (for the separation of CO2, CH4, C2H4,C2H6, C3H6, and C3H8) in parallel. The CH4 conversion (CCH4) andC2H4/C2H6/C3H6/C3H8 selectivity (SC2–3) were calculated using thestandard normalization method on the basis of carbon atom balanceand defined as follows (Eqs. 1 and 2)

Conversion %ð Þ¼ ∑converted reactants

∑ðconverted reactantsÞ þ∑remaining reactants� 100 ð1Þ

Selectivity %ð Þ¼ moles of carbon atom in product N

∑ðmoles of carbon atom in reaction productsÞ � 100 ð2Þ

No formation of carbon deposit was observed. The desired productsof the OCM reaction were C2H4, C2H6, C3H6, and C3H8. Their selec-tivity was described as the C2-C3 selectivity. Reaction data at each reac-tion condition were collected after running for at least 0.5 hour.

SUPPLEMENTARY MATERIALSSupplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/3/6/e1603180/DC1Supplementary Textfig. S1. CH4 conversion and C2-C3 selectivity for the pure supports and the supported Mn2O3-Na2WO4 catalysts.fig. S2. SEM and EDX mapping images.fig. S3. XRD patterns and testing results of the 2.7Mn2O3-5.0Na2WO4/Ti-MWW and 2.7Mn2O3-5.0Na2WO4/TS-1 catalysts under different reaction conditions.fig. S4. XRD patterns and testing results for various samples with different Si:Ti molar ratio (orTi content).fig. S5. XPS spectra of various samples.fig. S6. Raman spectra of various samples.fig. S7. Testing results of the catalysts with different active components.fig. S8. Ea calculations.fig. S9. Testing results of the catalysts with different active components prepared by thegrinding method.fig. S10. H2-TPR profiles and XRD patterns.fig. S11. Effects of Mn2O3 plus TiO2, and Na2WO4 loadings on the OCM performance for theMn2O3-TiO2-Na2WO4/SiO2 catalyst.fig. S12. Stability testing of the 6.2Mn2O3-6.3TiO2-10Na2WO4/SiO2 catalyst.fig. S13. Testing results and XRD patterns for the 6.2Mn2O3-6.3TiO2-10Na2WO4/SiO2 catalystunder different reaction conditions.table S1. Detailed treatment history of some catalysts for XRD and Raman measurements.table S2. Specific surface areas and real contents of Mn, W, Na, and Ti of all usedcatalysts.table S3. Surface contents of W, Mn, Na, Ti, Si, O, and C measured by XPS for the used catalysts.table S4. CH4 conversion and C2-C3 selectivity over representative catalysts.References (38–49)

REFERENCES AND NOTES1. R. Boswell, Is gas hydrate energy within reach? Science 325, 957–958 (2009).2. E. McFarland, Unconventional chemistry for unconventional natural gas. Science 338,

340–342 (2012).3. J. H. Lunsford, Catalytic conversion of methane to more useful chemicals and fuels: A

challenge for the 21st century. Catal. Today 63, 165–174 (2000).4. J. M. Fox III, The different catalytic routes for methane valorization: An assessment of

processes for liquid fuels. Catal. Rev. 35, 169–212 (1993).5. H. Koempel, W. Liebner, Lurgi’s methanol to propylene (MTP®) report on a successful

commercialisation. Stud. Surf. Sci. Catal. 167, 261–267 (2007).

8 of 9

Page 9: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

SC I ENCE ADVANCES | R E S EARCH ART I C L E

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from

6. P. Tian, Y. Wei, M. Ye, Z. Liu, Methanol to olefins (MTO): From fundamentals tocommercialization. ACS Catal. 5, 1922–1938 (2015).

7. F. Jiao, J. Li, X. Pan, J. Xiao, H. Li, H. Ma, M. Wei, Y. Pan, Z. Zhou, M. Li, S. Miao, J. Li, Y. Zhu,D. Xiao, T. He, J. Yang, F. Qi, Q. Fu, X. Bao, Selective conversion of syngas to light olefins.Science 351, 1065–1068 (2016).

8. L. Zhong, F. Yu, Y. An, Y. Zhao, Y. Sun, Z. Li, T. Lin, Y. Lin, X. Qi, Y. Dai, L. Gu, J. Hu, S. Jin,Q. Shen, H. Wang, Cobalt carbide nanoprisms for direct production of lower olefinsfrom syngas. Nature 538, 84–87 (2016).

9. J. Gao, Y. Zheng, J.-M. Jehng, Y. Tang, I. E. Wachs, S. G. Podkolzin, Identification ofmolybdenum oxide nanostructures on zeolites for natural gas conversion. Science 348,686–690 (2015).

10. X. Guo, G. Fang, G. Li, H. Ma, H. Fan, L. Yu, C. Ma, X. Wu, D. Deng, M. Wei, D. Tan, R. Si,S. Zhang, J. Li, L. Sun, Z. Tang, X. Pan, X. Bao, Direct, nonoxidative conversion ofmethane to ethylene, aromatics, and hydrogen. Science 344, 616–619 (2014).

11. H. Schwarz, Chemistry with methane: Concepts rather than recipes. Angew. Chem. Int. Ed.50, 10096–10115 (2011).

12. G. E. Keller, M. M. Bhasin, Synthesis of ethylene via oxidative coupling of methane:I. Determination of active catalysts. J. Catal. 73, 9–19 (1982).

13. A. A. Latimer, A. R. Kulkarni, H. Aljama, J. H. Montoya, J. S. Yoo, C. Tsai, F. Abild-Pedersen,F. Studt, J. K. Nørskov, Understanding trends in C–H bond activation in heterogeneouscatalysis. Nat. Mater. 16, 225–229 (2017).

14. K. D. Campbell, E. Morales, J. H. Lunsford, Gas-phase coupling of methyl radicals duringthe catalytic partial oxidation of methane. J. Am. Chem. Soc. 109, 7900–7901 (1987).

15. T. Ito, J. H. Lunsford, Synthesis of ethylene and ethane by partial oxidation of methaneover lithium-doped magnesium oxide. Nature 314, 721–722 (1985).

16. T. LeVan, M. Che, M. Kermarec, C. Louis, J. M. Tatibouët, Structure sensitivity of the catalyticoxidative coupling of methane on lanthanum oxide. Catal. Lett. 6, 395–400 (1990).

17. S. Sugiyama, Y. Matsumura, J. B. Moffat, A comparative study of the oxides of lanthanum,cerium, praseodymium, and samarium as catalysts for the oxidative dehydrogenationof methane in the presence and absence of carbon tetrachloride. J. Catal. 139, 338–350(1993).

18. X. Fang, S. Li, J. Lin, J. Gu, D. Yan, Preparation and characterization of catalyst for oxidativecoupling of methane. J. Mol. Catal. 6, 254–262 (1992).

19. S. Arndt, T. Otremba, U. Simon, M. Yildiz, H. Schubert, R. Schomäcker, Mn–Na2WO4/SiO2

as catalyst for the oxidative coupling of methane. What is really known?Appl. Catal. A 425–426, 53–61 (2012).

20. R. Ghose, H. T. Hwang, A. Varma, Oxidative coupling of methane using catalystssynthesized by solution combustion method: Catalyst optimization and kinetic studies.Appl. Catal. A 472, 39–46 (2014).

21. V. Fleischer, R. Steuer, S. Parishan, R. Schomäcker, Investigation of the surface reactionnetwork of the oxidative coupling of methane over Na2WO4/Mn/SiO2 catalyst bytemperature programmed and dynamic experiments. J. Catal. 341, 91–103 (2016).

22. J. Wu, H. Zhang, S. Qin, C. Hu, La-promoted Na2WO4/Mn/SiO2 catalysts for the oxidativeconversion of methane simultaneously to ethylene and carbon monoxide.Appl. Catal. A 323, 126–134 (2007).

23. V. I. Alexiadis, J. W. Thybaut, P. N. Kechagiopoulos, M. Chaar, A. C. Van Veen, M. Muhler,G. B. Marin, Oxidative coupling of methane: Catalytic behaviour assessment viacomprehensive microkinetic modeling. Appl. Catal. B 150–151, 496–505 (2014).

24. C. Mesters, A selection of recent advances in C1 chemistry. Annu. Rev. Chem. Biomol. Eng.7, 223–238 (2016).

25. Y. Cui, X. Shao, M. Baldofski, J. Sauer, N. Nilius, H.-J. Freund, Adsorption, activation, anddissociation of oxygen on doped oxides. Angew. Chem. Int. Ed. 52, 11385–11387 (2013).

26. P. Schwach, M. G. Willinger, A. Trunschke, R. Schlögl, Methane coupling over magnesiumoxide: How doping can work. Angew. Chem. Int. Ed. 52, 11381–11384 (2013).

27. W. Jeon, J. Y. Lee, M. Lee, J.-W. Choi, J.-M. Ha, D. J. Suh, I. W. Kim, Oxidative coupling ofmethane to C2 hydrocarbons on the Mg–Ti mixed oxide-supported catalysts at the lowerreaction temperature: Role of surface oxygen atoms. Appl. Catal. A 464–465, 68–77 (2013).

28. B. Liu, H. M. Chen, C. Liu, S. C. Andrews, C. Hahn, P. Yang, Large-scale synthesis oftransition-metal-doped TiO2 nanowires with controllable overpotential. J. Am. Chem. Soc.135, 9995–9998 (2013).

29. D. J. Wang, M. P. Rosynek, J. H. Lunsford, Oxidative coupling of methane over oxide-supported sodium-manganese catalysts. J. Catal. 155, 390–402 (1995).

30. V. I. Alexiadis, M. Chaar, A. van Veen, M. Muhler, J. W. Thybaut, G. B. Marin, Quantitativescreening of an extended oxidative coupling of methane catalyst library. Appl. Catal. B199, 252–259 (2016).

31. Y. Luo, Y.-Q. Deng, W. Mao, X.-J. Yang, K. Zhu, J. Xu, Y.-F. Han, Probing the surfacestructure of a-Mn2O3 nanocrystals during CO oxidation by operando Ramanspectroscopy. J. Phys. Chem. C 116, 20975–20981 (2012).

32. F. Buciuman, F. Patcas, R. Cracium, D. R. T. Zahn, Vibrational spectroscopy of bulk andsupported manganese oxides. Phys. Chem. Chem. Phys. 1, 185–190 (1999).

Wang et al., Sci. Adv. 2017;3 : e1603180 9 June 2017

33. S.-f Ji, T.-c Xiao, S.-b Li, C.-z Xu, R.-l Hou, K. S. Coleman, M. L. H. Green, The relationshipbetween the structure and the performance of Na-W-Mn/SiO2 catalysts for theoxidative coupling of methane. Appl. Catal. A 225, 271–284 (2002).

34. M. Yildiz, Y. Aksu, U. Simon, K. Kailasam, O. Görke, F. Rosowski, R. Schomäcker, A. Thomas,S. Arndt, Enhanced catalytic performance of MnxOy–Na2WO4/SiO2 for the oxidativecoupling of methane using an ordered mesoporous silica support. Chem. Commun. 50,14440–14442 (2014).

35. P. Wu, T. Tatsumi, T. Komatsu, T. Yashima, A novel titanosilicate with MWW structure.I. hydrothermal synthesis, elimination of extraframework titanium, and characterizations.J. Phys. Chem. B 105, 2897–2905 (2001).

36. M. Taramasso, G. Perego, B. Notari, “Preparation of porous crystalline synthetic materialcomprised of silicon and titanium oxides,” U.S. Patent 44l050l (1983).

37. M. Dindarsafa, A. Khataee, B. Kaymak, B. Vahid, A. Karimi, A. Rahmani, Heterogeneoussono-Fenton-like process using martite nanocatalyst prepared by high energyplanetary ball milling for treatment of a textile dye. Ultrason. Sonochem. 34, 389–399(2017).

38. Z. Stansch, L. Mleczko, M. Baerns, Comprehensive kinetics of oxidative coupling ofmethane over the La2O3/CaO catalyst. Ind. Eng. Chem. Res. 36, 2568–2579 (1997).

39. Z. Liu, G. Lu, Y. Guo, Y. Wang, Y. Guo, Catalytic performance of La1-xErxCoO3 perovskitefor the deoxidization of coal bed methane and role of erbium in a catalyst.Catal. Sci. Technol. 1, 1006–1012 (2011).

40. G. Kumar, S. L. J. Lau, M. D. Krcha, M. J. Janik, Correlation of methane activation and oxidecatalyst reducibility and its implications for oxidative coupling. ACS Catal. 6,1812–1821 (2016).

41. D. Dissanayake, J. H. Lunsiord, M. P. Rosynek, Oxidative coupling of methane over oxide-supported barium catalysts. J. Catal. 143, 286–298 (1993).

42. J. A. Sofranko, J. J. Leonard, C. A. Jones, A. M. Gaffney, H. P. Withers, Catalytic oxidativecoupling of methane over sodium-promoted Mn/SiO2 and Mn/MgO. Catal. Today 3,127–135 (1988).

43. V. R. Choudhary, V. H. Rane, S. T. Chaudhari, Factors influencing activity/selectivity of La-promoted MgO catalyst prepared from La- and Mg- acetates for oxidative couplingof methane. Fuel 79, 1487–1491 (2000).

44. H. Mimoun, A. Robine, S. Bonnaudet, C. J. Cameton, Oxidative coupling of methanefollowed by ethane pyrolysis. Chem. Lett. 18, 2185–2188 (1989).

45. E. V. Kondratenko, M. Schlüter, M. Baerns, D. Linke, M. Holena, Developing catalyticmaterials for the oxidative coupling of methane through statistical analysis of literaturedata. Catal. Sci. Technol. 5, 1668–1677 (2015).

46. J. M. Thomas, W. Ueda, J. Williams, K. D. M. Harris, New families of catalysts for theselective oxidation of methane. Faraday Discuss. Chem. Soc. 87, 33–45 (1989).

47. Y. Zeng, F. T. Akin, Y. S. Lin, Oxidative coupling of methane on fluorite-structuredsamarium–yttrium–bismuth oxide. Appl. Catal. A 213, 33–45 (2001).

48. A. Palermo, J. P. H. Vazquez, R. M. Lambert, New efficient catalysts for the oxidativecoupling of methane. Catal. Lett. 68, 191–196 (2000).

49. Z. C. Jiang, C. J. Yu, X. P. Fang, S. B. Li, H. L. Wang, Oxide/support interaction and surfacereconstruction in the sodium tungstate(Na2WO4)/silica system. J. Phys. Chem. 97,12870–12875 (1993).

Acknowledgments: We acknowledge P. Wu and Y. Liu at the East China Normal University fortheir assistance in supplying Ti-MWW and TS-1 zeolite samples. We also thank R. Prinsfrom the ETH-Zürich for the helpful discussion. Funding: This work has been funded by theNational Natural Science Foundation of China (21473057, U1462129, 21322307, 21273075,and 21076083) and the “973 Program” (2011CB201403) from the Ministry of Science andTechnology of the People’s Republic of China. Author contributions: Y.L. and G.Z. conceivedthe idea for the project. P.W. and Y.W. conducted the material synthesis. P.W. performedthe structural characterizations and catalytic test. Y.L., G.Z., and P.W. discussed the catalyticresults. Y.L. and G.Z. drafted the manuscript. P.W. and G.Z. contributed equally to thiswork. All authors discussed and commented on the manuscript. Competing interests: Y.L., P.W.,and G.Z. have a patent application related to this work filed with the Chinese Patent Officeon 13 March 2017 (201710146190.1). All other authors declare that they have no competinginterests. Data and materials availability: All data needed to evaluate the conclusions inthe paper are present in the paper and/or the Supplementary Materials. Additional datarelated to this paper may be requested from the authors.

Submitted 14 December 2016Accepted 10 April 2017Published 9 June 201710.1126/sciadv.1603180

Citation: P. Wang, G. Zhao, Y. Wang, Y. Lu, MnTiO3-driven low-temperature oxidative couplingof methane over TiO2-doped Mn2O3-Na2WO4/SiO2 catalyst. Sci. Adv. 3, e1603180 (2017).

9 of 9

Page 10: SCIENCE ADVANCES| RESEARCH ARTICLE · 3-driven low-temperature oxidative coupling of methane over TiO 2-doped Mn 2O 3-Na 2WO 4/SiO 2 catalyst Pengwei Wang,1* Guofeng Zhao,2*† Yu

catalyst2/SiO4WO2-Na3O2-doped Mn2-driven low-temperature oxidative coupling of methane over TiO3MnTiO

Pengwei Wang, Guofeng Zhao, Yu Wang and Yong Lu

DOI: 10.1126/sciadv.1603180 (6), e1603180.3Sci Adv 

ARTICLE TOOLS http://advances.sciencemag.org/content/3/6/e1603180

MATERIALSSUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2017/06/05/3.6.e1603180.DC1

REFERENCES

http://advances.sciencemag.org/content/3/6/e1603180#BIBLThis article cites 48 articles, 5 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science AdvancesYork Avenue NW, Washington, DC 20005. The title (ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 NewScience Advances

Copyright © 2017, The Authors

on Novem

ber 10, 2020http://advances.sciencem

ag.org/D

ownloaded from