7
Catalysis Science & Technology PAPER Cite this: DOI: 10.1039/c9cy02342a Received 19th November 2019, Accepted 22nd January 2020 DOI: 10.1039/c9cy02342a rsc.li/catalysis Nonmetallic boron nitride embedded graphitic carbon catalyst for oxidative dehydrogenation of ethylbenzeneJian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu * Oxidative dehydrogenation of ethylbenzene to styrene featured with a non-equilibrium-limit is an alternative to the direct dehydrogenation process, but often exhibits low styrene selectivity caused by deep-oxidation over metal oxide catalysts. Herein, we report a nonmetallic BN embedded graphitic carbon catalyst which was synthesized by a one-step co-pyrolysis method using boric acid, urea and ammonium iron citrate as the reactants. The obtained catalyst exhibits a high styrene selectivity of 94%. The existence of abundant carbonyl groups and BO species could correspondingly contribute more reaction sites and extra oxygen adsorption sites. A higher styrene formation rate of 5.8 mmol g cat 1 h 1 is thus achieved for our BN embedded graphitic carbon catalyst, whereas this value is 3.4 mmol g cat 1 h 1 for sole carbon nanotubes and 0.2 mmol g cat 1 h 1 for BN. Such promoted reactivity for oxidative dehydrogenation of ethylbenzene is derived from the synergetic and collaborative effect of the nonmetallic BN and graphitic carbon catalyst. 1 Introduction Oxidative dehydrogenation (ODH) of ethylbenzene (EB) to styrene (ST) is considered as an alternative route to the conventional direct dehydrogenation process since the exothermic and irreversible reaction nature of ODH can contribute to a higher EB conversion at relatively lower temperatures (400500 °C). Meanwhile, the direct dehydrogenation process catalysed by a potassium-promoted iron oxide catalyst was usually operated at higher temperatures (600650 °C) due to its endothermic nature. Thus, the catalyst often encounters inevitable deactivation caused by serious coke deposition. In this case, energy- intensive consumption of superheated steam is required to remove carbon deposits from the catalyst surface. 1,2 Despite many advantages of ODH, the ST selectivity is often lower than that of direct dehydrogenation due to the oxygen insertion and deep-oxidation of EB or ST to undesired products on metal oxide-based catalysts. 2,3 As a response to this problem, metal-free carbonaceous materials 4,5 have been proposed instead of metal-based catalysts such as vanadium oxides, 3,6 cerium oxides, 2 and metal phosphates 7 with the aim of suppressing deep- oxidation. Carbon-based catalysts with tunable surface acidity/basicity groups and electron density show higher ST selectivity compared to metal-based catalysts. 8 For example, carbon molecular sieves give a high ST selectivity (>90%) at a high EB conversion (>80%). 9 By thoroughly investigating the structure and catalytic performance of amorphous activated carbon and carbon nanotubes (CNTs) in this reaction system, the carbonyl groups on the surface are considered as the active sites to activate the reaction of ethylbenzene to styrene. 4,10 However, activated carbon with abundant micropores and low crystallinity usually encounters unavoidable deactivation due to coke formation in the micropores and gasification of the carbon substrate under oxidative conditions. 8,11 In contrast, mesoporous CNTs with a highly crystalline structure can overcome the aforementioned problems and show good catalytic stability. 12,13 But high graphitization always means few defect sites and low concentration of active surface oxygen groups, resulting in a low styrene formation rate. Oxidative post-treatments like HNO 3 refluxing, O 2 pre-oxidation or heteroatom modification can be employed to enrich the oxygen groups of CNTs, but non-selectivity and low efficiency for specific active oxygen groups often occur. 14 Fabrication of hybrid catalysts with the specific function of different components is often employed to improve the overall performance of catalysts. 15,16 For example, a sandwich structured MgOrGO hybrid catalyst showed enhanced ethylbenzene direct dehydrogenation reactivity. The special sandwich structure facilitated the exposure of active sites and Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2020 State Key Laboratory of Fine Chemicals, School of Chemical Engineering, Dalian University of Technology, Dalian 116024, P. R. China. E-mail: [email protected]; Tel: +86 411 84986112 Electronic supplementary information (ESI) available. See DOI: 10.1039/ c9cy02342a Published on 24 January 2020. Downloaded by Universite Paris Descartes on 2/12/2020 12:15:14 PM. View Article Online View Journal

Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

CatalysisScience &Technology

PAPER

Cite this: DOI: 10.1039/c9cy02342a

Received 19th November 2019,Accepted 22nd January 2020

DOI: 10.1039/c9cy02342a

rsc.li/catalysis

Nonmetallic boron nitride embedded graphiticcarbon catalyst for oxidative dehydrogenation ofethylbenzene†

Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu *

Oxidative dehydrogenation of ethylbenzene to styrene featured with a non-equilibrium-limit is an

alternative to the direct dehydrogenation process, but often exhibits low styrene selectivity caused by

deep-oxidation over metal oxide catalysts. Herein, we report a nonmetallic BN embedded graphitic carbon

catalyst which was synthesized by a one-step co-pyrolysis method using boric acid, urea and ammonium

iron citrate as the reactants. The obtained catalyst exhibits a high styrene selectivity of 94%. The existence

of abundant carbonyl groups and BO species could correspondingly contribute more reaction sites and

extra oxygen adsorption sites. A higher styrene formation rate of 5.8 mmol gcat−1 h−1 is thus achieved for

our BN embedded graphitic carbon catalyst, whereas this value is 3.4 mmol gcat−1 h−1 for sole carbon

nanotubes and 0.2 mmol gcat−1 h−1 for BN. Such promoted reactivity for oxidative dehydrogenation of

ethylbenzene is derived from the synergetic and collaborative effect of the nonmetallic BN and graphitic

carbon catalyst.

1 Introduction

Oxidative dehydrogenation (ODH) of ethylbenzene (EB) tostyrene (ST) is considered as an alternative route to theconventional direct dehydrogenation process since theexothermic and irreversible reaction nature of ODH cancontribute to a higher EB conversion at relatively lowertemperatures (400–500 °C). Meanwhile, the directdehydrogenation process catalysed by a potassium-promotediron oxide catalyst was usually operated at highertemperatures (600–650 °C) due to its endothermic nature.Thus, the catalyst often encounters inevitable deactivationcaused by serious coke deposition. In this case, energy-intensive consumption of superheated steam is required toremove carbon deposits from the catalyst surface.1,2 Despitemany advantages of ODH, the ST selectivity is often lowerthan that of direct dehydrogenation due to the oxygeninsertion and deep-oxidation of EB or ST to undesiredproducts on metal oxide-based catalysts.2,3

As a response to this problem, metal-free carbonaceousmaterials4,5 have been proposed instead of metal-basedcatalysts such as vanadium oxides,3,6 cerium oxides,2 andmetal phosphates7 with the aim of suppressing deep-

oxidation. Carbon-based catalysts with tunable surfaceacidity/basicity groups and electron density show higher STselectivity compared to metal-based catalysts.8 For example,carbon molecular sieves give a high ST selectivity (>90%) at ahigh EB conversion (>80%).9 By thoroughly investigating thestructure and catalytic performance of amorphous activatedcarbon and carbon nanotubes (CNTs) in this reaction system,the carbonyl groups on the surface are considered as theactive sites to activate the reaction of ethylbenzene tostyrene.4,10 However, activated carbon with abundantmicropores and low crystallinity usually encountersunavoidable deactivation due to coke formation in themicropores and gasification of the carbon substrate underoxidative conditions.8,11 In contrast, mesoporous CNTs with ahighly crystalline structure can overcome the aforementionedproblems and show good catalytic stability.12,13 But highgraphitization always means few defect sites and lowconcentration of active surface oxygen groups, resulting in alow styrene formation rate. Oxidative post-treatments likeHNO3 refluxing, O2 pre-oxidation or heteroatom modificationcan be employed to enrich the oxygen groups of CNTs, butnon-selectivity and low efficiency for specific active oxygengroups often occur.14

Fabrication of hybrid catalysts with the specific functionof different components is often employed to improve theoverall performance of catalysts.15,16 For example, a sandwichstructured MgO–rGO hybrid catalyst showed enhancedethylbenzene direct dehydrogenation reactivity. The specialsandwich structure facilitated the exposure of active sites and

Catal. Sci. Technol.This journal is © The Royal Society of Chemistry 2020

State Key Laboratory of Fine Chemicals, School of Chemical Engineering, Dalian

University of Technology, Dalian 116024, P. R. China.

E-mail: [email protected]; Tel: +86 411 84986112

† Electronic supplementary information (ESI) available. See DOI: 10.1039/c9cy02342a

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article OnlineView Journal

Page 2: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2020

the alkaline MgO promoted the desorption of styrene with ahigher ST selectivity.16

Boron nitride, with superior anti-oxidation ability andexcellent thermal conductivity, has been demonstrated to bean efficient catalyst in the oxidative dehydrogenation of lightalkanes to olefins with robust stability.17,18 And the BOspecies on the BN edges are considered to play an importantrole in the activation of alkane and oxygen molecules.19–24 Inthe case of the oxidative dehydrogenation of ethylbenzene,amorphous carbon (8–15 wt%) doped BN nanosheets arefound to effectively promote the formation of styrene. Thecarbonyl groups and the oxygen-terminated armchair edgesare supposed to jointly serve as active sites for this reactionand the BN edges could prevent the over-oxidation of theadsorbed species caused by highly reactive radicals.25

Besides, some theoretical calculations also point out that Bspecies can reinforce the activation ability of molecularoxygen in the oxidative dehydrogenation reaction.26 By takinginto account the advantage of the high selectivity of carbon-based catalysts and the oxygen-rich nature of BN, it'sreasonably expected that a nonmetallic hybrid catalystcontaining graphitic carbon and BN has great potential toachieve boosted reactivity for oxidative dehydrogenation.

Herein, we report a BN embedded mesoporous graphiticcarbon (BN-GC) nonmetallic hybrid catalyst synthesized by afacile one-step co-pyrolysis method without further complexoxidative post-treatments. BN-GC hybrids with abundantoxygen functional groups and BO species exhibit improvedcatalytic performance in the oxidative dehydrogenation ofethylbenzene to styrene with a 5.8 mmol gcat

−1 h−1 STformation rate compared to sole carbon nanotubes (3.4mmol gcat

−1 h−1) and BN (0.2 mmol gcat−1 h−1).

2 Experimental2.1 Chemical reagents and synthetic methods

All chemical reagents were employed as received without anypurification. Urea (99.0%), ammonium ironIJIII) citrate (CP)and boric acid (99.5%) were purchased from SinopharmChemical Reagent Co. Ltd.

Urea, boric acid and ammonium ironIJIII) citrate with molarratios of 120 :2 : 1, 120 :5 : 1, 120 :10 :1, and 120 :20 :1 were mixedinto pale yellow powder and then pyrolyzed in a tube furnace at900 °C in an Ar atmosphere (40 mL min−1) for 2 h. Then the as-obtained black powders were washed with 6 M HCl to removethe Fe residues, and then washed with deionized water to neutral.After drying at 50 °C for 24 h, the obtained hybrids were namedBN-GC-0.2, BN-GC-0.5, BN-GC-1, and BN-GC-2, respectively.

For comparison, one pure BN sample and one CNTsample were prepared. For BN preparation, urea and boricacid with a molar ratio of 6 : 1 were mixed and then calcinedin a tube furnace at 1000 °C in a N2 atmosphere for 1 h. Thenthe white powder was washed with hot deionized water (80°C) for 4 h and dried at 50 °C for 24 h to obtain sample BN.The CNT sample was obtained by purifying commercial

multiwall CNTs with acid washing with 6 M HCl for 24 h anddeionized water washing.

In order to demonstrate the effect of co-pyrolysis, onesample (named BN-CNT-mix) was prepared by mechanicallymixing the CNT and BN samples, in a mass ratio of CNTpowders and BN being 7 : 3 according to thermogravimetricanalysis.

2.2 Catalyst characterization

X-ray powder diffraction (XRD) measurements were performedon an X'Pert-3 powder diffractometer using Cu Kα radiation (λ= 0.15406 nm). Fourier transform infrared (FTIR) spectra werecollected using a Thermo Scientific Nicolet 6700 spectrometerby dispersing samples in KBr pellets. Transmission electronmicroscopy (TEM) images were recorded on a Tecnai G2 F30 S-Twin, operating at an accelerating voltage of 300 kV. Nitrogenadsorption isotherms were measured at 77 K with a Tristar3000 adsorption analyser (Micromeritics). All samples weredegassed at 200 °C for 6 h until the pressure was below 5 Pabefore the measurements. The Brunauer–Emmett–Teller (BET)method was used to calculate the specific surface areas (SBET).Total pore volumes were calculated from the amount adsorbedat a relative pressure P/P0 of 0.99. X-ray photoelectronspectroscopy (XPS) data were obtained with a PHI 5000Versaprobe spectrometer equipped with an Al Kα X-ray source.The binding energies of elements were calibrated using the C1s photoelectron peak at 284.6 eV. Thermogravimetric analysis(TGA) was performed from 40 to 800 °C with a heating rate of10 °C min−1 under an air flow, using a STA 449 F3 Jupiterthermogravimetric analyser (NETZSCH).

2.3 Catalytic testing

Oxidative dehydrogenation of ethylbenzene was carried outin a fixed-bed quartz reactor (i.d. 8 mm, length 420 mm) inthe range of 350–500 °C under atmospheric pressure with 50mg of catalyst. The reactant (2.8% EB, nitrogen as balance,total flow rate 10 mL min−1) was then fed to the reactor froma saturator kept at 40 °C. The reactants and products wereanalysed using an on-line gas chromatograph (GC7900,Techcomp) equipped with three columns: a Porapak Q 2 anda molecular sieve 5A column for the permanent gases and aPEG capillary column for the hydrocarbons, coupled to aTCD and a FID, respectively. The ethylbenzene conversion,styrene selectivity and styrene formation rate were calculatedaccording to the following equations:

EB Conversion = (FEBin − FEBout)/FEBin

ST Selectivity = FSTout/(FSTout + FBZout + FTOLout + FBAout + FCOxout)

ST Formation rate = (fraction of EB converted to ST (mmol))/(weight of the catalyst (g) × time (h))

Catalysis Science & TechnologyPaper

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article Online

Page 3: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

Catal. Sci. Technol.This journal is © The Royal Society of Chemistry 2020

FEB, FST, FBZ, FTU, FCOxand FBA denote the concentrations of

ethylbenzene, styrene, benzene, toluene, COx andbenzaldehyde, respectively.

The carbon balance was checked by comparing thenumber of moles of carbon in the outlet stream to thenumber of moles of carbon in the feed and the carbonbalance was close to 95–100% in all runs. The EB conversionof the blank experiment with quartz sand was lower than 1%in the whole test temperature range which can be ignored (asshown in Table S1†).

3 Results and discussion3.1 Structure characterization

The BN-GC hybrid catalysts were obtained by simple one-stepco-pyrolysis of the mixture of boric acid, urea andammonium ironIJIII) citrate in an Ar atmosphere. Fig. 1Ashows the XRD patterns of the freshly prepared hybridcatalysts and the control samples. The control sample BNshowed two characteristic peaks at 26.3° and 42.3°, ascribedto the (002) and (100) planes of hexagonal boron nitride (h-BN) (ICDD card no. 00-034-0421); whereas CNTs displayedtypical peaks at 26.0° and 42.8°, ascribed to the graphiticstructure of carbon (ICDD card no. 00-041-1487). For the BN-GC hybrid catalysts, due to the overlapping in the (002)planes of carbon and BN, the peak observed at 2θ = 26.0°could be assigned to the (002) plane of either carbon orhexagonal boron nitride. However, the peaks around 2θ = 77°were close to the (110) plane of the carbon material,suggesting that the main body of the BN-GC sample wascarbon. The sharp (002) peak of BN-GC around 26.0°indicated the highly crystalline structure of all the samples.

No diffraction patterns related to B2O3 were observed.Besides, there were also no obvious diffraction peaks of ironspecies detected in the hybrid samples, indicating thatalmost all the iron species were removed.

The Fourier transform infrared (FTIR) spectra of BN-GC-0.5 and BN (Fig. 1B) both exhibited two characteristicabsorption bands at 1380 and 790 cm−1, respectively, whichwere ascribed to the in-plane B–N stretching vibrations andout-of-plane B–N–B bending vibrations of BN,27 suggestingthe existence of BN in BN-GC. Besides, a broad peak at 3400cm−1 was attributed to the O–H stretching vibrations or watermolecules. The weak peak at 1590 cm−1 was assigned toCN absorption,27 indicating that carbon was doped intothe BN matrix or N was introduced into the carbon structure.

The nitrogen sorption isotherms (Fig. 1C) of the BN-GCsamples were type-IV isotherms with an H2 hysteresis loop,indicating their accessible mesoporous characteristics.Typically, the obtained BN-GC-0.5 has a surface area of 311m2 g−1 and a pore volume of 0.61 cm3 g−1. Such amesoporous structure is beneficial for the diffusion of thereactants and products, suppressing the coke formation. Thenitrogen sorption isotherms of the CNTs and BN are shownin Fig. S1,† and the specific surface area and pore volume arelisted in Table 1. These results suggest that all the preparedsamples have a similar specific surface area and mesoporousstructure.

The transmission electron microscopy (TEM) image(Fig. 1D) of BN-GC-0.5 showed a typical structure of graphiticcarbon with a lattice spacing of 0.334 nm (Fig. 1D, inset). Inthe co-pyrolysis process, we speculated that the dehydrationof H3BO3 to B2O3 and the pyrolysis of urea to CN gases(C2N2

+, C3N2+, and C3N3

+)28,29 might occur at the early stage.B2O3 may react with the decomposed product of urea to formBN. The metallic iron ions might be reduced to metallic ironby CN gases and then deposited on the BN substrate, whichcatalysed the growth of graphitic carbons. Finally, the BNembedded graphitic carbon hybrid structure was formed.30

The chemical structure of the BN-GC-0.5 sample wasfurther investigated by X-ray photoelectron spectroscopy(XPS) measurement. The survey spectra (Fig. 2A) revealed thatB, C, N and O elements were the primary surface species andthe relative atomic contents were respectively around 10.0%,72.8%, 9.9%, and 7.0%. The Fe signal around 710 eV wasnegligible, indicating the complete removal of the Fe residue.This result was in accordance with that of XRD. As Fig. 2Bshows, the B 1s spectrum was deconvoluted into two modesat ∼190.3 and ∼192.1 eV. The former peak was assigned toB–N and the latter sub-peak was attributed to B–O,potentially bonding on the edges.18 The C 1s spectrum(Fig. 2C) was deconvoluted into four peaks at around ∼284.6,∼286.0, ∼287.6 and ∼289.5 eV, corresponding to the typicalC–C/CC of graphitic carbon, CN, CO and COORspecies.30 As seen in Fig. 2D, the N 1s spectrum was dividedinto four typical peaks at ∼397.8, ∼398.6, ∼399.6 and ∼400.6eV, ascribed to N–B, and pyridinic, pyrrolic, and graphiticnitrogen, respectively.30,31 N–B was the main component.

Fig. 1 (A) X-ray diffraction patterns of the BN-GC samples, CNTs andBN. (B) FTIR spectra of BN-GC-0.5, CNTs and BN. (C) N2 sorptionisotherms and pore size distributions of BN-GC. (D) TEM image of BN-GC-0.5 (inset is the high resolution image of the graphitic structure).

Catalysis Science & Technology Paper

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article Online

Page 4: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2020

These results suggested that N was doped into the graphiticframework and some O bonded to B on the edges of BN overBN-GC during the one-step co-pyrolysis process.

3.2 Catalytic performance

To optimize the B content of BN-GC and the reactionconditions, the catalytic performances of the as-preparedcatalysts were evaluated. Fig. S2† shows the conversion-selectivity profiles of the BN-GC catalysts with different Bcontents. By altering the B contents, BN-GC-0.5 displayed thebest catalytic performance among the as-prepared catalystsdue to its highest ST selectivity at the same EB conversionunder identical reaction conditions. Further, the reactionconditions were optimized by changing the ratio of O2/EBand the reaction temperature at a constant space velocity(200 mL g−1 min−1) over the BN-GC-0.5 catalyst. The catalyticperformance dependence for various ratios of O2/EB at 400°C on time on stream is shown in Fig. S3A.† As the ratio of

O2/EB increased from 0.5 to 1.5, the tendency of EBconversion was ascending first and then descending. At O2/EB = 1, the maximum value of EB conversion (26%) wasobtained. As the reaction temperatures increased from 400°C to 450 °C at O2/EB = 1, an optimal catalytic performancewith 41% EB conversion and 94% ST selectivity was achievedat 425 °C (Fig. S3B†).

Fig. 3A shows the comparison of the catalytic performanceamong the as-prepared BN-GC-0.5, CNT and BN catalysts atdifferent temperatures. A relatively low catalytic reactivity(∼0.64 mmol gcat

−1 h−1) was observed for BN even at 465 °C,which was consistent with a previous report.25 BN-GC-0.5exhibited a higher EB conversion than CNTs in the wholetested temperature range and the ST selectivity almostmaintained the same level (>90%). Fig. 3B and Table 1present the stable catalytic performance of the as-preparedcatalysts at 425 °C after 8 h on stream. Among all thecatalysts, sample BN-GC-0.5 exhibited an EB conversion of41%, which was much higher than those for CNTs (23%) andBN (2%). The ST formation rate of the BN-GC-0.5 catalyst wascalculated to be 5.8 mmol gcat

−1 h−1, which was higher thanthose for CNTs (3.4 mmol gcat

−1 h−1) and BN (0.2 mmol gcat−1

h−1). In addition, the mechanically mixed sample (BN-CNT-mix) was also evaluated under the same reaction conditionsand offered a similar conversion rate (3.2 mmol gcat

−1 h−1)compared to CNTs, suggesting a possible synergistic effect of

Table 1 Summary of the physical and chemical properties of different catalysts and the calculated ST formation rates at 425 °C

CatalystsSBET

a

(m2 g−1)Vtotal

b

(cm3 g−1)Dp

c

(nm)EB con.(%)

ST Sel.(%)

ST formation rate

mmol gcat−1 h−1 μmol m−2 h−1

BN-GC-0.2 358 0.51 6.7 36 94 5.1 14.2BN-GC-0.5 311 0.61 8.7 41 94 5.8 18.6BN-GC-1 252 0.45 8.4 33 94 4.7 18.7BN-GC-2 186 0.32 8.4 35 91 4.8 25.8BN-CNT-mix — — — 23 94 3.2 —CNTs 207 0.80 17.5 24 95 3.4 16.4BN 252 0.16 2.4d 2 80 0.2 0.8

a The specific surface areas were calculated by the BET method. b The total volumes were estimated based on the adsorbed amount at arelative pressure of 0.99. c The pore sizes were calculated from the adsorption branch using the BJH model. d The pore sizes were calculatedfrom the adsorption branch using the DFT method.

Fig. 2 XPS spectra of the BN-GC-0.5 sample: (A) survey, (B) B 1s, (C) C1s, and (D) N 1s.

Fig. 3 Oxidative dehydrogenation of ethylbenzene over differentcatalysts: (A) influence of temperature on conversion and selectivity.Reaction conditions: 50 mg, 2.8% EB with N2 balance, O2/EB = 1, 200mL g−1 min−1. (B) Stable catalytic performance, data obtained after 8 hon stream. Reaction conditions: 50 mg, 2.8% EB with N2 balance, O2/EB = 1, 200 mL g−1 min−1, T = 425 °C.

Catalysis Science & TechnologyPaper

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article Online

Page 5: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

Catal. Sci. Technol.This journal is © The Royal Society of Chemistry 2020

the two components in the hybrid structure. Regarding thestyrene selectivity, it was higher than 90% over all the testedcatalysts except BN with 80% even at a low EB conversionlevel, indicating the low reactivity of boron nitride itself.Moreover, a comparative summary about oxidativedehydrogenation of ethylbenzene is listed in Table 2.Compared with the reported CNT catalysts, the as-preparedBN-GC catalyst showed a higher styrene selectivity andformation rate.

In order to further understand the enhanced catalyticperformance over the BN-GC catalysts, the spent catalystswere investigated. It can be seen that the spent BN-GC-0.5catalyst retained all the XRD reflections (Fig. S4†) of the freshsample, indicating that the general structure was wellmaintained after the ODH reaction. Thermogravimetricanalysis (Fig. S5†) was used to compare the weight loss of thefresh and spent catalysts. The identical final weight loss (72wt%) suggested that there was no obvious coke deposition onBN-GC-0.5. XPS was employed to explore variations of surfacecompositions, as shown in Fig. 4 and Table S2.† Theconcentration of the total surface oxygen in the fresh BN-GC-0.5 and CNT catalysts was estimated from the spectra to be7.0 at% and 1.9 at%, respectively. After the reaction, thesewere increased to 9.3 at% and 2.0 at%, respectively, indicatingthe regeneration of oxygen on the surface of the catalyst underthe reaction conditions.25 The deconvolution of the O 1sspectra of the fresh and spent CNTs (Fig. 4A) revealed threedifferent chemical environments of oxygen species, whichcould be respectively assigned to phenol groups (C–OH,∼533.7eV), carboxylic acid, anhydride, lactone and ester groups(OC–O, ∼532.6 eV), and carbonyl groups (CO, ∼531.1eV).31 As for CNT-BN-0.5 (Fig. 4B), it was divided into four peaksin addition to the B–O groups, located at ∼532.2 eV.23

Generally, the basic carbonyl groups are considered as activesites in the oxidative dehydrogenation reaction ofethylbenzene.4,10 As shown in Fig. S6 and S7,† a nearly linearrelationship was obtained between the ST formation ratesand the concentration of carbonyl groups, suggesting theimportant role of carbonyl groups. However, no obviousrelationship is found between the ST formation rate and B–Ogroups. The activation energies of BN-GC-0.5 and CNTs basedon the Arrhenius plot were further measured to be 66 kJmol−1 and 68 kJ mol−1, respectively (Fig. S8†), indicating thesame active sites of these catalysts. Therefore, the higherreaction rate observed for the BN-GC hybrid catalysts thanthat for CNTs was attributed to the concentration of carbonyl

groups. It is worth pointing out that the oxygen moleculescan be activated into peroxo species by the rich π electrons indefect sites and these can accumulate in the carbon pores,causing the formation of covalent carbon–oxygen bonds suchas carbonyl or quinone groups.37 This could also contributeto the total apparent reactivity. After the reaction, the totalcontent of CO was increased to 1.66% on the BN-GC-0.5hybrid catalyst. However, it almost remained unchanged onCNTs during the ODH process, probably due to the highcrystallinity with fewer defect sites of the CNTs. In addition,as given in Fig. S9,† the doping of nitrogen atom into thecarbon matrix can also increase the electron density andstrengthen the basicity of carbon materials. This may bebeneficial for the desorption of products and suppression ofover-oxidation, leading to an improvement in catalyticselectivity for styrene.38,39

On the other hand, the content of B–O was also increasedfrom 1.9% to 2.7% on BN-GC-0.5 and no B2O3 phase wasdetected in the XRD data (Fig. S4†) after the ODH process,suggesting the emergence of oxygen functionalization in BN.This observation was consistent with the deconvolution ofthe B 1s spectra (Fig. S10†). These results implied that theB–O groups might be involved in the ODH reactionnetwork.25,40 It was reported that the oxygen-containingboron species on the surface of boron nitride could activatemolecular oxygen in the ODH of light alkanes.18 Andethylbenzene molecules could be activated by BO speciessites and these sites could be regenerated by CO2.

40 In ourcase, as Fig. S11† shows, the reaction order of oxygenobserved over BN-GC-0.5 was 0.3, which was slightly higherthan that of CNTs (0.2), suggesting that the presence of BOspecies in BN-GC-0.5 could adsorb oxygen molecules39 and itcould possibly participate in the reaction.41 Unfortunately,

Table 2 Comparison of the catalytic performance of CNT catalysts reported in the literature

Catalysts O2/EB SV (mL g−1 min−1) T (°C) EB con. (%) ST Sel. (%) ST formation rate (mmol gcat−1 h−1) Ref.

BN-GC-0.5 1 200 425 41 94 5.8 This workCNTs 1 500 550 64 54 9.3 32CNFs 1 50 400 59 85 1.6 33CNTs 0.5 200 400 29 94 3.4 34oCNTs 1 55 400 46 91 1.4 35CNT monoliths 2.5 100 400 50 90 3.5 36Carbon-doped BN nanosheets 4 200 500 50 88 1.3 25

Fig. 4 The deconvolution of the O 1s spectra: (A) fresh and spentp-CNTs and (B) fresh and spent BN-GC-0.5.

Catalysis Science & Technology Paper

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article Online

Page 6: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2020

there was no concrete evidence for oxygen species from thecurrent experimental results. Thus, we speculated that theenhancement of the catalytic activity over the BN-GC catalystswas probably due to the increased concentration of carbonylgroups, and the existence of BO species might facilitate theadsorption of oxygen and possibly be involved in thereaction. These results revealed the existence of a synergisticeffect between carbon and BN in the oxidativedehydrogenation of ethylbenzene.

4 Conclusions

In summary, highly crystalline nonmetallic BN embeddedmesoporous graphitic carbon was synthesized by a one-stepco-pyrolysis method without further complex oxidative post-treatments and used as a novel nonmetallic catalyst for theoxidative dehydrogenation of ethylbenzene. Sample BN-GC-0.5 showed the best catalytic performance (41% EBconversion and 94% ST selectivity) with a 5.8 mmol gcat

−1 h−1

ST formation rate compared to sole CNTs (3.4 mmol gcat−1

h−1) and BN (0.2 mmol gcat−1 h−1). The mesopores allowed

easy desorption of styrene, consequently avoiding the deep-oxidation and the generation of by-products. The hybrid BN-GC structure could provide more oxygen functional groupson the surface of the graphitic carbon thus to promote theconversion of ethylbenzene, and the BO species at the BNedges could facilitate the adsorption of oxygen molecules toaccelerate the reaction cycle.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This study was supported by the State Key Program of theNational Natural Science Foundation of China (21733002), theJoint Sino-German Research Project (21761132011), and theCheung Kong Scholars Program of China (T2015036).

Notes and references

1 N. Mimura, I. Takahara, M. Saito, T. Hattori, K. Ohkumaand M. Ando, Catal. Today, 1998, 45, 61–64.

2 J. Xu, B. Xue, Y.-M. Liu, Y.-X. Li, Y. Cao and K.-N. Fan, Appl.Catal., A, 2011, 405, 142–148.

3 W. Oganowski, J. Hanuza and L. Kepiński, Appl. Catal., A,1998, 171, 145–154.

4 M. F. R. Pereira, J. J. M. Orfao and J. L. Figueiredo, Appl.Catal., A, 1999, 184, 153–160.

5 G. Mestl, N. I. Maksimova, N. Keller, V. V. Roddatis and R.Schlogl, Angew. Chem., Int. Ed., 2001, 40, 2066–2068.

6 W. S. Chang, Y. Z. Chen and B. L. Yang, Appl. Catal., A,1995, 124, 221–243.

7 F. M. Bautista, J. M. Campelo, D. Luna, J. M. Marinas, R. A.Quirós and A. A. Romero, Appl. Catal., B, 2007, 70, 611–620.

8 W. Qi and D. Su, ACS Catal., 2014, 4, 3212–3218.

9 F. Cavani and F. Trifirò, Appl. Catal., A, 1995, 133, 219–239.10 W. Qi, W. Liu, B. Zhang, X. Gu, X. Guo and D. Su, Angew.

Chem., Int. Ed., 2013, 52, 14224–14228.11 D. Chen, A. Holmen, Z. Sui and X. Zhou, Chin. J. Catal.,

2014, 35, 824–841.12 M. F. R. Pereira, J. L. Figueiredo, J. J. M. Órfão, P. Serp, P.

Kalck and Y. Kihn, Carbon, 2004, 42, 2807–2813.13 W. Qi, P. Yan and D. S. Su, Acc. Chem. Res., 2018, 51,

640–648.14 K. A. Wepasnick, B. A. Smith, K. E. Schrote, H. K. Wilson,

S. R. Diegelmann and D. H. Fairbrother, Carbon, 2011, 49,24–36.

15 H. Liu, J. Diao, Q. Wang, S. Gu, T. Chen, C. Miao, W. Yangand D. Su, Chem. Commun., 2014, 50, 7810–7812.

16 J. Diao, Y. Zhang, J. Zhang, J. Wang, H. Liu and D. S. Su,Chem. Commun., 2017, 53, 11322–11325.

17 J. T. Grant, C. A. Carrero, F. Goeltl, J. Venegas, P. Mueller,S. P. Burt, S. E. Specht, W. P. McDermott, A. Chieregato andI. Hermans, Science, 2016, 354, 1570–1573.

18 L. Shi, D. Wang, W. Song, D. Shao, W.-P. Zhang and A.-H.Lu, ChemCatChem, 2017, 9, 1788–1793.

19 J. Tian, J. Tan, M. Xu, Z. Zhang, S. Wan, S. Wang, J. Lin andY. Wang, Sci. Adv., 2019, 5, 8063.

20 Y. Zhou, J. Lin, L. Li, X. Pan, X. Sun and X. Wang, J. Catal.,2018, 365, 14–23.

21 L. Shi, Y. Wang, B. Yan, W. Song, D. Shao and A.-H. Lu,Chem. Commun., 2018, 54, 10936–10946.

22 L. Shi, D. Wang and A.-H. Lu, Chin. J. Catal., 2018, 39,908–913.

23 L. Shi, B. Yan, D. Shao, F. Jiang, D. Wang and A.-H. Lu, Chin.J. Catal., 2017, 38, 389–395.

24 R. Huang, B. Zhang, J. Wang, K.-H. Wu, W. Shi, Y. Zhang, Y.Liu, A. Zheng, R. Schlögl and D. S. Su, ChemCatChem,2017, 9, 3293–3297.

25 F. Guo, P. Yang, Z. Pan, X. N. Cao, Z. Xie and X. Wang,Angew. Chem., Int. Ed., 2017, 56, 8231–8235.

26 S. Shang, P.-P. Chen, L. Wang, Y. Lv, W.-X. Li and S. Gao,ACS Catal., 2018, 8, 9936–9944.

27 J. Wu, L. Wang, B. Lv and J. Chen, ACS Appl. Mater.Interfaces, 2017, 9, 14319–14327.

28 Y. Yan, J. Miao, Z. Yang, F. X. Xiao, H. B. Yang, B. Liu and Y.Yang, Chem. Soc. Rev., 2015, 44, 3295–3346.

29 J. Liu, T. Zhang, Z. Wang, G. Dawson and W. Chen, J. Mater.Chem., 2011, 21, 14398–14401.

30 B. He, W.-C. Li, Y. Zhang, X.-F. Yu, B. Zhang, F. Li and A.-H.Lu, J. Mater. Chem. A, 2018, 6, 24194–24200.

31 J. L. Figueiredo and M. F. R. Pereira, Catal. Today, 2010, 150,2–7.

32 D. S. Su, N. Maksimova, J. J. Delgado, N. Keller, G. Mestl,M. J. Ledoux and R. Schlögl, Catal. Today, 2005, 102–103,110–114.

33 T.-J. Zhao, W.-Z. Sun, X.-Y. Gu, M. Rønning, D. Chen, Y.-C.Dai, W.-K. Yuan and A. Holmen, Appl. Catal., A, 2007, 323,135–146.

34 J. Zhang, D. Su, A. Zhang, D. Wang, R. Schlögl and C.Hébert, Angew. Chem., 2007, 119, 7460–7464.

Catalysis Science & TechnologyPaper

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article Online

Page 7: Catalysis Science & Technology - dlut.edu.cnanhuilu.dlut.edu.cn/ky/article/shengjian2020.pdf · 2020. 3. 28. · Jian Sheng, Bing Yan, Bin He, Wen-Duo Lu, Wen-Cui Li and An-Hui Lu

Catal. Sci. Technol.This journal is © The Royal Society of Chemistry 2020

35 N. V. Qui, P. Scholz, T. Krech, T. F. Keller, K. Pollok and B.Ondruschka, Catal. Commun., 2011, 12, 464–469.

36 J. Zhang, R. Wang, E. Liu, X. Gao, Z. Sun, F. S. Xiao, F.Girgsdies and D. S. Su, Angew. Chem., Int. Ed., 2012, 51,7581–7585.

37 F. Atamny, J. Blöcker, A. Dübotzky, H. Kurt, O. Timpe, G.Loose, W. Mahdi and R. Schlögl, Mol. Phys., 1992, 76,851–886.

38 Z. Zhao and Y. Dai, J. Mater. Chem. A, 2014, 2,13442–13451.

39 W. Liu, C. Wang, F. Herold, B. J. M. Etzold, D. Su and W. Qi,Appl. Catal., B, 2019, 258, 117982.

40 L. Wang, C. Wang, Z. Zhang, J. Wu, R. Ding and B. Lv, Appl.Surf. Sci., 2017, 422, 574–581.

41 Y.-R. Liu, X. Li, W.-M. Liao, A.-P. Jia, Y.-J. Wang, M.-F. Luoand J.-Q. Lu, ACS Catal., 2019, 9, 1472–1481.

Catalysis Science & Technology Paper

Publ

ishe

d on

24

Janu

ary

2020

. Dow

nloa

ded

by U

nive

rsite

Par

is D

esca

rtes

on

2/12

/202

0 12

:15:

14 P

M.

View Article Online