Sie sind auf Seite 1von 43

Accepted Manuscript

Coal tar- and residual oil-derived porous carbon as metal-free catalyst for nitroarene
reduction to aminoarene

Qinhong Wei, Fangfang Qin, Qingxiang Ma, Wenzhong Shen

PII: S0008-6223(18)30907-2
DOI: 10.1016/j.carbon.2018.09.087
Reference: CARBON 13515

To appear in: Carbon

Received Date: 9 August 2018


Revised Date: 23 September 2018
Accepted Date: 29 September 2018

Please cite this article as: Q. Wei, F. Qin, Q. Ma, W. Shen, Coal tar- and residual oil-derived porous
carbon as metal-free catalyst for nitroarene reduction to aminoarene, Carbon (2018), doi: https://
doi.org/10.1016/j.carbon.2018.09.087.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Graphical abstract

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

Coal tar- and residual oil-derived porous carbon as metal-free

catalyst for nitroarene reduction to aminoarene

Qinhong Weia, Fangfang Qina, Qingxiang Mab, Wenzhong Shena,*

PT
a

RI
State Key Laboratory of Coal Conversion, Institute of Coal Chemistry,

Chinese Academy of Sciences, Taiyuan 030001, People’s Republic of

SC
China.

U
b
State Key Laboratory of High-efficiency Coal Utilization and Green
AN
Chemical Engineering, Ningxia University, Yinchuan 750021, People’s

Republic of China.
M
D

*Corresponding author:
TE

E-mail addresses: shenwzh@sxicc.ac.cn


C EP
AC
ACCEPTED MANUSCRIPT

1 Abstract

2 Nitroarenes reduction is an important technology in the industrial production of

3 aminoarenes. In this work, we presented an environment-friendly and low-cost green

4 synthesis route for preparation of oxygen and nitrogen co-doped porous carbon

PT
5 (ONPC) via acid oxidation and alkali activation methods using coal tar- and residual

RI
6 oil-based as starting materials, and the prepared ONPC was employed as metal-free

7 carbon catalyst for nitroarenes reduction to aminoarenes in the presence of hydrazine

SC
8 hydrate. This ONPC catalyst showed much higher catalytic activity as compared to

U
9 those of un-doped porous carbon (PC), activated carbon (AC) and carbon black,
AN
10 which was attributed to its large surface area and developed pore structure as well as

11 O and N co-doping. Additionally, it also exhibited a versatility in various aromatic


M

12 nitro-compounds reduction to relative aromatic amines. Experimental results by using


D

13 model catalysts to simulate different carbons with various oxygen-containing groups


TE

14 proved that carbonyl groups were more favorable for nitrobenzene reduction to

15 aniline. Upon co-doping O and N into PC, the two kinds of introduced species
EP

16 synergistically promoted the catalytic activity of PC. Good performance together with
C

17 low-cost preparation makes oxygen and nitrogen co-doped porous carbon a potential
AC

18 substitution of supported metal catalyst for nitroarenes reduction.

19

20 Keywords: Asphaltene, porous carbon, carbon catalyst, reduction, nitroarenes

21

22

1
ACCEPTED MANUSCRIPT

1 1. Introduction

2 Metal-catalyzed chemical reactions are playing an important role in the

3 production of industrial chemicals and fuels [1-4]. As well known, the representative

4 industrial reactions of catalytic reforming and catalytic hydrocracking of petroleum

PT
5 processing using noble metal Pt and Pd catalysts are being used to produce liquid and

RI
6 gas fuels [5,6]. Ammonia synthesis, an extremely significant chemical reaction

7 catalyzed by non-noble metal iron catalyst, promote the economic development and

SC
8 social progress [7]. In addition, metal catalysts used for heterogeneous reactions like

U
9 nitro compounds hydrogenation [8], alcohol oxidation and carbon-carbon coupling
AN
10 occupy an important position in the production of fine chemicals [9-10]. However,

11 metal-based catalysts tend to suffer from high cost, poor durability, loss of active
M

12 components, and susceptibility to poisoning. To overcome this issue, catalytic


D

13 reactions based on carbon materials have been a research hotspot over the past ten
TE

14 years. For example, carbon nanotubes (CNTs), graphene, and active carbon (AC)

15 employed as metal-free catalysts have exhibited excellent catalytic activity for many
EP

16 heterogeneous catalytic reactions [11-13].


C

17 The virtue of carbon materials-based catalysis is that catalytic stability can be


AC

18 well maintained because of the intrinsic active sites of carbons themselves. Apart

19 from above-mentioned carbons, more carbon materials which are originated from

20 abundant raw materials like coal, petroleum and biomass have been developed and

21 designed to serve as catalyst support, catalytic material as well as electrode material

22 [14,15]. Although various carbon materials, such as CNTs, graphene and ordered

2
ACCEPTED MANUSCRIPT

1 mesoporous carbons with large surface area, porosity and rich surface chemical

2 properties, have been studied, their preparation process are relatively complex and

3 expensive [16,17]. Compared to them, coal tar- and residual oil-derived asphaltene

4 used as raw material to prepare carbon materials presents comparable superiority,

PT
5 where asphaltene can be treated by chemical modification to obtain diverse carbon

RI
6 materials as well as realize heteroatoms doping. In addition to this, asphaltene

7 obtained from the deep processing of coal tar and/or residual oil is widely available

SC
8 and low in cost. Thus, utilizing asphaltene to prepare porous carbon materials is of

U
9 important significance. At present, the researches on asphaltene-based porous carbon
AN
10 materials are scarcely reported. These porous carbon materials are mostly applied in

11 the studies of electrodes for supercapacitors [18-20], while their applications in


M

12 catalysis have not yet been reported. Here, we adopted a common strategy for
D

13 preparation of porous carbon by acid oxidation and alkali activation of asphaltene.


TE

14 The prepared porous carbon was directly used as metal-free catalyst, for the first time,

15 to probe into the hydrogenation of nitroarenes to aminoarenes.


EP

16 In terms of carbon catalytic materials, chemical modification has significant


C

17 impact on their catalytic performances [21]. Oxygen and nitrogen are most widely
AC

18 used to dope and modify carbon materials. It is generally accepted that the introduced

19 oxygen functional groups act as active species to enhance catalytic activity and the

20 doped N can increase the electrical conductivity of carbon body as well as adsorption

21 to reactants [22,23]. Wang et al. prepared carbon cloth supported graphene

22 electrocatalyst by Ar plasma etching method. The defect-rich and

3
ACCEPTED MANUSCRIPT

1 oxygen-functionalized graphene presented excellent electrochemical activity, which

2 was attributed to the synergetic effect of O doping and defect [24]. Su et al. used

3 chemical vapor deposition method with ammonia as nitrogen source to prepare

4 nitrogen-doped carbon nanotubes. They were observed to be effective catalysts for the

PT
5 oxidation of glycerol to dihydroxyacetone [25].

RI
6 Selective reduction of nitroarenes to aminoarenes is considered as an important

7 and fundamental chemical reaction. The produced aminoarenes are key chemical

SC
8 intermediates in the production of pharmaceuticals, agrochemicals, dyes and fine

U
9 chemicals [26,27] Traditional production routes of aminoarenes are mainly carried out
AN
10 by reduction method using non-noble metals Ni [28], Co [29], Fe [30] and sulfides [31]

11 as reduction agents. However, these catalytic processes pose considerable problems


M

12 such as low production efficiency, equipment corrosion and environmental pollution.


D

13 Nowadays, catalytic hydrogenation has served as a good alternative to the


TE

14 conventional recipe for the reduction of nitro compounds because this route can not

15 only enhance the yield of aminoarenes but also greatly reduce environmental
EP

16 pollution. Supported noble-metal-based catalysts have been widely applied for the
C

17 catalytic hydrogenation of nitroarenes to aminoarenes [32-36], and they exhibited


AC

18 high catalytic activities. However, the noble catalysts-catalyzed reactions present

19 many problems such as relatively low selectivity of aminoarenes because of the

20 excessive hydrogenation as well as high production cost. Considering high catalytic

21 activity together with remarkable selectivity, metal-free carbon catalysts have been

22 investigated for chemoselective reduction of nitroarenes to aminoarenes [37,38]. Arai

4
ACCEPTED MANUSCRIPT

1 et al. have reported that a surface-functionalized AC treated by hydrogen peroxide and

2 ammonia exhibited excellent catalytic ability for nitroarenes hydrogenation, where

3 carbon acted as active sites to catalyze this reaction smoothly [39].

4 In present work, we developed a generalized strategy on the preparation of O and

PT
5 N co-doped porous carbons by acid oxidation and alkali activation methods using coal

RI
6 tar- and residual oil-based asphaltene as starting materials. The prepared porous

7 carbon catalytic material with larger surface area and developed porous structure was

SC
8 used as metal-free catalyst for chemoselective reduction of nitroarenes to aromatic

U
9 amines in the presence of hydrazine hydrate. Owing to its large surface area and
AN
10 excellent porosity, the reactant molecules could rapidly transfer to exposed carbon

11 active sites and achieve fast conversion. Moreover, the doped oxygen and nitrogen
M

12 species had the ability of fabricating structure defects and changing electronic
D

13 conductivity of the inherent carbons, consequently resulting in high catalytic activity.


TE

14 2. Experimental section
EP

15 2.1 Catalyst preparation


C

16 2.1.1 Preparation of asphaltene


AC

17 Coal tar- and residual oil-derived residual as feedstock of synthesizing

18 asphaltene was first prepared by co-treatment of coal tar and residual oil in a sealed

19 stainless steel autoclave of 5 liter, where the mixture was stirred in N2 upon high

20 temperature of 420 oC and high pressure of 5 MPa for 4 h.

21 Asphaltene was then prepared by extracting the resulting residual with

5
ACCEPTED MANUSCRIPT

1 tetrahydrofuran (THF) as solvent. In a typical process, 1000 g of residual and 1000 ml

2 of THF were added to a sealed home-made high-temperature filter and stirred for 1 h.

3 Then, the mixture was heated to 100 oC and retained for 2 h, followed by being

4 filtered using a sieve with 800 mesh to collect extracted solubles. By removing THF

PT
5 from solubles using evaporation way, the asphaltene was obtained.

RI
6 2.1.2 Preparation of O and N co-doped porous carbon (ONPC)

7 The as-prepared asphaltene was used as parent carbon to prepare O and N

SC
8 co-doped porous carbon by nitric acid oxidation and potassium hydroxide combined

U
9 with urea activation. A weighted asphaltene of 10 g was dispersed in 150 ml of
AN
10 concentrated nitric acid in a flask, followed by being stirred at 80 oC for 16 h, washed

11 with distilled water until neutrality and dried at 100 oC overnight. The oxidized
M

12 asphaltene was dispersed in a mixed solution of potassium hydroxide and urea (with a
D

13 20:17:5 of mass ratio of KOH:urea), then the mixture was stirred at 80 oC for 10 h,
TE

14 dried at 100 oC for 12 h and activated at 800 oC in Ar for 1 h. Finally, the calcined

15 sample was washed and dried to acquire O- and N-co-doped porous carbon (designed
EP

16 as ONPC). Meanwhile, ONPC with various doped contents of O and N were prepared
C

17 by nitric acid oxidation for respective 12 h, 20 h, 24 h and 30 h as well as potassium


AC

18 hydroxide along with urea activation for 2 h. The ONPC prepared by nitric acid

19 oxidation of 16 h and alkali activation of 1 h was used in the whole content, unless

20 otherwise specified.

21 It is important to point out that in this process of preparation, the KOH was used

22 as activating agent, in which it reacted with carbon under high temperature. As a

6
ACCEPTED MANUSCRIPT

1 result, the porous carbon with developed pore structure was fabricated.

2 For comparative studies, nonporous carbon (defined as non-C) was prepared by

3 direct carbonization of porous carbon precursor without the procedures of nitric acid

4 oxidation and alkali activation; porous carbon (designated as PC) and O-doped porous

PT
5 carbon (named as OPC) were obtained as follow: PC was prepared by KOH activation

RI
6 under Ar atmosphere for 1 h; OPC was prepared by oxidizing PC with H2O2 as

7 oxidant for 24 h. The O and N doped carbon nanosheet was also synthesized using

SC
8 Schiff-base method in a mixed salt. Typically, after asphaltene being oxidized by

U
9 nitric acid for 16 h, the oxidized asphaltene was mixed with melamine, lithium
AN
10 chloride and potassium chloride (mass ratio of 1:1:9), in which the Schiff-base

11 reaction occurred between the aldehyde groups of oxidized asphaltene and the animo
M

12 groups of melamine to form a polymer. Subsequently, the polymer was gradually


D

13 pyrolyzed in the mixed salt of LiCl and KCl at 600 oC in Ar for 2h. Finally, the
TE

14 resulting carbons were washed with deionized water to obtain the O and N doped

15 nanosheets.
EP

16 2.2 Catalyst characterization


C

17 The textural property of porous carbon materials was measured by N2 physical


AC

o
18 adsorption-desorption at -196 C on a JW-BK-122W apparatus. Prior to

19 measurements, all samples were pre-treated under vacuum condition at 200 oC for 5 h.

20 The specific surface area of porous carbons was based on the Brunauer-Emmet-Teller

21 method. The crystalline structure of samples was characterized by X-ray diffraction

22 (XRD) on a D8 Advance diffractometer with a Cu-Kα radiation at an operating

7
ACCEPTED MANUSCRIPT

1 condition of 40 kV and 40 mA. The surface morphologies and element compositions

2 of samples were observed by scanning electron microscopy (SEM) on JSM-7001F

3 instrument of JEOL attached with an energy-diffusive spectroscopy (QX200, EDS)

4 and scanning transmission electron microscopy (TEM) on Tecnai G2 F20 S-Twin of

PT
5 FEI operated at 200 kV accelerating voltage. The thickness of nanosheet porous

RI
6 carbon was measured by using an atomic force microscope (AFM) on Dimension

7 Edge of Bruker. The surface properties of samples were revealed by X-ray

SC
8 photoelectron spectroscopy (XPS) on AXIS ULTRA DLD apparatus with a

U
9 monochromatic radiation source of Al Kα, and XPS peaks were deconvolved by peak
AN
10 fitting. The structure defects of samples were examined by Raman spectrum on

11 Renishaw InVia Reflex spectrometer with a laser of 514 nm. The O- and N-doped
M

12 functional groups on samples were discerned by Fourier transform infrared spectra


D

13 (FT-IR) on TENSOR-27. Before tests, samples were diluted with KBr, pressed into
TE

14 wafers and dried.

15 2.3 Catalytic tests


EP

16 The catalytic activity of O- and N-doped porous carbon catalysts was tested for
C

17 catalytic hydrogenation of nitroarenes to anilines in a 10 ml flask with a reflux


AC

18 condensation. In a typical procedure, 10 mg catalyst, 4.5 mmol substrate, 2 ml

19 hydrazine hydrate (30 %) and 1 ml isopropanol (solvent) were introduced into the

20 flask, and then the reaction mixture was heated to 80 oC for catalytic hydrogenation

21 reaction of 4 h. Other reaction conditions performed (such as reaction temperature and

22 various substrates as well as their dosage) will be illustrated in content. After reactions,

8
ACCEPTED MANUSCRIPT

1 50 µL methyl-phenoxide as internal standard and 2 ml ethanol as diluent were added

2 to the resulting mixture. The liquid sample was collected using nylon filtration

3 membrane and then analyzed by gas chromatograph (GC-250) with a flame ionization

4 detector and a HP-5MS capillary column packed with (5%-phenyl)-95% methyl

PT
5 polysiloxane. The conversions of nitroarenes and selectivities of anilines were

RI
6 calculated based on the internal standard.

SC
7 3. Results and discussion

U
8 3.1 Catalyst characterizations
AN
9 Oxygen and nitrogen co-doped porous carbon (ONPC) was fabricated by acid

10 oxidation and alkali activation method, in which porous carbon precursor was first
M

11 oxidized by nitric acid to introduce oxygen groups and then activated by a mixture of
D

12 KOH and urea to dope nitrogen species. The surface morphologies of asphaltene and
TE

13 as-prepared ONPC samples are shown in Fig. 1. As clearly observed in Fig. 1a,

14 asphaltene possesses a blocky structure with rugged surface, whose surface area is
EP

15 negligible. The ONPC obtained exhibits a morphology of small fragment compared to


C

16 that of asphaltene shown in Fig. 1b. Nitrogen adsorption-desorption isotherm of


AC

17 ONPC displays two types of curves that are composed of classic I type in the

18 low-pressure region and IV type along with an apparent H1 hysteresis loop in Fig. 1c,

19 meaning the coexistence of mesopore and micropore [40]. Pore distribution curve

20 further corroborates the presence of mesopore and micropore in Fig. 1d, which of

21 these pores are mainly concentrated on 2.5 nm and 0.8 nm. The specific surface area

9
ACCEPTED MANUSCRIPT

1 of ONPC is up to 1649 m2/g accompanied by large pore volume of 1.0 cm3/g. BET

2 result indicates that using KOH activator can smoothly fabricate developed porous

3 structure by which large surface area and abundant pores were formed [41]. TEM

4 image of ONPC is displayed in Fig. 1e. It can be distinctly observed that the whole

PT
5 section was populated with micropores, which was consistent with the pore

RI
6 distribution of micropore. EDS analysis verified that O and N were successfully

7 doped into porous carbon by nitric acid oxidation and KOH combined with urea

SC
8 activation shown in Fig. 1f and Fig. S1. For the ONPC owning large surface area and

U
9 excellent pore properties, it is expected that more catalytically active sites are exposed
AN
10 to reactants and developed pore structure facilitate to mass transfer, both of which will

11 accelerate reaction rate.


M

12
D
TE
C EP
AC

10
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE

1
EP

2 Fig. 1. SEM images of asphaltene (a) and ONPC (b); N2 adsorption-desorption

3 isotherms (c), pore size distribution curves (d), TEM image (e) and EDS (f) of ONPC.
C
AC

4 For carbon-based metal-free catalysts, structure defects play critical role in

5 catalytic reactions because they can trigger coordination unsaturation of adjacent

6 carbon atoms [42]. Activated carbon (AC) and carbon black, due to the inherent

7 structure defects, are commonly used as metal-free catalysts for catalytic reactions.

8 Here, in order to characterize defect level, Raman spectra of various samples were

9 investigated in Fig. 2. Two characteristic Raman peaks are observed for all samples,
11
ACCEPTED MANUSCRIPT

1 which are assigned to D band and G band at around 1347 cm-1 and 1590 cm-1

2 respectively. The D band is caused by disorder degree of the poorly-organized crystal

3 structure on account of the breathing vibration of sp2 carbon atoms in amorphous

4 carbon and/or at edge planes of the graphitic domains [43], while G band results from

PT
5 in-plane vibration of sp2 carbon atoms in rings and chains [44]. Therefore, the

RI
6 strength ratio of D peak and G peak (designated as ID/IG) is used to decide the

7 structure defect degree of carbon materials. Among these carbon materials of

SC
8 asphaltene, non-C, PC, OPC and ONPC, asphaltene as initial carbon possesses large

U
9 value of ID/IG = 2.25, larger than those of non-C (1.86), PC (2.03), OPC (2.06) and
AN
10 ONPC (2.15). When asphaltene was activated to prepare PC, the value of ID/IG

11 decreased to 1.86. This could be due to a fact that high-temperature activation and
M

12 carbonization promoted higher graphitization degree. When oxygen and nitrogen were
D

13 both introduced to PC, the prepared ONPC obtained higher ID/IG up to 2.15.
TE

14 Combined with Raman analysis, this result indicates that heteroatoms doped into

15 carbon domains are capable of initiating structure defects remarkably.


EP

16 XRD patterns of non-C, OPC and ONPC samples are shown in Fig. S2. It is
C

17 clear that a broad diffraction peak at about 25o is observed for non-C, which
AC

18 corresponds to (002) crystal surface of graphitic carbon. The presence of broad (002)

19 diffraction peak means the formation of a certain extent of graphitic structure in

20 carbon body, which arose from high-temperature carbonization. However, compared

21 to non-C, OPC and ONPC exhibit dispersed peaks with low intensity shifting to small

22 diffraction angle, indicating that N and/or O dopings may result in framework

12
ACCEPTED MANUSCRIPT

1 expansion and destroy graphitic structure to some extent, which is in agreement with

2 Raman analysis. However, with regard to the asphaltene carbonization to carbon with

3 a certain degree of graphitization under high temperature, the carbon structure was

4 actually transformed from amorphous form to crystalline. The graphitization

PT
5 mechanism can be elaborated from the perspective of the molecule structure

RI
6 transformation of asphaltene. The basic structural unit of asphaltene is the condensed

7 aromatic molecule that consisting of multiple aromatic rings along with side-chain

SC
8 alkyls. With the increase of carbonization temperature, the condensed aromatics occur

U
9 condensation reaction and then form the mesophase, a crossing-linking large molecule.
AN
10 Finally, the mesophase is transformed to carbon with a certain degree of

11 graphitization via high-temperature carbonization [45].


M

12
D

ID IG
TE

e ID/IG = 2.15
Intensity (a.u.)

d ID/IG = 2.03
EP

c ID/IG = 2.06

b ID/IG = 1.86
C

a ID/IG = 2.25
AC

500 1000 1500 2000


-1
13 Raman Shift (cm )

14 Fig. 2. Raman spectra of asphaltene (a), non-C (b), PC (c), OPC (d) and ONPC (e).

15 In general, catalytic activities of metal-free carbon catalysts are closely

16 associated with its textural properties, surface chemical properties as well as the

17 number of active sites. Specially, doped heteroatoms have the ability of making defect
13
ACCEPTED MANUSCRIPT

1 sites and changing electronic state of carbon atoms, thereby enhancing catalytic

2 activity of carbon catalyst. It has been confirmed by previous reports that nitric

3 acid-oxidized CNT exhibited higher catalytic performance for nitrobenzene reduction

4 to aniline than that of pristine CNT, which was attributed to the introduced

PT
5 oxygen-containing functional groups such as carbonyl, carboxylic and hydroxyl

RI
6 groups [46]. In the case of N-doped carbon catalyst, a suitable amount of

7 nitrogen doped into carbon was more beneficial to adsorption of reactants as

SC
8 well as improved electrical conductivity of carbon itself, both of which

U
9 contributed to increased catalytic activity [23].
AN
10 To clarify the O and N species in as-prepared samples, XPS measurements of PC,

11 OPC and ONPC were conducted, as shown in Fig. 3. XPS survey spectra for all
M

12 samples exhibit C 1s and O 1s characteristic peaks located respectively at about 286


D

13 eV and 532.5 eV in Fig. 3a. In addition, N 1s XPS peak was examined in ONPC, and
TE

14 the results affirmed that O and N were co-doped into carbon. In order to detect the

15 existent forms of O and N species, their XPS peaks were deconvolved by fitting peaks.
EP

16 O 1s was deconvolved into three individual peaks for all samples, which are assigned
C

17 to carbonyl group (C-OH, at around 531.5 eV), carboxylic group (O=C-O, at around
AC

18 532.9 eV) and hydroxyl group (C=O, at around 533.9 eV), respectively in Fig. 3b [24].

19 The relative concentrations of surface oxygen and different oxygen-containing groups

20 are summarized in Table 1. Compared to PC owning 18.8 atom% surface oxygen

21 content, OPC presents higher oxygen content of 22.8 atom%, demonstrating that

22 using nitric acid to oxidize PC effectively introduced oxygen into carbon domains.

14
ACCEPTED MANUSCRIPT

1 However, for ONPC, surface oxygen content is only 6.1 atom%. The main reason for

2 the reduced oxygen content is due to the high-temperature activation under Ar, during

3 which part of oxygen groups are susceptible to decomposition upon high temperature

4 process. In addition to this, the N doping is also a cause of leading to the decrease of

PT
5 surface O content, in which the doped N species, due to the interaction with carbon

RI
6 atoms, unavoidably inhibits the binding of O species with carbon atoms, consequently

7 giving rise to the decrease of O content. N 1s XPS spectrum of ONPC is displayed in

SC
8 Fig. 3c. Consistent with the previous reports, N 1s is deconvolved into three XPS

U
9 peaks positioned at 398. 7 eV, 400.1 eV, and 401.6 eV, which is ascribed to pyridinic
AN
10 N, pyrrolic N and graphitic N, respectively [39,47]. For the three kinds of N species

11 formed on ONPC, pyrrolic N is of higher content as compared to those of pyridinic N


M

12 and graphitic N by calculating peak areas of N 1s deconvolution peaks. Based on


D

13 above Raman, XRD and XPS characterizations, we proposed the molecular structure
TE

14 of ONPC catalyst as illustrated in Fig. 3d.


C EP
AC

15
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
1 Fig. 3. XPS survey spectra (a) and O 1s XPS spectra (b) of PC, OPC and ONPC; N 1s
M

2 spectrum of ONPC (c); Schematic illustration of ONPC graphitic structure (d).


3
D

4 Table 1 Surface O contents and O 1s XPS data of PC, OPC and ONPC samples
TE

5
Total O
Catalyst Carbonyl O Carboxylic O Hydroxyl O
atom %
EP

PC 18.8 34.7 37.5 27.8

OPC 22.8 38.5 35.1 26.4


C

ONPC 6.1 34.6 36.7 28.7


AC

6
7 3.2 Catalytic performance evaluation

8 All the prepared carbon materials were tested as metal-free catalysts for

9 nitrobenzene hydrogenation to aniline in the presence of hydrazine hydrate

10 (N2H4·H2O) conducted at 80 oC, and their catalytic activities are shown in Table 2.

11 Neither the blank test nor the absence of hydrazine hydrate were active (Table 2,
16
ACCEPTED MANUSCRIPT

1 entries 1-2), indicating that nitrobenzene reduction was conducted by means of

2 catalyst and reducing agent. When AC was employed as catalyst for nitrobenzene

3 reduction, nitrobenzene conversion was obtained at 71.4 %, along with aniline

4 selectivity in 94.2 %. Similarly, using carbon black as catalyst, nitrobenzene

PT
5 conversion and aniline selectivity reached 76.7 % and 95.1 %, respectively (Table 2,

RI
6 entries 3-4).

7 Larsen et al. reported the reduction of nitrobenzene to aniline using N2H4·H2O as

SC
8 reductant catalyzed by carbon. In the process of reaction, carbon played two roles of

U
9 adsorbent and electrical conductor, ensuring the reaction to perform smoothly [23,48].
AN
10 The nitrobenzene reduction is seen as a six-electron reduction process (as listed in

11 following Eqs. (1-3)), in which the formed intermediates, nitrosobenzene and


M

12 phenylhydroxylamine, gave rise to other by-products such as azobenzene and


D

13 azoxybenzene [48,49]. Despite the ambiguous mechanism by carbon, its electronic


TE

14 property as well as structure defect indeed had a significant effect on this catalytic

15 reaction. In present experiments, AC- and carbon black-catalyzed nitrobenzene


EP

16 reduction produced by-products nitrosobenzene and phenylhydroxylamine


C

17 concomitantly, which may be primarily relative to the inherent surface properties and
AC

18 structure defects as characterized by XPS and Raman spectra in Fig. S3 and Fig. S4.
19
20 (1)
21
22 (2)
23
(3)
24

17
ACCEPTED MANUSCRIPT

1 When ONPC was used as carbon catalyst for nitrobenzene reduction carried out

2 at 80 oC (Table 2, entry 5), nitrobenzene was converted at conversion of 100 %, and

3 aniline selectivity of 100 % was also yielded. By investigating reaction temperatures

4 in a temperature range of 40-70 oC (Table 2, entries 6-9), it can be seen clearly that

PT
5 the catalytic activity of ONPC drops gradually, while aniline selectivity is still

RI
6 maintained in 100 %. Although nitrobenzene reduction to aniline is a

7 thermodynamically exothermic reaction in which low temperature is conductive to

SC
8 catalytic conversion, nitrobenzene conversion increased with the increase of reaction

U
9 temperature. As well known, high reaction temperature can help to speed up reaction
AN
10 rate by which to enhance conversion of nitrobenzene. Again, all of these reacted

11 nitrobenzene upon different temperatures were completely converted to key product


M

12 aniline, without any detectable byproducts. However, supported metal catalysts as


D

13 reference obtained from previous studies exhibit relatively low catalytic activity as
TE

14 well as poor selectivity for nitrobenzene hydrogenation to aniline conducted in H2

15 atmosphere with a certain of pressure (Table 2, entries 10-13) [50-52]. By comparison,


EP

16 the prepared O- and N-co-doped porous carbon, without the presence of metal
C

17 components, presents excellent catalytic activity in conversion of nitrobenzene to


AC

18 aniline upon mild reaction conditions. This may be interpreted that catalytic

19 hydrogenation over metal based catalysts is more vulnerable to reaction temperature

20 and H2 pressure. In addition, metal-catalyzed hydrogenation reactions more easily

21 produce byproducts as compared to that over carbocatalysts, which is due to the

22 excessive hydrogenation resulting from the high hydrogenation performance of metal

18
ACCEPTED MANUSCRIPT

1 catalysts.

PT
Nitrobenzene Aniline
Entry Catalyst t (h)
Conv. (%) Sel. (%)

1 Blank 4 1.52 100

RI
2b AC 4 - -

SC
3 AC 4 71.9 94.2

U
4 Carbon black AN 4 76.7 95.1

5 ONPC (80 oC) 4 100 100

6 ONPC (70 oC) 4 95.7 100


M

7 ONPC (60 oC) 4 88.5 100


D

8 ONPC (50 oC) 4 79.6 100


TE

9 ONPC (40 oC) 4 59.7 100


EP

10 Fe(BF4)24·6H2O/L2c 2 49 >99

11 Ni/CeO2-CASd 5 30.1 100


C

12 PtCo nanoparticlee 1 98.2 90.9


AC

13 Commercial Pt/Cf 30 100 32.4

3 Table 2 Catalytic hydrogenation of nitrobenzene to aniline over different catalystsa


a
4 Reaction conditions: 10 mg catalyst, 4.5 mmol substrate, 2 ml N2H4·H2O, 1 ml
5 isopropanol (solvent), 4 h, Entry (1-5) at 80 oC, Entry (6-9) at 40-70 oC. b Without the
c
6 adding of hydrazine hydrate to reaction system. 0.5 mmol nitrobenzene, 1 mol%
7 catalyst (based on metal), 120 oC, 30 bar H2, 1.5 ml THF (solvent). d 30 mg catalyst,
19
ACCEPTED MANUSCRIPT

1 2.4 mmol nitrobenzene, 150 oC, 1.5 Mpa, 40 ml ethanol (solvent). e,f
0.5 mmol
2 nitrobenzene, 0.3 mol% catalyst (based on metal), room temperature, 1 bar H2, 3 ml
c,d e,f
3 THF (solvent). data derived from the references [50,51] and data derived from
4 the references [52].

5 As for carbon catalytic materials, textural structure and surface chemical nature

PT
6 pose significant effect on catalytic activity. Carbon catalytic materials with small

RI
7 specific surface area and poor pore property usually exhibit low catalytic activity. The

8 nonporous carbon (non-C) without nitric acid oxidation and alkali activation

SC
9 presented low nitrobenzene conversion of 42.1 % (Table 3, entry 1), which was

U
10 attributable to its poor textural structure in Fig. S5. Compared to non-C, porous
AN
11 carbon (PC) converted nitrobenzene in a comparable conversion of 62.7 % (Table 3,

12 entry 2). This is because that using KOH to activate PC could fabricate a certain
M

13 porous structure from BET analysis (189.7 m2/g) in Fig. S6, whereby the obtained
D

14 surface area provided active sites to bring out high nitrobenzene conversion. Upon
TE

15 doping O into PC, the catalytic activity of the resulting OPC increased to 83.2 %

16 (Table 3, entry 3), higher than those of non-C and PC. Su et al. recently reported that
EP

17 O-modified CNTs was used as catalyst for nitrobenzene reduction to aniline. The
C

18 catalytic activity of O-doped CNTs notably increased. In particular, with a high


AC

19 concentration of oxygenated functional groups on CNTs, the nitrobenzene conversion

20 was as high as 95.5 %, far higher than that of pristine CHTs with 40.2 %. We also

21 tested the catalytic activity of ONPC nanosheet for nitrobenzene reduction.

22 Unexpectedly, the ONPC nanosheet also exhibited amazing catalytic activity in good

23 to conversion of 100 %. Its physicochemical features, such as thickness of thin

20
ACCEPTED MANUSCRIPT

1 lamella at about 11.3 nm, large specific surface area (267.8 m2/g) and heteroatoms (O

2 and N) modification may account for the excellent catalytic performance as displayed

3 in Fig. S7, S8 and S9.

PT
4

RI
5

Table 3 Nitrobenzene reduction to aniline over various carbon catalysts

SC
6

Second
Nitrobenzene Aniline
KOHa

U
Entry Catalyst First (HNO3) Third (H2O2)
Conv. (%) Sel. (%)
(melamine + KCl + LiCl)b
AN
1 non-C no no no 42.1 100
M

2 PC no activation (1 h)a no 62.7 100


D

3 OPC no activation (1 h)a 24 83.2 100


TE

4 ONPC oxidation (16 h) activation (2 h)b no 100 100


(nanosheet)
EP

7 Reaction conditions: 10 mg catalyst, 4.5 mmol substrate, 2 ml N2H4 H2O, 1 ml


8 isopropanol (solvent), 4 h, 80 oC.
C

9 To further demonstrate the fact that oxygen groups play key role in enhancing
AC

10 catalytic conversion, we investigated relative model molecules, including anthracene,

11 phenanthrene, anthraquinone, hydroquinone and phthalid used as catalysts for

12 nitrobenzene reduction to aniline. As displayed in Table 4, anthracene and

13 phenanthrene model catalysts which have no any oxygenated groups exhibited low

14 nitrobenzene conversions in separate 58.6 % and 34.2 %. When anthraquinone,


21
ACCEPTED MANUSCRIPT

1 hydroquinone and phthalid respectively possessing carbonyl (C=O), phenol hydroxyl

2 (C-OH) and carboxyl (O=C-O) groups were used as catalysts, nitrobenzene

3 conversions were remarkably increased. The anthraquinone catalyst showed 100 %

4 conversion of nitrobenzene, along with high aniline selectivity in 98.3 %. In view of

PT
5 the comparative results, one can say that oxygen functional groups really have

RI
6 significant effect on promoting catalytic activity. Simultaneously, carbonyl group is

7 particularly deemed as preferable oxygen species to help to realize high conversion of

SC
8 nitrobenzene to aniline. Moreover, by comparing experimental results between

U
9 anthracene and phenanthrene model catalysts, it is found that two types of edge
AN
10 defects of armchair and zigzag showed discernible catalytic activities in nitrobenzene

11 reduction, where armchair defect might have more positive function on this reaction
M

12 [42]. Accordingly, a certain amount of armchair and zigzag edge defects formed on
D

13 graphite layers of the as-prepared catalysts could play a positive role in reduction of
TE

14 nitrobenzene.

15 In addition to oxygen species, nitrogen doped into carbon catalytic materials can
EP

16 significantly improve catalytic activity. N-doped carbons usually comprise diverse


C

17 nitrogen species namely pyrrolic-, pyridinic-, graphitic-, and oxidized nitrogen. These
AC

18 formed nitrogen species contribute differently to the integral material properties [39].

19 At present, N-doped carbon catalytic materials have been widely used in the synthesis

20 of fine chemicals, including aerobic oxidation of alcohols to aldehydes, benzene

21 hydroxylation to phenol, and acetylene hydrochlorination to chloroethylene, etc

22 [23,53,54]. Lin et al. gave a detailed report with regard to high-performance

22
ACCEPTED MANUSCRIPT

1 nitrogen-doped carbon catalytic material used for acetylene hydrochlorination [23].

2 Studies by investigating reaction mechanism and active sites disclosed that pyrrolic-N

3 groups affected the electrical conductivity of carbon domains as well as its adsorption

4 to reactants. By keeping a balance of the relationship between this two properties,

PT
5 high catalytic performance could be accomplished. As expected, our prepared porous

RI
6 carbon with N doping catalyst indeed presented higher catalytic activity compared to

7 any of PC and OPC. For such a result, it is convinced that the introduced N species

SC
8 has significant effect in enhancing catalytic activity. N 1s XPS spectrum by

U
9 deconvolution reveals that pyrrolic N has the highest content among the three species.
AN
10 Does the pyrrolic N group significantly promote the catalytic performance?

Table 4 Investigation of model catalysts for nitrobenzene hydrogenation to aniline


M

11
D
TE
C EP
AC

23
ACCEPTED MANUSCRIPT
Nitrobenzene Aniline
Entry Model Catalysts
Conv. (%) Sel. (%)

1 Anthracene 58.6 100

PT
2 Phenanthrene 34.2 86.4

RI
3 Anthraquinone 100 98.3

SC
O

U
4 hydroquinone OH OH 72.4 99
AN
O
6 Phthalide 59.5 97.1
O
M

1 Reaction conditions: 0.1 g model catalyst, 1 mmol substrate, 2 ml N2H4·H2O, 1 ml


2 isopropanol (solvent), 4 h, 80 oC.
D
TE

3 To illustrate reasons for the excellent catalytic activity of ONPC catalyst, we

4 investigated ONPC catalysts with diverse O and N contents for nitrobenzene


EP

5 reduction. O 1s and N1s XPS peaks were deconvolved in Fig. 4, and their contents
C

6 were calculated by fitting peak areas, as listed in Table 5. From the test results, the
AC

7 catalytic activity did not increase strictly with the increase of oxygen or nitrogen

8 functional group content despite high concentration of pyrrolic N and carbonyl O in

9 each ONPC, which might be due to the synergistic effect between the coexisting

10 oxygen and nitrogen species. Nevertheless, it is noticeable that the catalytic activity of

11 ONPC is susceptible to the total amount of oxygen and nitrogen species, and a

24
ACCEPTED MANUSCRIPT

1 suitable amount of O and N doped into ONPC is beneficial to the enhanced catalytic

2 performance. In metal-free carbon catalyst, carbon serves as active sites and plays

3 critical role in catalytic reaction. In spite of O and N dopants which could promote

4 catalytic reaction, excessive doping resulted in a decrease of active sites because some

PT
5 catalytically active carbons inevitably bonded with introduced O and N groups.

RI
6 Moreover, the very low doing of O and N on ONPC also led to the poor catalytic

7 activity. To affirm it, the ONPC catalyst carbonized at 1000 oC was also tested for

SC
8 nitrobenzene reduction, and the conversion is only 78.3 % (in Table S1) lower than

U
9 those tested on ONPC catalysts with high N and O content (in Table 5). The poor
AN
10 performance can be associated to the low N and O content, with respective 0.76 atom%

11 and 3.6 atom% calculated by XPS (in Table S1 and Fig. S10). With regard to the
M

12 ONPC catalyst carbonized at 1000 oC, the N and O species were easily subjected to
D

13 decomposition by which to lower the doping amount of N and O dopants, thus


TE

14 resulting in low catalytic activity.

15 Until now, although it is difficult to elucidate catalytic mechanism of carbon


EP

16 catalysis, it is an indisputable fact that carbon acting as electrical conductor must be


C

17 for catalytic reaction to occur [48]. For N-doped carbon catalyst, N doping can change
AC

18 overall chemical property of carbon, this effect being propitious to catalytic reaction.

19 Ikeda et al. pointed out that carbon atoms bonded to nitrogen atoms were more in

20 favor of activating reactant molecules [55]. In this work, the experimental results also

21 indicated that the O- and N-co-doped porous carbon did indeed showed the high

22 catalytic activity in comparison to those of non-doped and N or O doped porous

25
ACCEPTED MANUSCRIPT

1 carbon in nitrobenzene reduction to aniline.

3 (a) O=C-O
C-OH C=O
No.1
4

PT
Intensity (a.u.)
No.2
5

RI
6 No.3

SC
No.4

8 536 535 534 533 532 531 530


Binding energy (eV)

U
9 (b) Pyridinic N
Pyrrolic N
Graphitic N
AN
No. 1
10
Intensity (a.u.)

No. 2
11
M

12 \ No. 3
D

13 No. 4
TE

14 404 402 400 398 396


Binding energy (eV)
EP

15

16 Fig. 4. O 1s XPS spectra (a) and N 1s XPS spectra (b) of ONPC catalysts with
C

17 different contents of O and N species. Note: No.1 (acid oxidation 20 h, alkali


AC

18 activation 1 h; No.2 (acid oxidation 12 h, alkali activation 1 h); No.3 (acid oxidation

19 30 h, alkali activation 2 h); No.4 (acid oxidation 24 h, alkali activation 1 h).

20

21 Table 5 Content of surface oxygen and nitrogen species calculated by XPS in ONPC
22 catalysts

Sample Conv. % Total O Total N Oxygen species Nitrogen species


26
ACCEPTED MANUSCRIPT
atom % atom % Carbonyl Carboxylic Hydroxyl Pyridinic Pyrrolic Graphitic
O O O N N N

No.1 92.1 6.7 2.6 30.9 38.7 30.4 20.5 61.4 18.1

No.2 100 4.6 2.7 34.6 36.7 28.7 17.6 71.3 11.1

No.3 90.8 8.5 4.6 19.3 42.4 38.3 32.1 52.7 15.2

PT
No.4 100 7.1 1.9 29.9 40.5 29.6 26.8 39.9 33.3

RI
1 In terms of nitrobenzene reduction to aniline in the presence of N2H4·H2O, the

SC
2 decomposition of N2H4·H2O is one of the vital factors. To investigate the ability of

U
3 ONPC to decompose N2H4·H2O, the decomposition experiments of N2H4·H2O over
AN
4 different catalysts were studied in Fig. 5. Unlike nitrobenzene reduction which is a

5 complicated reaction process by carbon catalysis, single N2H4·H2O decomposition is


M

6 relatively simple as N2H4·H2O decomposes to produce NH3 and N2 gases. The


D

7 decomposition rate was recorded by gauging gas production. It is obvious that ONPC
TE

8 exhibits the strongest ability of decomposing N2H4·H2O among these samples. At the
EP

9 beginning, the gas production rate over ONPC is as high as 3.5 ml/min. With the

10 increase of decomposition time, the production rate of gas dropped. When


C

11 decomposition time increases to 3 h, gas production rate decreased to 1.9 ml/min. The
AC

12 fast N2H4·H2O decomposition over ONPC explained why ONPC presented higher

13 catalytic performance for nitrobenzene reduction. As explained, large surface area of

14 ONPC provided adequate carbon active sites for nitrobenzene reduction; O and N

15 dopings changed overall electrical property of carbon. Both factors contributed to the

16 excellent catalytic activity. AC shows quite low decomposition rate of N2H4·H2O in

27
ACCEPTED MANUSCRIPT

1 around 0.2 ml/min inferior to that of PC, but AC presented higher nitrobenzene

2 conversion of 71.9 % compared to that of 62.7 % over PC. In a combination of

3 N2H4·H2O decomposition and catalytic activity, overquick decomposition rate of

4 N2H4·H2O is inclined to generate NH3 and N2 gases, while a mild decomposition rate

PT
5 ensures efficient catalytic reduction of nitrobenzene, which is similar to a previous

RI
6 report related to the H2O2 decomposition over reduced graphene oxide [50].

3.5

SC
3.0
Volume flow rate/ml⋅min-1

ONPC
PC
2.5

2.0

U
1.5
AN
1.0
AC Asphaltene
0.5

0.0
0 30 60 90 120 150 180
M

7 t/min

8 Fig. 5. Decomposition of N2H4·H2O over different catalysts. Decomposition tests


D

9 were carried out in a sealed glass container of 100 ml with a soaping flow tube.
TE

10 Experimental conditions: catalyst 1g, hydrazine hydrate 50 ml, 80 oC, 3 h.


EP

11 Catalyst stability is an important characteristic in practical application. Here, the

12 catalytic reusability of ONPC catalyst was investigated for nitrobenzene reduction.


C
AC

13 Fig. 6 gives the test results for repeated reduction runs with eight times. It is observed

14 clearly that the ONPC catalyst exhibits considerable catalytic stability accompanied

15 by selectivity of 100 %, and there is no distinct activity loss during nitrobenzene

16 conversion. These results intensely indicate that the ONPC is an extremely stable

17 carbon catalyst in nitrobenzene reduction to aniline. It is worth mentioning that

18 graphene oxide (GC) as hot carbon material has been used in many catalytic reactions.
28
ACCEPTED MANUSCRIPT

1 However, its surface oxygen groups regarded as active species easily lose upon

2 hydrothermal or reaction conditions, giving rise to the unstable catalytic activity [56].

3 For the ONPC carbon catalyst, it was prepared by first nitric acid oxidation and

4 subsequent potassium hydroxide activation. Then, these surface oxygen groups

PT
5 obtained via high-temperature activation had strong interaction with carbon atoms, the

RI
6 surface feature being key to contribute to stable catalytic activity.

7 To prove this fact, the chemical composition and structure of the used ONPC

SC
8 catalyst after recycling tests were characterized by XRD pattern, TEM image, Raman

U
9 spectrum and XPS spectrum, as displayed in Fig. S11. The TEM image shows that a
AN
10 mass of worm-like shapes was observed, indicating that the well-developed porous

11 structure of ONPC catalyst were maintained after the 8 times recycling tests. The
M

12 Raman spectrum reveals that the ID/IG value (2.11) is close to that of 2.15 on fresh
D

13 ONPC, meaning that the structure defect of ONPC catalyst has almost no change
TE

14 before and after reaction. For the surface composition of the used ONPC, the result of

15 XPS analysis exhibits that the surface content of N and O species presents tiny
EP

16 difference as compared to that of the fresh ONPC, and the analysis result is listed in
C

17 Table S2. It is found that the O and N content as well as the composition of various N
AC

18 and O groups are nearly the same as that of the fresh ONPC. This could be explained

19 that unlike rigorous catalytic hydrogenation under high temperature and pressure, the

20 reduction of nitrobenzene to aniline catalyzed over ONPC was carried out in the mild

21 reaction conditions of atmosphere pressure and low temperature (80 oC), therefore,

22 this reaction did not pose a serious effect on the structure and surface chemical

29
ACCEPTED MANUSCRIPT

1 composition of the ONPC, exhibiting stable catalytic activity.

2
Conversion Selectivity
100
3

Conversion/Selectivity (%)
80
4
60

PT
5
40

RI
6 20

7 0

SC
1 2 3 4 5 6 7 8
Run nunbers
8

U
9 Fig. 6. The recycling tests of ONPC for nitrobenzene reduction to aniline. Reaction
AN
10 conditions: 10 mg catalyst, 4.5 mmol substrate, 2 ml N2H4·H2O, 1 ml isopropanol

11 (solvent), 80 oC, 4 h for each run.


M

12 3.3 Reduction of various aromatic nitro-compounds


D

13 To explore the potential and versatility of this ONPC catalyst, various aromatic
TE

14 nitro-compounds reduction to relative anilines over ONPC were also carried out at 80
EP

o
15 C. As shown in Table 6, with different reactants catalyzed by ONPC, their

16 conversions are strikingly distinct. As for 1-chloro-3-nitrobenzene, p-nitrophenol and


C

17 1-bromo-4-nitrobenzene reactants (Table 6, entries 1-3), they were completely


AC

18 converted in 100% conversion, and relative aromatic amines were selectively yielded

19 with high selectivity in respective 96.3 %, 99.2 % and 90%. Nevertheless, in the case

20 of p-iodonitrobenzene reduction (Table 6, entry 4), conversion is only 69.2 %, besides,

21 the selectivity of key product p-iodoaniline is 76.4 %. Interestingly, in the process of

22 p-iodonitrobenzene reduction, part of iodine groups was also hydrogenated with the

30
ACCEPTED MANUSCRIPT

1 exception of the hydrogenation of nitro group. As a result, nitrobenzene was produced

2 due to the reduction of iodine group, leading to the decreased selectivity of

3 p-iodoaniline. Different from those reactions mentioned above, conversions of

4 p-nitrotoluene, p-nitroaniline and 2-nitroaniline are lower (Table 6, entries 5-7),

PT
5 especially the latter two in only 15.8 % and 5.5 %. How can these significant

RI
6 differences be understood? The reason for this is that the doping of O and N into

7 carbon was capable of fabricating some polar sites, as oxygen and nitrogen both have

SC
8 stronger polarity compared to that of carbon. These formed polar sites tend to be more

U
9 likely to adsorb nitro groups because of its polar feature and, thereby, the N/O doped
AN
10 porous carbon acting as carbocatalyst exhibits excellent catalytic performance for the

11 reduction of nitrobenzene to aniline [57]. In contrast, for the reduction of


M

12 p-nitrotoluene, the N/O doped porous carbon presented low catalytic activity because
D

13 the presence of methyl group that possesses non-polar property suppressed its
TE

14 adsorption on the polar surface of N/O doped porous carbon. Similarly, amino groups,

15 due to basic property, made p-nitroaniline and 2-nitroaniline adsorbed difficulty on


EP

16 the surface of ONPC, thus inhibiting effective catalytic reaction.


C
AC

17

18

19 Table 6 Reduction of diverse aromatic nitro-compounds to relative anilines over


20 ONPC catalyst

Entry Substrate Conversion (%) Product Selectivity (%)

31
ACCEPTED MANUSCRIPT
Cl NO2 Cl NH2

1 100 96.3

NO2 NH2

2 100 99.2
OH OH

NO2 NH2

PT
3 100 90.2
Br Br

RI
NO2 NH2

4 69.2 76.4
I I

SC
NO2 NH2

5 49.6 100

U
CH3 CH3

NO2 NH2
AN
6 15.8 100
NH2 NH2

NO2 NH2
M

7 5.5 100
NH2 NH2
D

1 Reaction conditions: 10 mg catalyst, 1.3 mmol substrate, 2 ml N2H4 H2O, 1 ml


TE

2 isopropanol (solvent), 4 h, 80 oC.


EP

3 Based on physicochemical characterizations and test results, the proposed

4 reaction mechanism for reduction of nitroarenes to aminoarenes over ONPC using


C

5 hydrazine hydrate as reductant is illustrated in Fig. 7. The nitroarenes and hydrazine


AC

6 hydrate molecules adsorbed occurred at reduction reaction on ONPC, in which the

7 doped O and N served as active species to promote the catalytic reduction of

8 nitroarenes. Decomposition experiments of hydrazine hydrate indicated that ONPC

9 showed strong ability of decomposing hydrazine hydrate. However, unlike the single

10 decomposition of hydrazine hydrate over ONPC, the reaction system produced more

32
ACCEPTED MANUSCRIPT

1 H+ which partook in the reduction of nitroarene. As raised in Eqs. (1-3), nitroarene

2 reduction is deemed as a six electron reduction process, during which nitroarene is

3 reduced step by step to aminoarene.

PT
5

RI
6

SC
7

U
AN
9
M

10
D

11 Fig. 7. Proposed reaction mechanism for reduction of nitroarene to aminoarene over


TE

12 ONPC catalyst.
EP

13 4. Conclusions

14 An oxygen and nitrogen co-doped porous carbon (ONPC) was prepared using
C

15 coal tar- and residual oil-derived asphaltene as raw material by oixdation of nitric acid
AC

16 and activation of potassium hydroxide combined with urea. The prepared ONPC with

17 large specific surface area and developd pore structure used as metal-free carbon

18 catalyst provided abundant carbon active sites, effecitvely enhancing its catalytic

19 activity and strong stability for nitrobenzene reduction to aniline, better than those of

20 un-doped porous carbon (PC), AC as well as carbon black. The dopings of O and N

33
ACCEPTED MANUSCRIPT

1 significantly improved cataltyic activity of PC. A fundamental consideration using

2 model chemicals to imitate carbon catalysts with different oxygenated groups was

3 performed to reveal the function of oxygen species. The resutls indicated that

4 carbonyl groups is the most preferable oxygen active species to promote the reduction

PT
5 of nitrobenzene. When nitrogen was doped into PC, the formed N groups

RI
6 synergistically cooperated with oxygen species, strikingly promoting catalytic activity

7 of PC in nitrobenzene reduction to aniline as well as other nitroarenes reduction to

SC
8 relative aminoarenes.

9 Acknowledgements
U
AN
10 The work was financially supported by National Key R&D Program of China
M

11 (2016YFE0203500), National Natural Science Foundation of China (Grant No.

12 U1510122), Foundation of State Key Laboratory of High-efficiency Utilization of


D

13 Coal and Green Chemical Engineering (Grant No. 2017-K11), and China Postdoctoral
TE

14 Science Foundation (Grant No. 2018M631777).


EP

15 References
[1] G.G. Liu, Q.J. Chen, E. Oyunkhand, S.Y. Ding, N. Yamane, G.H. Yang, Y.
C

16

17 Yoneyama, N. Tsubaki. Nitrogen-rich mesoporous carbon supported iron catalyst


AC

18 with superior activity for Fischer-Tropsch synthesis, Carbon 130 (2017) 304-314.
19 [2] Q.N. Wang, L. Shi, W. Li, W.C. Li, R. Si, F. Schüth, A.H. Lu, Cu supported on
20 thin carbon layer-coated porous SiO2 for efficient ethanol dehydrogenation, Catal.
21 Sci. Technol. 8 (2018) 472-479.
22 [3] A. Tanimu, S. Jaenicke, K. Alhooshani, Heterogeneous catalysis in contiuous
23 flow microreactors: A review of methods and applications, Chem. Eng. J. 327
24 (2017) 792-821.
34
ACCEPTED MANUSCRIPT

1 [4] J. Sun, X.G. Li, A. Taguchi, T. Abe, W.Q. Niu, P. Lu, Y. Yoneyama, N. Tsubaki,
2 Highly-dispersed metallic ru nanoparticles sputtered on h-beta zeolite for directly
3 converting syngas to middle isoparaffins, ACS Catal. 4 (2013) 1-8.
4 [5] M.R. Rahimpour, M. Jafari, D. Iranshahi, Progress in catalytic naphtha reforming
5 process: a review, Appl Ener 109 (2013) 79-93.

PT
6 [6] P. Bai, M. Xie, U.J. Etim, W. Xing, P. Wu, Y. Zhang, B. Liu, Y. Wang, K. Qiao, Z.
7 Yan, Zeolite Y mother liquor modified γ-al2o3 with enhanced brönsted acidity as

RI
8 active matrix to improve the performance of fluid catalytic cracking catalyst, Ind.
9 Eng. Chem. Res. 57 (2018) 1389-1398.

SC
10 [7] C. Song, Q. Liu, N. Ji, Y. Song, Y. Kitamura, Conceptual design and process
11 feasibility analysis of a novel ammonia synthesis process by efficient heat

U
12 integration, ACS Sustainable Chem. Eng. 5 (2017) 7420-7432.
AN
13 [8] R.V. Jagadeesh, A.E. Surkus, H. Junge, M.M. Pohl, J. Radnik, J. Rabeah, H.M.
14 Huan, V. Schünemann, A. Brückner, M. Beller, Nanoscale Fe2O3-based catalysts
M

15 for selective hydrogenation of nitroarenes to anilines, Science 342 (2013)


16 1073-1076.
D

17 [9] K. Nogi, T. Fujihara, J. Terao, Y. Tsuji, Carboxyzincation employing carbon


TE

18 dioxide and zinc powder: cobalt-catalyzed multicomponent coupling reactions


19 with alkynes, J. Am. Chem. Soc. 138 (2016) 5547-5550.
EP

20 [10] J. Rabeah, U. Bentrup, R. Stosser, A. Bruckner, Mechanistic insights by


21 simultaneously coupled operando EPR/UV-Vis/ATR-IR spectroscopy, Angew.
Chem. Int. Ed. 54 (2015) 11791-11794.
C

22

23 [11] O.S.G.P. Soares, R.P. Rocha, A.G. Gonçalves, J.L. Figueiredo, J.J.M. Órfão,
AC

24 M.F.R. Pereira, Highly active n-doped carbon nanotubes prepared by an easy ball
25 milling method for advanced oxidation processes, Appl. Catal. B: Environ. 192
26 (2016) 296-303.
27 [12] Y. Ito, W. Cong, T. Fujita, Z. Tang, M. Chen, High catalytic activity of nitrogen
28 and sulfur co-doped nanoporous graphene in the hydrogen evolution reaction,
29 Angew. Chem. Int. Ed. 54 (2015) 2131-2136.
30 [13] B. Zhang, Z. Wen, S. Ci, S. Mao, J. Chen, Z. He, Synthesizing nitrogen-doped
35
ACCEPTED MANUSCRIPT

1 activated carbon and probing its active sites for oxygen reduction reaction in
2 microbial fuel cells, ACS Appl. mater. interfaces 6 (2014) 7464-7470.
3 [14] M. Li, F. Xu, H. Li, Y. Wang, Nitrogen-doped porous carbon materials:
4 promising catalysts or catalyst supports for heterogeneous hydrogenation and
5 oxidation, Catal. Sci. Technol. 6 (2016) 3670-3693.

PT
6 [15] S. Gao, B.S. Villacorta, L. Ge, T.E. Rufford, Z. Zhu, Effect of sonication and
7 hydrogen peroxide oxidation of carbon nanotube modifiers on the microstructure

RI
8 of pitch-derived activated carbon foam discs, Carbon 124 (2017) 142-151.
9 [16] R. Lv, T. Cui, M.S. Jun, Q. Zhang, A. Cao, D.S. Su, Z. Zhang, S.H. Yoon, J.

SC
10 Miyawaki, I. Mochida, F. Kang, Open-ended, N-doped carbon
11 nanotube-graphene hybrid nanostructures as high-performance catalyst support,

U
12 Adv. Funct. Mater. 21 (2011) 999-1006.
AN
13 [17] Y.S. Liu, Q. Cai, H.D. Li, J.M. Zhang, Fabrication and characterization of
14 mesoporous carbon nanosheets using halloysite nanotubes and polypyrrole via a
M

15 template-like method, J. Appl. Polym. Sci. 128 (2013) 517-522.


16 [18] T.Y. Kim, G. Jung, S. Yoo, K.S. Suh, R.S. Ruoff, Activated graphene-based
D

17 carbons as supercapacitor electrodes with macro- and mesopores, ACS Nano 7


TE

18 (2013) 6899-6905.
19 [19] K.S. Lee, M. Park, S. Choi, J.D. Kim, Preparation and characterization of N,
EP

20 S-codoped activated carbon-derived asphaltene used as electrode material for an


21 electric double layer capacitor, Colloids Surf. A 529 (2017) 107-112.
[20] Q. Wang, J. Yan, Y. Wang, T. Wei, M. Zhang, X. Jing, Z. Fan, Three-dimensional
C

22

23 flower-like and hierarchical porous carbon materials as high-rate performance


AC

24 electrodes for supercapacitors, Carbon 67 (2014) 119-127.


25 [21] D.H. Deng, K.S. Novoselov, Q. Fu, N.F. Zheng, Z.Q. Tian, X.H. Bao, Catalysis
26 with two-dimensional materials and their heterostructures, Nat. Nanotech. 11
27 (2016) 218-230.
28 [22] M. Inagaki, M. Toyoda, Y. Soneda, T. Morishita, Nitrogen-doped carbon
29 materials. Carbon 132 (2018) 104-140.
30 [23] R. Lin, S.K. Kaiser, R. Hauert, J. Pérez-Ramírez, Descriptors for
36
ACCEPTED MANUSCRIPT

1 high-performance nitrogen-doped carbon catalysts in acetylene


2 hydrochlorination, ACS Catal. 8 (2018) 1114-1121.
3 [24] Z. Liu, Z. Zhao, Y. Wang, S. Dou, D. Yan, D. Liu, Z. Xia, S. Wang, In situ
4 exfoliated, edge-rich, oxygen-functionalized graphene from carbon fibers for
5 oxygen electrocatalysis, Adv. Mater. 29 (2017) 1606207-1606213.

PT
6

7 [25] N. Gupta, O. Khavryuchenko, A. Villa, D. Su, Metal-free oxidation of glycerol

RI
8 over nitrogen-containing carbon nanotubes, ChemSusChem 10 (2017)
9 3030-3034.

SC
10 [26] K. Natte, R.V. Jagadeesh, M. Sharif, H. Neumann, M. Beller, Synthesis of nitriles
11 from amines using nanoscale co3o4-based catalysts via sustainable aerobic

U
12 oxidation, Org. Biomol. Chem. 14 (2016) 3356-3359.
AN
13 [27] M. Johannsen, K.A. Jørgensen, Allylic amination, Chem. Rev. 98 (1998)
14 1689-1708.
M

15 [28] X. Meng, H. Cheng, Y. Akiyama, Y. Hao, W. Qiao, Y. Yu, F. Zhao, S. Fujita, M.


16 Arai, Selective hydrogenation of nitrobenzene to aniline in dense phase carbon
D

17 dioxide over Ni/γ-Al2O3: significance of molecular interactions, J. Catal. 264


TE

18 (2009) 1-10.
19 [29] U. Sharma, P. Kumar, N. Kumar, V. Kumar, B. Singh, Highly chemo- and
EP

20 regioselective reduction of aromatic nitro compounds catalyzed by recyclable


21 copper (II) as well as Cobalt (II) phthalocyanines, Adv. Synth. Catal. 352 (2010)
1834-1840.
C

22

23 [30] M. Tian, X. Cui, M. Yuan, J. Yang, J. Ma, Z. Dong, Efficient chemoselective


AC

24 hydrogenation of halogenated nitrobenzenes over an easily prepared


25 γ-Fe2O3-modified mesoporous carbon catalyst, Green Chem. 19 (2017)
26 1548-1554.
27 [31] C. Zhang, Z. Zhang, X. Wang, M. Li, J. Lu, R. Si, F. Wang, Transfer
28 hydrogenation of nitroarenes to arylamines catalysed by an oxygen-implanted
29 MoS2 catalyst, Appl. Catal. A: Gen. 525 (2016) 85-93.
30 [32] J.B. Xi, H.Y. Sun, D. Wang, Z.Y. Zhang, X.M. Duan, J.W. Xiao, F. Xia, L.M. Liu,
37
ACCEPTED MANUSCRIPT

1 S. Wang, Confined-interface-directed synthesis of palladium single-atom


2 catalysts on graphene/amorphous carbon, Appl. Catal. B: Environ. 225 (2018)
3 291-297.
4 [33] J.B. Xi, J.W. Xiao, F. Xiao, Y.X. Jin, Y. Dong, F. Jing, S. Wang,
5 Mussel-inspired functionalization of cotton for nano-catalyst support and its

PT
6 application in a fixed-bed system with high performance, Sci. Rep. 6 (2016)
7 21904.

RI
8 [34] X.M. Duan, M.C. Xiao, S. Liang, Z.Y. Zhang, Y. Zeng, J.B. Xi, S. Wang,
9 Ultrafine palladium nanoparticles supported on nitrogen-doped carbon

SC
10 microtubes as a high-performance organocatalyst, Carbon 119 (2017) 326-331.
11 [35] S. Furukawa, Y. Yoshida, T. Komatsu, Chemoselective hydrogenation of

U
12 nitrostyrene to aminostyrene over Pd- and Rh-based intermetallic compounds,
AN
13 ACS Catal. 4 (2014) 1441-1450.
14 [36] J. Liu, J.F. Hao, C.C. Hu, B.J. He, J.B. Xi, J.W. Xiao, S. Wang, Z.W. Bai,
M

15 Palladium nanoparticles anchored on amine-functionalized silica nanotubes as a


16 highly effective catalyst, J. Phys. Chem. C 122 (2018) 2696-2703.
D

17 [37] J.B. Xi, Q.J. Wang, J. Liu, L. Huan, Z.L. He, Y. Qiu, J. Zhang, C.Y. Tang, J.
TE

18 Xiao, S. Wang, N,P-dual-doped multilayer graphene as an efficient carbocatalyst


19 for nitroarene reduction: A mechanistic study of metal-free catalysis, J. Catal.
EP

20 359 (2018) 233-241.


21 [38] X.D. Duan, J. Liu, J.F. Hao, L.M. W, B.J. He, Y. Qiu, J. Zhang, Z.L. He, J.B. Xi,
S. Wang, Magnetically recyclable nanocatalyst with synergetic catalytic effect
C

22

23 and its application for 4-nitrophenol reduction and Suzuki coupling reactions,
AC

24 Carbon 130 (2018) 806-813.


25 [39] H. Watanabe, S. Asano, S. Fujita, H. Yoshida, M. Arai, Nitrogen-doped,
26 metal-free activated carbon catalysts for aerobic oxidation of alcohols, ACS
27 Catal. 5 (2015) 2886-2894.
28 [40] H. Sun, P. Peng, Y. Wang, C. Li, F. Subhan, P. Bai, W. Xing, Z. Zhang, Z. Liu, Z.
29 Yan, Preparation, scale-up and application of meso-ZSM-5 zeolite by sequential
30 desilication-dealumination, J. Porous Mater. 24 (2017) 1513-1525.
38
ACCEPTED MANUSCRIPT

1 [41] S.E. Chun, J.F. Whitacre, Formation of micro/mesopores during chemical


2 activation in tailor-made nongraphitic carbons, Microporous Mesoporous Mater.
3 251 (2017) 34-41.
4 [42] Y. Gao, D. Ma, C. Wang, J. Guan, X. Bao, Reduced graphene oxide as a catalyst
5 for hydrogenation of nitrobenzene at room temperatur, Chem. Commun. 47

PT
6 (2011) 2432-2434.
7 [43] A. Kaniyoor, T.T. Baby, T. Arockiadoss, N. Rajalakshmi, S. Ramaprabhu,

RI
8 Wrinkled graphenes: a study on the effects of synthesis parameters on
9 exfoliation-reduction of graphite oxide, J. Phys. Chem. C 115 (2011)

SC
10 17660-17669.
11 [44] M.S. Dresselhaus, A. Jorio, M. Hofmann, G. Dresselhaus, R. Saito, Perspectives

U
12 on carbon nanotubes and graphene raman spectroscopy, Nano lett. 10 (2010)
AN
13 751-758.
14 [45] M. Inagaki, F. Kang, M. Toyoda, H. Konno, Advanced Materials Science and
M

15 Engineering of Carbon, Elsevier, London, 2014, pp. 87-107.


16 [46] S.C. Wu, G.D. Wen, J. Wang, J.F. Rong, B.N. Zong, R. Schlöglc, D.S. Su,
D

17 Nitrobenzene reduction catalyzed by carbon: does the reaction really belong to


TE

18 carbocatalysis, Catal. Sci. Technol. 4 (2014) 4183-4187.


19 [47] Z. Li, L. Yin, Nitrogen-doped mof-derived micropores carbon as immobilizer for
EP

20 small sulfur molecules as a cathode for lithium sulfur batteries with excellent
21 electrochemical performance, ACS Appl. Mater. interfaces 7 (2015) 4029-4038.
[48] J.W. Larsen, M. Freund, K.Y. Kim, M. Sidovar, J.L. Stuart, mechanism of the
C

22

23 carbon catalyzed reduction of nitrobenzene by hydrazine, Carbon 38 (2000)


AC

24 655-661.
25 [49] A. Demortier, A.J. Bard, Electrochemical reactions of organic compounds in
26 liquid ammonia. I. reduction of benzophenone, J. Am. Chem. Soc. 95 (1973)
27 3495-3500.
28 [50] G. Wienhöfer, M. Baseda-Krüger, C. Ziebart, F.A. Westerhaus, W. Baumann, R.
29 Jackstell, K. Junge, M. Beller, Hydrogenation of nitroarenes using defined iron–
30 phosphine catalysts, Chem. Comm. 49 (2013) 9089-9091.
39
ACCEPTED MANUSCRIPT

1 [51] W. She, T.Q.J. Qi, M.X. Cui, P.F. Yan, N.S. Weng, W.Z. Li, G.M. Li, High
2 catalytic performance of CeO2-supported Ni catalyst for hydrogenation of
3 nitroarenes fabricated via coordination-assisted strategy, ACS Appl. mater.
4 interfaces 10 (2018) 14698-14707.
5 [52] H. Miao, S.W. Hu, K.L. Ma, L. Sun, F.F. Wu, H.T. Wang, H.Q. Li, Synthesis of

PT
6 PtCo nanoflowers and its catalytic activity towards nitrobenzene hydrogenation,
7 Catal. Commun. 109 (2018) 33-37.

RI
8 [53] J.H. Yang, G. Sun, Y. Gao, H. Zhao, P. Tang, J. Tan, A.-H. Lu, D. Ma, Direct
9 catalytic oxidation of benzene to phenol over metal-free graphene-based catalyst,

SC
10 Energy Environ. Sci. 6 (2013) 793.
11 [54] S. Zhu, Y. Cen, M. Yang, J. Guo, C. Chen, J. Wang, W. Fan, Probing the intrinsic

U
12 active sites of modified graphene oxide for aerobic benzylic alcohol oxidation,
AN
13 Appl. Catal. B: Environ. 211 (2017) 89-97.
14 [55] T. Ikeda, M. Boero, S.F. Huang, K. Terakura, M. Oshima, J. Ozaki, Carbon alloy
M

15 catalysts: active sites for oxygen reduction reaction, J. Phys. Chem. C 112 (2008)
16 14706-14709.
D

17 [56] X. Zhao, J. Wang, C. Chen, Y. Huang, A. Wang, T. Zhang, Graphene oxide for
TE

18 cellulose hydrolysis: how it works as a highly active catalyst, Chem. Commun.


19 50 (2014) 3439-3442.
EP

20 [57] S. Fujita, H. Watanabe, A. Katagiri, H. Yoshida, M. Arai, Nitrogen and


21 oxygen-doped metal-free carbon catalysts for chemoselective transfer
hydrogenation of nitrobenzene, styrene, and 3-nitrostyrene with hydrazine, J.
C

22

23 Mol. Catal. A: Chem. 393 (2014) 257-262.


AC

40

Das könnte Ihnen auch gefallen