You are on page 1of 8

Chemical Engineering Journal 262 (2015) 356–363

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Fabrication of Pd/c-Al2O3 catalysts for hydrogenation


of 2-ethyl-9,10-anthraquinone assisted by plant-mediated
strategy
Huimei Chen a, Dengpo Huang b, Xiyao Su b, Jiale Huang b,⇑, Xiaolian Jing a, Mingming Du b,
Daohua Sun b, Lishan Jia b, Qingbiao Li a,b,c,⇑
a
Environmental Science Research Center, College of the Environment & Ecology, Xiamen University, Xiamen 361005, PR China
b
Department of Chemical and Biochemical Engineering, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, PR China
c
College of Chemistry & Life Science, Quanzhou Normal University, Quanzhou 362000, PR China

h i g h l i g h t s

 Green Pd/Al2O3 catalysts for 2-ethyl-9,10-anthraquinone hydrogenation were fabricated.


 The maximum H2O2 yield of 96.4% could be achieved at lower reaction temperature.
 The Pd/c-Al2O3 catalysts exhibited their excellent durability.

a r t i c l e i n f o a b s t r a c t

Article history: Pd/c-Al2O3 catalysts efficient for the liquid-phase hydrogenation of 2-ethyl-9,10-anthraquinone were
Received 27 May 2014 prepared via an adsorption–reduction method using Cacumen Platycladi extract as both reductive and
Received in revised form 6 September 2014 protective agent. The catalysts were characterized by N2 physisorption, X-ray diffraction, transmission
Accepted 30 September 2014
electron microscopy and X-ray photoelectron spectroscopy. The results showed that Pd nanoparticles
Available online 8 October 2014
(NPs) measuring <10 nm were uniformly distributed on c-Al2O3. Based on the catalyst with the Pd load-
ing of 0.5 wt.%, H2O2 yield of 85.6% could be achieved at the optimum reaction conditions (H2 flow rate
Keywords:
40 mL/min, working solution volume 10 mL, hydrogenation time 1.5 h and reaction temperature 40 °C).
Bioreduction
Pd/c-Al2O3
And the reaction temperature was lower than those in the previous reports. Additionally, the catalysts
Hydrogenation showed much better catalytic activities after pretreatment of the support by HCl, NaOH, Na2CO3 and Na2-
Anthraquinone SiO3 solutions. Particularly, after c-Al2O3 was pretreated with Na2SiO3 solution, the H2O2 yield could
reach 96.4% and the catalysts exhibited excellent durability while the degradation products were inhib-
ited to some extent. Furthermore, the Pd NPs thereof did not aggregate after reuse. Therefore, the cata-
lysts assisted by plant-mediated strategy may be promising for the industrial application.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction and non-polar solvents [4,5]. As presented in Scheme 1, during


the reaction, eAQ is first catalytically hydrogenated to produce
Hydrogen peroxide (H2O2) is an environmentally friendly the corresponding 2-ethyl-9,10-anthrahydroquinone (eAQH2) and
chemical that is widely used in many chemical processes [1]. Com- 2-ethyl-5,6,7,8-tetrahydro-9,10-anthrahydroquinone (H4eAQH2).
mercially, H2O2 is produced by the anthraquinone auto-oxidation Then, the as-obtained eAQH2 and H4eAQH2 are oxidized by O2/air
process [2,3] in which 2-ethyl-9,10-anthraquinone (eAQ) is dis- to produce H2O2 and some other active anthraquinones (mixture
solved in an organic solvent system consisted of a mixture of polar of the original eAQ and 2-ethyl-5,6,7,8-tetrahydro-9,10-anthraqui-
none (H4eAQ)) [1,2,6–8]. However, in cyclic hydrogenation/oxida-
tion operations, due to the consumption via the parallel reaction
pathway I and II (Scheme 1), the content of eAQH2 slowly
⇑ Corresponding authors at: Department of Chemical and Biochemical Engineering,
College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005,
decreases and produces products termed as ‘‘degradation prod-
PR China (Q. Li and J. Huang). Tel.: +86 592 2183088; fax: +86 592 2184822. ucts’’ [9,10]. Among these, the reaction pathway I is a consecutive
E-mail addresses: cola@xmu.edu.cn (J. Huang), kelqb@xmu.edu.cn (Q. Li). process of aromatic ring saturation from eAQH2 to H4eAQH2 and

http://dx.doi.org/10.1016/j.cej.2014.09.117
1385-8947/Ó 2014 Elsevier B.V. All rights reserved.
H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363 357

Scheme 1. Reaction routes for hydrogenation of eAQ.

2-ethyl-1,2,3,4,5,6,7,8-octahydro-9,10-anthrahydroquinone (H8- Hydrogen peroxide (3 wt.%) was purchased from Beijing J&K Scien-
eAQH2). And the reaction pathway II comprises the hydrogenolysis tific Co., Ltd., China. 2-ethylanthraquinone (99%) was bought from
reactions of C–O bonds from eAQH2, such as 2-ethyloxanthrone Henan Hongye Chemical Co., Ltd., China. The Cacumen Platycladi
(OXO-E), 2-ethylanthrone (eAN), 2-ethylanthracene (eANT) and biomass and c-Al2O3 were purchased from Xiamen Jiuding Drug-
dianthrone, etc. [11–16]. It has been demonstrated that the selectiv- store and CNOOC Tianjin Chemical Research & Design Institute,
ity of the desired reaction (i.e., hydrogenation of quinone to hydro- China, respectively.
quinone) should be above 99% in industrial practice [3]. This
clearly shows that the side reactions at the cost of active anthraqui- 2.2. Preparation of CP extract and measurement of main components
nones must be strongly inhibited. However, even for the active Pd
catalysts prepared by chemical reduction methods (impregnation, In order to obtain CP extract, the carefully weighted sundried
H2 reduction, etc.), the consumption of eAQH2 to produce degrada- biomass, which was screened with a 50 mesh sieve, 1.0 g was
tion products is inevitable [9,10,17]. Thus, it is imperative to develop added to 100 mL deionized water in a conical flask of 250 mL
an efficient and durable catalyst with high selectivity in the hydro- capacity. The mixture was thereafter shaken (150 rpm, 2 h) at
genation of the carbonyl group rather than the aromatic ring. 30 °C and filtrated to get filtrate, i.e. the CP extract. The extract
Green plant-mediated biosynthesis of metal nanoparticles (NPs) was reserved in a refrigerator (4 °C) for further experiments. The
in which plant extract played dual roles as reductive and protective reducing sugar, flavonoid, polysaccharides and protein contents
agents has been regarded as an alternative to chemical methods in before and after the reaction were measured by the same methods
the last decade [18–22]. Interestingly, the biosynthetic metal NPs over in our previous study [33].
proper supports exhibited high activity and stability toward some
reactions [23–31]. As far as biosynthetic Pd NPs were concerned, 2.3. Preparation of Pd/c-Al2O3 catalysts
few reports were devoted to their catalytic application. In our previ-
ous study, the biogenic Pd NPs by antioxidants in Gardenia jasminoides The original support c-Al2O3 was calcined at 500 °C prior to its
Ellis as long lifetime nanocatalysts for p-nitrotoluene hydrogenation use. Then c-Al2O3 was soaked in 0.5 M HCl, 0.5 M NaOH, 0.05 M
was demonstrated. Very recently, biosynthesis of Pd NPs using Delo- Na2CO3 or 0.007 M Na2SiO3 solution under stirring for 2 h, respec-
nix regia leaf extract and their catalytic applications to nitro-aromatics tively. Next, the slurry was filtered, washed and dried to obtain
hydrogenation were reported [32]. Herein, we aimed at expanding pretreated c-Al2O3 for further experiments. In a typical preparation
the green plant-mediated strategy to highly active and stable Pd nan- of Pd/c-Al2O3 catalyst, an adsorption–reduction method in which
ocatalysts for the liquid-phase hydrogenation of eAQ. In this work, Pd/ Pd(NO3)2 adsorbed onto c-Al2O3 support was reduced with the
c-Al2O3 catalysts were prepared via an adsorption-reduction method CP extract was adopted. Firstly, 0.5 g support was immersed in
with Cacumen Platycladi (CP) extract without calcination. In order to aqueous Pd(NO3)2 (30 mL, 0.78 mM) for 0.5 h. Then, 30 mL of CP
optimize the catalytic performance, effect of reaction temperature, extract was added to the solution under stirring. After 1 h, the
hydrogenated time, working solution volume and H2 flow rate was slurry was filtered, and the solid (0.5 wt.% Pd/c-Al2O3 catalysts)
investigated. In particular, pretreatments of c-Al2O3 support with was washed thoroughly with distilled water, dried in vacuum oven
HCl, NaOH, Na2CO3 and Na2SiO3 solutions on the catalytic perfor- at 50 °C overnight (without calcination) and their catalytic activity
mance were also studied. The Pd/c-Al2O3 catalysts were characterized for hydrogenation of 2-eAQ was then tested. The actual Pd loading
by N2 physisorption (BET), X-ray diffraction (XRD), transmission elec- of the catalyst was analyzed by AAS (Pgeneral, China) by dissolving
tron microscopy (TEM) and X-ray photoelectron spectroscopy (XPS). the catalyst in aqua regia for 24 h.
And atomic absorption spectrophotometry (AAS) was used to mea-
sure the actual Pd loading of the catalyst. 2.4. Characterizations of Pd/c-Al2O3 catalysts

2. Materials and methods N2 physisorption isotherms were recorded on a porosimetry


analyzer (Micromeritics TriStar 3000) at 77 K. Before measure-
2.1. Materials and reagents ments, Pd/c-Al2O3 samples were outgassed at 300 °C for 3 h. The
TEM observation was performed on a Tecnai F30 microscope at
Palladium nitrate (AR), 2-octanol (AR) and xylol (AR) were 300 keV with a high-angle annular dark field detector (HAADF).
purchased from Sinopharm Chemical Reagent Co., Ltd., China. The average size and size distribution of Pd NPs were measured
358 H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363

on the basis of TEM images by the previous method [33]. Powder Table 1
X-ray diffraction (XRD) patterns were performed on an X’pert Structural properties of c-Al2O3 before and after pretreatments.

PRO diffractionmeter at 40 kV and a current of 30 mA with Cu- Sample SBET/m2 g1 VP/cm3 g1 DP/nm
Ka radiation (k = 1.5418 Å). X-ray photoelectron spectroscopy Al2O3 208 0.35 3.3
(XPS) was performed on a Quantum 2000 spectrometer, and the Al2O3–HCl 245 0.35 2.8
Al-Ka line was used as the excitation source. Al2O3–NaOH 230 0.33 2.9
Al2O3–Na2CO3 237 0.32 2.7
Al2O3–Na2SiO3 227 0.32 2.8
2.5. Hydrogenation of 2-ethylanthraquinone and analytical procedure

Firstly, the catalytic activity for hydrogenation of 2-eAQ was That is, when the sum of n(eAQ)/n0(eAQ) and nðH4 eAQ Þ /n0(eAQ) was
tested in agitated glass reactor at 40 °C using xylene and 2-octanol much closer to 1, the content of degradation products was much
(volume ratio 1:1) as mixing solvent, with constant eAQ concentra- less.
tion 20 g/L, H2 flow rate 30 mL/min, catalyst 0.1 g. Subsequently,
about the solution volume 3 mL was oxidized by oxygen (flow rate:
40 mL/min) for about 15 min in the other glass reactor without agi- 3. Results and discussion
tation. By observing the color of the solution, one can determine
whether the oxidization was complete or not. In the previous stud- 3.1. Characterization of c-Al2O3 and the bioreduced Pd/c-Al2O3
ies, the hydrogen peroxide amount in the aqueous extract was catalysts
determined by titration with standard potassium permanganate
solution [1,6,8,11,13–15,34,35]. However, this method caused a Nitrogen physisorption was used to evaluate pore structure
large error in our system because some biological components parameters of c-Al2O3 before and after pretreatments, including
might be extracted and oxidized by potassium permanganate. Thus, specific area (SBET), pore volume (VP) and pore diameter (DP).
in this work, H2O2 amount in the aqueous extract was analyzed by Fig. 1a and b shows the N2 adsorption/desorption isotherms and
HPLC using a chromatograph (Agilent, USA) with an UV detector pore size distributions, respectively. As shown, all the samples fol-
(200 nm), column 25 cm long, C18 fraction dp = 5 lm, and mixing lowed Langmuir type IV isotherms, revealing that they were typi-
eluent composed of 80% acetonitrile and 20% ultrapure water. The cal mesoporous materials. The pore size distribution (Fig. 1b)
retention time of standard hydrogen peroxide was about 2.43 min, showed a shoulder peak of pore size at about 2.9 nm after pretreat-
and the peak of hydrogen peroxide was well identified. ment, indicating some small mesopores were created. In addition,
The yield of total hydrogenated anthraquinones was deter- pore structure parameters in Table 1 indicated that the SBET slightly
mined by the following equations: increased while DP decreased after pretreatment.
Y H2 O2 =% ¼ nðH2 O2 Þ =n0ðH2 O2 Þ  100% ¼ nðH2 O2 Þ =n0ðeAQ Þ  100% ð1Þ As far as the preparation of Pd/c-Al2O3 catalysts, Pd(II) ions
were firstly adsorbed on the c-Al2O3 support, then CP extract
Average yield ¼ n0ðH2 O2 Þ  Y H2 O2  M H2 O2 =ðmPd  hÞ ð2Þ was introduced into reduce Pd(II) ions. In order to clarify the mech-
anism of bio-reduction, the contents of flavonoids, polysaccharides,
where n0ðH2 O2 Þ and nðH2 O2 Þ are the initial and the obtained mole of reducing sugars and proteins in the extract before and after the
H2O2, respectively. MH2 O2 and mPd represent the H2O2 molar mass reaction were measured. Fig. S1 (Supporting Information) shows
(34 g/mol) and the Pd (g) mass in the catalyst, respectively. The con- that all the contents of the flavonoids and reducing sugars (actually
centrations of both eAQ and H4eAQ were determined with a gas included in saccharides) decreased largely, especially the content
chromatograph (GC 9160, Shanghai Ouhua, China) equipped with of flavonoids. Hence, these flavonoids and reducing sugars in the
FID detector in conditions: column FFAP, 0.53 mm  50 m  extract were responsible for the reduction of Pd(II). The morphol-
0.4 lm, isothermal run 245 °C, which were analyzed by internal ogy of Pd/c-Al2O3 catalysts with theoretical Pd loading of
standard method. Dibutyl phthalate was chosen as an internal stan- 0.5 wt.% was obtained by HAADF-STEM in Fig. 2a, in which bright
dard substance. The sum of n(eAQ) and nðH4 eAQ Þ was smaller than the spots (Orange box) represented the Pd NPs within size range of
initial content of n0(eAQ). And the percentage of degradation prod- 4–10 nm. Even after 8-run reuse, Pd NPs were still well distributed
ucts (ND) over n0(eAQ) could be calculated according to the law of over the support and did not aggregate (Fig. 2b). According to the
mass conservation: AAS results, the actual Pd loading was 0.41 wt.%, which was
ND ¼ 1  nðeAQ Þ =n0ðeAQ Þ  nðH4 eAQ Þ =n0ðeAQ Þ ð3Þ slightly lower than the theoretical one. XPS technique was further
used to determine the electronic state of Pd, and the spectra of Pd

Fig. 1. (a) Nitrogen adsorption/desorption isotherms and (b) the corresponding pore size distributions of c-Al2O3 before and after pretreatments.
H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363 359

Fig. 2. STEM images of 0.5 wt.% Pd catalysts on the support pretreated by Na2SiO3 solution (a) before reaction and (b) after 8-run reuse.

Fig. 3. (a) XPS spectra of Pd 3d regions recorded on support adsorbed Pd(NO3)2 (a0 ), bioreduced catalyst with 0.5 wt.% Pd loading (b0 ) before reaction and (c0 ) after reaction, (b)
Powder XRD patterns of catalysts with 0.5 wt.% Pd loading on pure c-Al2O3 and various pretreated supports.

3d region are presented in Fig. 3a. It was evident that a doublet of increased (Fig. 4b). Thus, the optimal reaction temperature was
the Pd 3d peak consists of the 3d5/2 and 3d3/2 peaks. When 40 °C, which was lower than those in the previous studies
Pd(NO3)2 was adsorbed by the support at the beginning, the bind- [6,38,39].
ing energy of Pd 3d5/2 peak was around 336.5 eV (a0 ), indicating
that Pd(II) had not yet been reduced. The shift of the Pd 3d5/2 peak 3.2.2. Effect of hydrogenated time
to a lower position at 335.6 eV (b0 ) indicated that Pd(II) ions were With the reaction time going on, a consecutive process of aro-
reduced to elemental Pd [36,37], and the position of the peak little matic ring saturation eAQH2 ! H4 eAQH2 ! H8 eAQH2 or hydrog-
changed after the recycling test. Different from the 3d5/2 peak, the enolysis reactions of C-O bonds to generate degradation products
3d3/2 peak possessed higher binding energy, i.e. 340.85 eV. Fig. 3b occurred [11]. Fig. 5 illustrates the effect of hydrogenation time
shows XRD patterns of the support before and after pretreatment. on the H2O2 yield and the active anthraquinone content. When
All the supports presented predominant peaks (2h = 36.2°, 45.4°, the hydrogenation time was 1.5 h, the H2O2 yield reached the high-
66.5°) of the c-Al2O3, while the Bragg reflections of Pd were not est value of 85.6%. Nevertheless, as the reaction time proceeded,
distinct because of alumina diffractions. the average yield decreased accordingly. When the hydrogenation
time was just 0.5 h, the quite high average yield of 95.8 g H2 O2 /
3.2. Catalytic activity studies (gPd h) was achieved. However, the H2O2 yield was only 83.2%.
Accordingly, ND increased with the hydrogenation time going on.
3.2.1. Effect of reaction temperature
The hydrogenation of eAQ was carried out in a 3-phase system. 3.2.3. Effect of working solution volume
And the rate of mass-transport process was associated with the Comparative experiments were also conducted to study the
reaction parameters (reaction temperature, hydrogenated time, effect of working solution volume on the H2O2 yield and the active
working solution volume and H2 flow rate). First, as this reaction anthraquinone content. As shown in Fig. 6, when the working solu-
was exothermic, the temperature had a significant effect on the tion volume was 10 mL, the highest catalytic activity was achieved.
eAQ conversion. The H2O2 yield and the evolution of eAQ and By contrast, when it was much lower than 10 mL, deep hydrogena-
H4eAQ as a function of reaction temperature in the range of 30– tion as well as generation of degradation products were resulted
70 °C are plotted in Fig. 4a and b, respectively. As shown, the max- because of much less eAQ and effective intermediate, given the
imum H2O2 yield of 85.6% and average H2O2 yield of 32.9 g H2 O2 / same H2 flow rate. In this case, ND was about 10%. On the other
(gPd h) were attained at 40 °C, respectively, while the eAQ conver- hand, when the working solution volume was larger than 10 mL,
sion was 90%. If the temperature increased, the H2O2 yield did not ND became much lower, while the H2O2 yields were less as well.
increase further while the percentage of degradation products ND According to equation (2), n0ðH2 O2 Þ were much higher when the
360 H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363

Fig. 4. Correlations of (a) the H2O2 yield and (b) the active anthraquinone content with reaction temperature, respectively. (Conditions: 0.5 wt.% Pd/c-Al2O3 catalysts 0.1 g,
hydrogenated time 1.5 h, working solution volume 10 mL, H2 flow rate 30 mL/min.)

Fig. 5. Correlations of (a) the H2O2 yield and (b) the active anthraquinone content with hydrogenation time, respectively. (Conditions: 0.5 wt.% Pd/c-Al2O3 catalysts 0.1 g,
temperature 40 °C, working solution volume 10 mL, H2 flow rate 30 mL/min.)

Fig. 6. Correlations of (a) the H2O2 yield and (b) the active anthraquinone content with working solution volume, respectively. (Conditions: 0.5 wt.% Pd/c-Al2O3 catalysts
0.1 g, hydrogenated time 1.5 h, temperature 40 °C, H2 flow rate 30 mL/min.)

working solution volume was larger than 10 mL, leading to higher were carried out above stoichiometric line (Fig. S2). Then the
average yield though the corresponding yields were lower. How- amount of H2, that is, the H2 flow rate had some effect on the
ever, the concentration of catalysts was changed and simulta- H2O2 yield and the active anthraquinone content, given the same
neously the gas–liquid surface was also changed by changing the reaction time (1.5 h). It can be seen from Fig. 7 that the H2O2 yield
volume of working solution. Thus, the mass-transport effect might increased slightly with increasing the flow rate of hydrogen gas at
be complicated and remains unclear. Therefore, it was necessary to the beginning. When the H2 flow rate exceeded 30 mL/min, the
control the working solution volume when the H2 flow rate was H2O2 yield, however, began to decrease. At the same time, ND
constant. started to increase. According to a previous report, hydrogen was
firstly adsorbed onto the Pd and subsequently reacted with eAQ
3.2.4. Effect of H2 flow rate [9]. If H2 flow rate was large enough, more H2 flowed out of the
It should be mentioned that the hydrogenation of quinone to liquid phase prior to its dissociation on Pd surface and did not par-
hydroquinone is very easy and commonly limited by mass transfer ticipate in the reaction, thus leading to the decrease in yield and
of H2 [6,12,40]. Thus, in this work, the hydrogenation experiments increase in ND. Thus it was quite significant to maintain the H2 flow
H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363 361

Fig. 7. Correlations of (a) the H2O2 yield and (b) the active anthraquinone content with H2 flow rate, respectively. (Conditions: 0.5 wt.% Pd/c-Al2O3 catalysts 0.1 g,
hydrogenated time 1.5 h, working solution volume 10 mL, temperature 40 °C.)

rate in the 3-phase system of hydrogenation of eAQ. By contrast, of Pd loading on the Na2SiO3-pretreated support was further stud-
the ring hydrogenation and C–O hydrogenolysis, are slower reac- ied. As shown in Fig. 9a, the H2O2 yield significantly increased with
tions so that the mass transfer of H2 had not so much important theoretical Pd loading increasing from 0.1 to 0.5 wt.%. If further
[12]. increasing the Pd loading, the yield did not increase significantly
but decreased slightly. According to the TEM images in Fig. 9b
and c, when Pd loading was as low as 0.3 wt.%, Pd NPs were too
3.2.5. Effect of pretreatment of the support
small to observe, even with HADDF-STEM. However, when Pd load-
The pretreatments on c-Al2O3 also influenced the catalytic
ing reached 1.6 wt.%, well dispersed Pd NPs on the support were
activity of the catalysts. In previous work, Drelinkiewicz et al.
observed with the bright field TEM, and the size was much larger
had done a lot of work on the pre-impregnation of alumina with
compared with the Pd NPs in the case of 0.5 wt.% Pd loading, as
Na2SiO3 [11,35], while we focused on catalytic activities of our
seen in Fig. 2. When Pd loading was below 0.5 wt.%, internal sur-
green catalysts with various pretreatments. Compared with the
face of catalyst (Pd NPs may be located inside the channel of sup-
unpretreated support, the catalysts using c-Al2O3 pretreated with
port according to Table 1) was inefficiently utilized in the main
HCl, NaOH, Na2CO3 and Na2SiO3 solutions, showed much better
reaction eAQ to eAQH2, thus leading to lower H2O2 yield. Neverthe-
catalytic activities, as shown in Fig. 8a. Especially, the Na2SiO3-pre-
less, when Pd loading was over 0.5 wt.%, larger size of Pd NPs was
treated catalyst achieved the highest H2O2 yield of 96.4%, a little bit
obtained, giving rise to less H2O2 yield. According to the active
higher than that of NaOH-pretreated catalyst, which was about
anthraquinone content, ND was much higher than ND of 0.5 wt.%
10% higher than that achieved with the unpretreated catalyst. It
Pd loading (data not shown). This may be due to deep hydrogena-
was reported that acid-base sites of the catalyst, which can act as
tion or hydrogenolysis reactions if much higher Pd loading was
adsorption sites for eAQ molecules, played an important role in
adopted. Therefore, the main reaction strongly depended on the
the hydrogenation of eAQ [9]. Drelinkiewicz et al. reported the role
Pd particle sizes, which was consistent with the previous report
of alkali modifiers was associated with stronger interactions
[35].
between the catalyst and quinone reagents, and degradation prod-
ucts are remarkably inhibited. The acidic sites of the support acted
as adsorption sites for the aromatic compound, and once these
3.2.7. Durability of the Pd/c-Al2O3 catalysts and the comparison with
adsorbed species were activated, they can react with hydrogen
other Pd catalysts
spilt from the Pd metal [9]. On the other hand, the formation of
In order to assess the durability of the bioreduced Pd catalyst,
degradation products was strongly inhibited on these pretreated
recycle tests were conducted using the Na2SiO3-pretreated cata-
catalysts, as shown in Fig. 8b.
lyst. The catalyst was washed with ethanol after reaction, then
dried and reused for the next reaction under the same condition.
3.2.6. Effect of Pd loading As shown in Fig. 10a, the H2O2 yield remained stable during eight
As discussed above, the maximum H2O2 yield was obtained if cycles, and the content of degradation products did not increase
the catalyst support was pretreated with Na2SiO3. Thus, the effect significantly (Fig. 10b), indicating that bioreduced Pd catalyst

Fig. 8. Effect of pretreatment of the support on (a) the H2O2 yield and (b) the active anthraquinone content, respectively. (Conditions: 0.5 wt.% Pd/c-Al2O3 catalysts 0.1 g,
hydrogenated time 1.5 h, working solution volume 10 mL, H2 flow rate 30 mL/min and temperature 40 °C.)
362 H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363

Fig. 9. (a) Effect of Pd loading of c-Al2O3–Na2SiO3 support on the H2O2 yield and TEM images of Pd/c-Al2O3–Na2SiO3 with Pd loading of (b) 0.3 wt.% and (c) 1.6 wt.%,
respectively. (Conditions: Pd/c-Al2O3–Na2SiO3 catalysts 0.1 g, hydrogenated time 1.5 h, working solution volume 10 mL, H2 flow rate 30 mL/min and temperature 40 °C.) The
insert histogram in (c) indicates the corresponding size distribution.

Fig. 10. The catalytic performance of recycling bioreduced Pd catalysts: (a) the H2O2 yield and (b) the active anthraquinone content, respectively. (Conditions: 0.5 wt.% Pd/c-
Al2O3–Na2SiO3 catalysts 0.1 g, hydrogenated time 1.5 h, working solution 10 mL, H2 flow rate 40 mL/min and temperature 40 °C.)

showed high catalytic stability. In addition, even after 8-run reuse, working solution 10 mL and H2 flow rate 40 mL/min. In particular,
the Pd NPs almost retained stable on the support, as illustrated by if the supports were pretreated with HCl, NaOH, Na2CO3 and Na2-
the TEM image in Fig. 2b. According to the AAS result, little Pd suf- SiO3 solutions, the H2O2 yield could increase by about 5–10%.
fered leaching during the reaction and the actual Pd loading remain And the best H2O2 yield of 96.4% could be achieved when the sup-
0.40 wt.% after 8-run reuse. For comparison, the catalytic activity of port was pretreated with Na2SiO3 solution, and even after 8-run
Pd/c-Al2O3 catalyst prepared by H2 reduction was also studied. reuse, the catalysts exhibited remarkable durability and the con-
However, the results showed that the H2O2 yield was only about tent of side products did not increase significantly. Also, Pd NPs
50%. Therefore, the bioreduced Pd/c-Al2O3 catalyst herein exhib- on the support were stable and did not aggregate after reuse.
ited much better catalytic performance. Furthermore, comparisons Therefore, in terms of these advantages, the catalysts assisted by
of the catalytic performance between the bioreduced Pd/c-Al2O3 plant-mediated strategy may be promising for future industrial
catalyst and other Pd catalysts reported in literatures were also application.
carried out [3,6]. Zhang et al. reported a Pd/Al2O3/cordierite could
obtain average yield of 15.16 g H2 O2 /(gPd h) [6], while Chen listed
Acknowledgements
APC-Q-1 spherical Pd catalyst could get more than 66.67 g H2 O2 /
(gPd h) compared with a commercial Pd catalyst could achieve
This work was supported by the Natural Science Foundation of
more than 45.83 g H2 O2 /(gPd h) [3]. Obviously, the catalytic perfor-
China (21036004 and 21106117) and Research Fund for the Doctoral
mance of the bioreduced Pd catalysts was comparable to other
Program of Higher Education of China (20110121120018). The
Pd catalysts. Moreover, when the hydrogenation time was only
authors are grateful to Dr. Xianxue Li for his helpful discussion.
0.5 h for one-way reaction, the average yield of 101.3 g H2 O2 /(gPd h)
could be achieved, and the H2O2 yield was 80.3%.
Appendix A. Supplementary data
4. Conclusion
Supplementary data associated with this article can be found, in
In summary, the uncalcined Pd/c-Al2O3 catalysts for the liquid- the online version, at http://dx.doi.org/10.1016/j.cej.2014.09.117.
phase hydrogenation of 2-ethyl-9,10-anthraquinone were pre-
pared via an adsorption–reduction method using Cacumen Platycla-
References
di extract. Based on the catalyst with the Pd loading of 0.5 wt.%, the
optimum reaction conditions were reaction temperature 40 °C [1] R. Halder, A. Lawal, Experimental studies on hydrogenation of anthraquinone
(lower than those in the literature), hydrogenated time 1.5 h, derivative in a microreactor, Catal. Today 125 (2007) 48–55.
H. Chen et al. / Chemical Engineering Journal 262 (2015) 356–363 363

[2] Y. Hou, Effects of lanthanum addition on Ni-B/c-Al2O3 amorphous alloy [23] L.S. Jia, Q. Zhang, Q.B. Li, H. Song, The biosynthesis of palladium nanoparticles
catalysts used in anthraquinone hydrogenation, Appl. Catal. A: Gen. 259 (2004) by antioxidants in Gardenia jasminoides Ellis: long lifetime nanocatalysts for p-
35–40. nitrotoluene hydrogenation, Nanotechnology 20 (2009) 385601.
[3] Q. Chen, Development of an anthraquinone process for the production of [24] Y.L. Hong, X.L. Jing, J.L. Huang, D.H. Sun, T. Odoom-Wubah, F. Yang, M.M. Du,
hydrogen peroxide in a trickle bed reactor-from bench scale to industrial scale, Q.B. Li, Biosynthesized bimetallic Au–Pd nanoparticles supported on TiO2 for
Chem. Eng. Process. Process Intensif. 47 (2008) 787–792. solvent-free oxidation of benzyl alcohol, ACS Sustainable Chem. Eng. 2 (2014)
[4] J. Petr, L. Kurc, Z. Belohlav, L. Cerveny, Catalytic hydrogenation of 2-ethyl-9,10- 1752–1759.
anthrahydroquinone, Chem. Eng. Process. 43 (2004) 887–894. [25] G.W. Zhan, J.L. Huang, M.M. Du, D.H. Sun, I. Abdul-Rauf, W.S. Lin, Y.L. Hong,
[5] A. Drelinkiewicz, Kinetics and mechanism of hydrogenation of 2-ethyl-9,10- Q.B. Li, Liquid phase oxidation of benzyl alcohol to benzaldehyde with novel
anthraquinone in liquid phase. I. Introductory study, Bull. Pol. Acad. Sci. Chem. uncalcined bioreduction Au catalysts: high activity and durability, Chem. Eng.
39 (1991) 63–72. J. 187 (2012) 232–238.
[6] J. Zhang, D. Li, Y. Zhao, Q. Kong, S. Wang, A Pd/Al2O3/cordierite monolithic [26] M.M. Du, G.W. Zhan, X. Yang, H.X. Wang, W.S. Lin, Y. Zhou, J. Zhu, L. Lin, J.L.
catalyst for hydrogenation of 2-ethylanthraquinone, Catal. Commun. 9 (2008) Huang, D.H. Sun, L.S. Jia, Q.B. Li, Ionic liquid-enhanced immobilization of
2565–2569. biosynthesized Au nanoparticles on TS-1 toward efficient catalysts for
[7] Y.J. Hou, Y.Q. Wang, F. He, S. Han, Z.T. Mi, W. Wu, E.Z. Min, Liquid phase propylene epoxidation, J. Catal. 283 (2011) 192–201.
hydrogenation of 2-ethylanthraquinone over La-doped Ni-B amorphous alloy [27] J.L. Huang, C. Liu, D.H. Sun, Y.L. Hong, M.M. Du, T. Odoom-Wubah, W.P. Fang, Q.
catalysts, Mater. Lett. 58 (2004) 1267–1271. Li, Biosynthesized gold nanoparticles supported over TS-1 toward efficient
[8] B. Liu, M.H. Qiao, J.Q. Wang, K.N. Fan, Highly selective amorphous Ni–Cr–B catalyst for epoxidation of styrene, Chem. Eng. J. 235 (2014) 215–223.
catalyst in 2-ethylanthraquinone hydrogenation to 2- [28] D.H. Sun, G.L. Zhang, X.D. Jiang, J.L. Huang, X.L. Jing, Y.M. Zheng, J. He, Q.B. Li,
ethylanthrahydroquinone, Chem. Commun. (2002) 1236–1237. Biogenic flower-shaped Au–Pd nanoparticles: synthesis, SERS detection and
[9] R. Kosydar, A. Drelinkiewicz, E. Lalik, J. Gurgul, The role of alkali modifiers (Li, catalysis towards benzyl alcohol oxidation, J. Mater. Chem. A 2 (2013) 1767–
Na, K, Cs) in activity of 2% Pd/Al2O3 catalysts for 2-ethyl-9,10-anthraquione 1773.
hydrogenation, Appl. Catal. A: Gen. 402 (2011) 121–131. [29] V. Reddy, R.S. Torati, S. Oh, C. Kim, Biosynthesis of gold nanoparticles assisted
[10] A. Drelinkiewicz, A. Waksmundzka-Gora, W. Makowski, J. Stejskal, Pd/ by Sapindus mukorossi Gaertn. Fruit pericarp and their catalytic application for
polyaniline(SiO2) a novel catalyst for the hydrogenation of 2- the reduction of p-nitroaniline, Ind. Eng. Chem. Res. 52 (2012) 556–564.
ethylanthraquinone, Catal. Commun. 6 (2005) 347–356. [30] G.W. Zhan, Y.L. Hong, F.F. Lu, A.R. Ibrahim, M.M. Du, D.H. Sun, J.L. Huang, Q.B.
[11] A. Drelinkiewicz, A. Pukkinen, R. Kangas, R. Laitinen, Hydrogenation of 2- Li, J. Li, Kinetics of liquid phase oxidation of benzyl alcohol with hydrogen
ethylanthraquinone over Pd/SiO2 and Pd/Al2O3 in the fixed-bed reactor. The peroxide over bio-reduced Au/TS-1 catalysts, J. Mol. Catal. A: Chem. 366 (2013)
effect of the type of support, Catal. Lett. 94 (2004) 157–170. 215–221.
[12] R. Kosydar, A. Drelinkiewicz, J.P. Ganhy, Degradation reactions in [31] G.W. Zhan, Y.L. Hong, V.T. Mbah, J.L. Huang, A.R. Ibrahim, M.M. Du, Q.B. Li,
anthraquinone process of hydrogen peroxide synthesis, Catal. Lett. 139 Bimetallic Au–Pd/MgO as efficient catalysts for aerobic oxidation of benzyl
(2010) 105–113. alcohol: a green bio-reducing preparation method, Appl. Catal. A: Gen. 439
[13] A. Drelinkiewicz, R. Laitinen, R. Kangas, J. Pursiainen, 2-Ethylanthraquinone (2012) 179–186.
hydrogenation on Pd/Al2O3 – the effect of water and NaOH on the degradation [32] P. Dauthal, M. Mukhopadhyay, Biosynthesis of palladium nanoparticles using
process, Appl. Catal. A Gen. 284 (2005) 59–67. Delonix Regia leaf extract and its catalytic activity for nitro-aromatics
[14] A. Drelinkiewicz, A. Waksmundzka-Gora, Hydrogenation of 2-ethyl-9,10- hydrogenation, Ind. Eng. Chem. Res. 52 (2013) 18131–18139.
anthraquinone on Pd/SiO2 catalysts – the role of humidity in the [33] J.L. Huang, G.W. Zhan, B.Y. Zheng, D.H. Sun, F.F. Lu, Y. Lin, H.M. Chen, Z.D.
transformation of hydroquinone form, J. Mol. Catal. A: Chem. 258 (2006) 1–9. Zheng, Y.M. Zheng, Q.B. Li, Biogenic silver nanoparticles by Cacumen Platycladi
[15] A. Drelinkiewicz, A. Waksmundzka-Gora, J.W. Sobczak, J. Stejskal, extract: synthesis, formation mechanism, and antibacterial activity, Ind. Eng.
Hydrogenation of 2-ethyl-9,10-anthraquinone on Pd-polyaniline (SiO2) Chem. Res. 50 (2011) 9095–9106.
composite catalyst – the effect of humidity, Appl. Catal. A: Gen. 333 (2007) [34] F. Wang, X.L. Xu, K.P. Sun, Hydrogenation of 2-ethylanthraquinone over Pd/
219–228. ZrO2-gamma-Al2O3 catalyst, React. Kinet. Catal. Lett. 93 (2008) 135–140.
[16] A. Drelinkiewicz, Kinetic aspects in the selectivity of deep hydrogenation of 2- [35] A. Drelinkiewicz, R. Kangas, R. Laitinen, A. Pukkinen, J. Pursiainen,
ethylanthraquinone over Pd/SiO2, J. Mol. Catal. A: Chem. 101 (1995) 61–74. Hydrogenation of 2-ethylanthraquinone on alumina-supported palladium
[17] A. Drelinkiewicz, A. Waksmundzka-Gora, Investigation of 2- catalysts – the effect of support modification with Na2SiO3, Appl. Catal. A:
ethylanthraquinone degradation on palladium catalysts, J. Mol. Catal. A: Gen. 263 (2004) 71–82.
Chem. 246 (2006) 167–175. [36] M. Bunge, L.S. Søbjerg, A.E. Rotaru, D. Gauthier, A.T. Lindhardt, G. Hause, K.
[18] P.X. Zhao, N. Li, D. Astruc, State of the art in gold nanoparticle synthesis, Coord. Finster, P. Kingshott, T. Skrydstrup, R.L. Meyer, Formation of palladium (0)
Chem. Rev. 257 (2013) 638–665. nanoparticles at microbial surfaces, Biotechnol. Bioeng. 107 (2010) 206–215.
[19] D. Hebbalalu, J. Lalley, M.N. Nadagouda, R.S. Varma, Greener techniques for the [37] V.S. Coker, J.A. Bennett, N.D. Telling, T. Henkel, J.M. Charnock, G. van der Laan,
synthesis of silver nanoparticles using plant extracts, enzymes, bacteria, R.A.D. Pattrick, C.I. Pearce, R.S. Cutting, I.J. Shannon, Microbial engineering of
biodegradable polymers, and microwaves, ACS Sustainable Chem. Eng. 1 nanoheterostructures: biological synthesis of a magnetically recoverable
(2013) 703–712. palladium nanocatalyst, ACS Nano 4 (2010) 2577–2584.
[20] L.F. Wu, W.W. Wu, X.L. Jing, J.L. Huang, D.H. Sun, T. Odoom-Wubah, H.Y. Liu, [38] A. Drelinkiewicz, M. Hasik, M. Kloc, Pd/polyaniline as the catalysts for 2-
H.T. Wang, Q. Li, Trisodium citrate-assisted biosynthesis of silver nanoflowers ethylanthraquinone hydrogenation. The effect of palladium dispersion, Catal.
by Canarium album foliar broths as a platform for SERS detection, Ind. Eng. Lett. 64 (2000) 41–47.
Chem. Res. 52 (2013) 5085–5094. [39] A. Drelinkiewicz, M. Hasik, 2-Ethyl-9,10-anthraquinone hydrogenation over
[21] B.Y. Zheng, T. Kong, X.L. Jing, T. Odoom-Wubah, X.X. Li, D.H. Sun, F.F. Lu, Y.M. Pd/polymers – effect of polymers–Pd(II) chlorocomplexes interactions, Appl.
Zheng, J.L. Huang, Q. Li, Plant-mediated synthesis of platinum nanoparticles Catal. A: Gen. 177 (2001) 149–164.
and its bioreductive mechanism, J. Coll. Interf. Sci. 396 (2013) 138–145. [40] E. Santacesaria, M. Di Serio, A. Russo, U. Leone, R. Velotte, Kinetic and catalytic
[22] G.L. Zhang, M.M. Du, Q.B. Li, X.L. Li, J.L. Huang, X.D. Jiang, D.H. Sun, Green aspects in the hydrogenation peroxide production via anthraquinone, Chem.
synthesis of Au–Ag alloy nanoparticles using Cacumen platycladi extract, RSC Eng. Sci. 54 (1999) 2799–2806.
Adv. 3 (2013) 1878–1884.

You might also like