Vous êtes sur la page 1sur 4

Catalysis Communications 19 (2012) 115118

Contents lists available at SciVerse ScienceDirect

Catalysis Communications
journal homepage: www.elsevier.com/locate/catcom

Short Communication

Direct conversion of cellulose into sorbitol using dual-functionalized catalysts in


neutral aqueous solution
Joung Woo Han, Hyunjoo Lee
Department of Chemical and Biomolecular Engineering, Yonsei University, Seoul 120-749, Republic of Korea

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

Article history: Direct and selective conversion of cellulose into sorbitol was carried out over dual-functionalized catalysts
Received 3 November 2011 containing both sulfonate groups and Ru nanoparticles. A high sorbitol yield of 71.1% was obtained in a neu-
Received in revised form 23 December 2011 tral aqueous solution without an aid of liquid phase acid at an intermediate reaction temperature of 165 C
Accepted 23 December 2011
via synergy of sulfonate groups and Ru sites on the catalyst surface. No deactivation was observed even
Available online 30 December 2011
after 5th repeated reactions. The effect of reaction time, temperature, and the Ru amount was also evaluated
Keywords:
for the sorbitol production. The cellulose was decomposed by simultaneous hydrolysis and hydrogenation
Biomass producing cello-oligomers with partially hydrogenated end groups.
Cellulose 2011 Elsevier B.V. All rights reserved.
Sorbitol
Heterogeneous catalyst
Dual-functionalized catalyst

1. Introduction with yields of 25% and 6%, respectively, over Pt/-Al2O3 at 190 C
[3]. C. Luo et al. reported a 38% of cellulose conversion and a 22% of
Recently, biomass has drawn a lot of attention as an environmental- sugar alcohol yield obtained over a Ru/C catalyst in hot water at
ly friendly and sustainable resource for the production of fuel and 245 C [4]. Deng et al. reported the cracking of cellulose into sorbitol
chemical products. Because the CO2 generated by the consumption of with a 69% sorbitol yield, which is, to the best of our knowledge, the
biomass-derived chemicals can be recovered through the growth of highest value to date, over a Ru/CNT catalyst after the cellulose has
more biomass, biomass is called as a carbon-neutral resource. Unlike ed- been pretreated with phosphoric acid at a reaction temperature of
ible types of biomass such as starch or oil, which has raised ethical issues, 185 C [5]. Generally, the conversion of cellulose into sorbitol necessi-
cellulosic biomass cannot be digested by humans and is very abundant in tates both acid and metal catalysts, for hydrolysis and hydrogenation,
nature [1]. Therefore, cellulose is the most promising natural resource for respectively. Liquid-phase mineral acids or heteropolyacids have been
the conversion into more valuable chemicals. used together with metal catalysts containing Pt or Ru [8,9]. These
Cellulose can be converted into sugar alcohols [29], oxygenated acids, however, are difcult to recover and can cause corrosion of the re-
bio-oil [10], and hydrocarbons [11] by various chemical transforma- actor. In addition, a large amount of waste sludge is formed in the acid
tions. Especially, sorbitol, which is the hydrogenated form of glucose, neutralization process. Water at temperatures above 190 C, which
was targeted in this work because it is a good model system to study can donate an H+ ion reversibly [4,6], was used for hydrolysis as well.
both hydrolysis and hydrogenation. Sorbitol is widely used as a But it is better to avoid such high reaction temperatures to minimize
sweetener, a moisture controller in cosmetics, and in medical applica- the energy input.
tions. It also has been studied as resource for the production of hydro- In this study, dual-functionalized catalysts were developed by de-
gen, alkane, and value-added chemicals such as ethylene glycol and positing Ru nanoparticles on carbon supports treated with sulfuric
propylene glycol [1214]. acid. The sulfonate groups were shown to be effective for the hydro-
Since Yan et al. presented the conversion of cellobiose into sorbitol lysis of cellulose to form glucose [1517]. Ru nanoparticles, which
over an acidic Ru colloidal solution [2], several groups have also act as catalysts for hydrogenation, were deposited on these carbon
reported the conversion of cellulose into sorbitol. Fukuoka et al. supports by chemical reduction method to preserve the surface sulfo-
reported that cellulose can be cracked into sorbitol and mannitol nate groups. The conventional calcination method would destroy the
surface functional groups. The prepared dual-functionalized catalysts
produced sorbitol directly from cellulose via a one-pot process with a
Corresponding author. Tel.: + 82 2 2123 5759; fax: + 82 2 312 6401. high yield, and the reaction proceeded in a neutral aqueous solution
E-mail address: azhyun@yonsei.ac.kr (H. Lee). at an intermediate temperature.

1566-7367/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.catcom.2011.12.032
116 J.W. Han, H. Lee / Catalysis Communications 19 (2012) 115118

2. Experimental section
a) b)
2.1. Catalyst preparation

AC-SO3H was prepared by treating 300 mg of activated carbon


(AC, NORIT) with 30 ml of concentrated sulfuric acid (96%, Duksan)
in a three-neck ask with mild stirring (500 rpm) and reuxing at
200 C. After 24 h, the AC-SO3H was ltered and washed with plenty
of water until the ltrate pH became 7. Then, the washed AC-SO3H
was dried in a convection oven at 80 C. To obtain 10 wt.% Ru/
AC-SO3H, 100 mg of dried AC-SO3H, 43.8 mg of Ru(acac)3 (Aldrich, 50nm 2nm
97%) and 20 ml of ethylene glycol (Aldrich, 99%) were put in a
three-neck ask. The ask was heated to 240 C for 1.5 h with vigor-
ous stirring and reuxing. Then, the solution was washed with ethyl
c) S 2p d) Ru 3p3/2
alcohol (Duksan, 99.9%), ltered, and dried in a convection oven. Pt/
AC-SO3H, Pd/AC-SO3H, and Ni/AC-SO3H were prepared by using the
same method except the metal precursors (Pt(acac)2 (Aldrich, 97%),
Pd(acac)2 (Aldrich, 99%) and Ni(acac)2 (Aldrich, 95%) used instead).
The prepared catalysts were analyzed by TEM (Philips Tecnai 20, ac-
celeration voltage: 200 kV) and XPS (Thermo VG Escalab 220i-XL).

2.2. Hydrogenation reactions of cellulose

The cellulose hydrogenation reactions were carried out in a stain- Fig. 1. (a) TEM, (b) HR-TEM images, and XPS spectra for (c) S 2p and (d) Ru 3p3/2 of
less steel autoclave reactor (Hanwoul Eng). 50 mg of ball-milled cel- 10 wt.% Ru/AC-SO3H catalysts.
lulose, 20 mg of the prepared catalyst, and 12 ml of water were
introduced into the reactor, and the reactor was purged with 50 bar
of hydrogen gas. Then, the reactor was heated to 165 C with stirring In the cases in which no catalyst was used (entry 1), or activated
at 450 rpm. After the reaction, the product solution was centrifuged, carbon (entry 2) was used, no sugar alcohols were observed, although
and the supernatants were ltered. The ltered water-soluble prod- conversions were approximately 40%. The cellulose appeared to be
ucts were analyzed by high-performance liquid chromatography converted into cello-oligomers that dissolved in the water or into gas-
(HPLC; Young Lin YL-9100 series) with an ELSD detector (Schambeck eous products such as CO and CO2. (For the detailed results, see Table
SFD ZAM3000). The columns used in this study were an Agilent S1 and Fig. S2) When sulfonated activated carbon was used (entry 3),
Zorbax-NH2 column (mobile phase; acetonitrile:water = 3:1, 1 ml/ a yield of 11.4% glucose was obtained, but no sugar alcohols were ob-
min, 40 C) and a Waters Sugar-Pak1 column (mobile phase; water, served due to the absence of hydrogenation. In contrast, when Ru
0.4 ml/min, 85 C). The recyclability of the catalysts was tested after supported on activated carbon (Ru/AC) was used without acidic
removing the unreacted solid cellulose by dissolving in sulfuric acid sites (entry 4), a high conversion of 79.2% was observed, but no
at room temperature under otherwise the same reaction condition. sugar alcohols were obtained. In the absence of acidic groups, proper
This dissolution process did not affect the sulfonate group on the cat-
alyst. Ru and S content in the catalysts showed little difference before
and after the dissolution process.
Table 1
Conversion of the cellulose to sugar alcohols over various solid catalystsa.
3. Results and discussion
Entry Catalyst Product yield (%) Cellulose
conversion
TEM and XPS data of Ru/AC-SO3H are shown in Fig. 1. Fig. 1(a) Sorbitol Mannitol Xylitol Glucose Othersb
(%)c
shows that the size of Ru nanoparticles was approximately 10 nm.
1 No catalyst 0 0 0 0 39.4 39.4
The high-resolution transmission electron microscopy (HR-TEM)
2 Activated 0 0 0 0 41.3 41.3
image in Fig. 1(b) clearly shows a Ru nanoparticle supported on the carbon (AC)
carbon support. The X-ray photoelectron spectra (XPS) of Ru/ 3 AC-SO3H 0 0 0 11.4 49.3 60.7
AC-SO3H in Fig. 1(c) and (d) show a large S 2p peak around 168 eV 4 10 wt.% Ru/ 0 0 0 0 79.2 79.2
and a Ru 3p3/2 peak around 462 eV, indicating the presence of sulfo- AC
5 10 wt.% Ru/ 58.7 6.7 6.8 0 8.8 81.0
nate groups (SO3H) and metallic Ru, respectively [7,18]. AC-SO3H
To reduce the crystallinity of the cellulose for enhancing the contact 6 d
10 wt.% Ru/ 8.6 3.6 3.5 0 3.9 19.6
between the catalysts and the cellulose, ball-milling was performed as a AC-SO3H
e,f
pretreatment procedure. The detailed process is presented in the sup- 7 Ru NP + AC- 34.5 7.5 4.7 0 24.6 71.3
SO3H
plementary data. Fig. S1 shows that the crystallinity of the cellulose
8 10 wt.% Pt/ 51.0 4.0 4.7 0 24.5 84.2
was signicantly reduced after the ball-milling. When the cellulose AC-SO3H
obtained after 72 h of ball-milling was treated at 165 C and 50 bar of 9 10 wt.% Pd/ 2.2 0.4 0.7 9.5 57.4 70.2
H2 for 24 h with the catalysts and water, the results shown in Table 1 AC-SO3H
were obtained. The cellulose conversion was calculated based on the 10 10 wt.% Ni/ 1.8 0.4 0.3 8.7 54.9 66.1
AC-SO3H
weight difference of solid cellulose before and after the reaction. The
a
solid cellulose separated from the solution phase after the reaction Ball-milled cellulose 50 mg, catalyst 20 mg, water 12 ml, 165 C, 50 bar H2, 24 h.
b
Unidentied molecules including cello-oligomers and gaseous products.
was surely dried at 80 C overnight before the weight measurement. c
Calculated by weight difference of solid cellulose before and after the reaction.
The yields of the products in the solution were measured using high- d
Reaction with microcrystalline cellulose without ball-milling.
performance liquid chromatography (HPLC) calibrated for each sugar e
The amount of the used Ru metal was the same as the other systems.
f
alcohol and for glucose. The method for Ru nanoparticle synthesis was adapted from Yan's work [2].
J.W. Han, H. Lee / Catalysis Communications 19 (2012) 115118 117

hydrolysis did not occur at relatively moderate temperature of 165 C, a) 100 100
resulting in destructive dissociation.

Cellulose Conversion (%)


For dual-functionalized catalysts containing 10 wt.% Ru and sulfo- 80 80
nate groups (entry 5), sorbitol was obtained with a yield of 58.7% at a

Sorbitol Yield (%)


cellulose conversion of 81%. Small amounts of mannitol and xylitol
60 60
were also observed. The yield calculated for all detected sugar alcohol
molecules (sorbitol + mannitol + xylitol) was very high, at 72.2%. No
40 40
glucose was observed. When microcrystalline cellulose was used di-
rectly without the ball-milling process (entry 6), the sorbitol yield
was much lower, at 8.6%. The colloidal Ru nanoparticles were mixed 20 20
with sulfonated carbon supports and used for the conversion of cellu-
lose into sorbitol (entry 7). The sorbitol yield was relatively high, at
120 1 40 160 180
34.5%. However, it was difcult to separate the Ru nanoparticles
Temperature (oC)
(~2 nm) dispersed in the solution from the products in this case. It
should be noted that all of the reactions were performed in neutral b) 100 100
aqueous solutions. The solution pH did not differ before and after

Cellulose Conversion (%)


the reaction. The effect of different metals was also investigated. Pt/
80 80
AC-SO3H (entry 8) showed a comparable catalytic activity to Ru/

Sorbitol Yield (%)


AC-SO3H. In the cases of Pd (entry 9) and Ni (entry 10), however,
60 60
the sorbitol yield was much lower.
The stability of the dual-functionalized catalysts (Ru/AC-SO3H)
40 40
was tested by repeating the reaction several times, as shown in
Fig. 2. No leached Ru was detected in the product solution. The S con-
tent in the catalyst also showed little difference. The cellulose conver- 20 20
sion and the sorbitol yield showed no difference up to the 5th cycle.
The dual-functionalized catalysts exhibited no deactivation and
6 12 18 24 30 36 42 48 54
were stable for the conversion of cellulose into sorbitol after repeated
Reaction Time (h)
reactions.
The effect of the reaction temperature and time on the sorbitol
Fig. 3. Effect of (a) reaction temperature (for 24 h), and (b) reaction time (at 165 C)
yield was investigated, as shown in Fig. 3. The cellulose conversion in- for the conversion of cellulose into sorbitol.
creased as the temperature increased, but the sorbitol yield showed a
maximum at 165 C. The products were degraded at higher reaction
temperatures. The sorbitol yield increased with time, showing a max- S1, rather than sequentially, with total hydrolysis followed by hydro-
imum of 71.1% at 36 h. For longer reaction times, the sorbitol yield de- genation with the scheme of cellulose glucose sorbitol.
creased, although the cellulose conversion approached 100%. The Different amounts of Ru were loaded on the activated carbon sup-
sorbitol was degraded with longer reaction times. A conversion ports, and the effect of the Ru amount on the sorbitol yield was also
mechanism of cellulose into sorbitol was also investigated. The prod- studied as shown in Table S2. The use of 1 wt.% Ru/AC-SO3H yielded
ucts after 6, 12, 18, and 24 h of reaction were analyzed by liquid chro- only 7.3% sorbitol; 6.8% of glucose was also detected. In this case,
matography mass spectrometry (LCMS). As shown in Fig. S3, the the amount of Ru seemed to be insufcient to fully hydrogenate the
peaks for cello-oligomers consisting of sorbitol and glucose units products into sorbitol. Among 5, 10, and 15 wt.% Ru/AC-SO3H, the
(sorbitolglucose (m/z: 367.27) and sorbitolglucoseglucose (m/z: 10 wt.% case showed the highest sorbitol yield and the highest selec-
529.27)) were clearly observed, in addition to sorbitol. The cello- tivity for sorbitol. These results indicate that a balanced Ru-to-S ratio
oligomer peaks were much bigger in the early stage of the reaction is important for the effective conversion of cellulose into sorbitol.
but became smaller over time and then almost disappeared after
24 h. This observation indicates that the cellulose polymer undergoes
hydrolysis and hydrogenation simultaneously, as proposed in Scheme 4. Conclusions

A dual-functionalized catalyst containing both acidic groups and


metal active sites for hydrolysis and hydrogenation, respectively,
was prepared. This catalyst was very effective for the direct and selec-
100 Sorbitol Yield tive conversion of cellulose into sorbitol with a maximum yield of
Cellulose Conversion 71.1% in neutral aqueous solution and an intermediate temperature
of 165 C. No deactivation was observed after repeated reactions.
Yield & Conversion (%)

80

60 Acknowledgment

40 This work was supported by the Basic Science Research Program


through the National Research Foundation of Korea (NRF) funded
20 by the Ministry of Education, Science and Technology (2010-
0009174) and the S-Oil company.
0
1st 2nd 3rd 4th 5th
Appendix A. Supplementary data
Recycling number

Fig. 2. Recyclability of dual-functionalized catalysts (Ru/AC-SO3H) up to 5th repeated Supplementary data to this article can be found online at doi:10.
reactions. 1016/j.catcom.2011.12.032.
118 J.W. Han, H. Lee / Catalysis Communications 19 (2012) 115118

References [9] R. Palkovits, K. Tajvidi, J. Procelewska, R. Rinaldi, A. Ruppert, Green Chemistry 12


(2010) 972.
[1] M. Jarvis, Nature 426 (2003) 611. [10] A. Corma, G.W. Huber, Angewandte Chemie International Edition 46 (2007) 7184.
[2] N. Yan, C. Zhao, C. Luo, P.J. Dyson, H.C. Liu, Y. Kou, Journal of the American Chem- [11] R.W. Thring, S.P.R. Katikaneni, N.N. Bakhshi, Fuel Processing Technology 62
ical Society 128 (2006) 8714. (2000) 17.
[3] A. Fukuoka, P.L. Dhepe, Angewandte Chemie International Edition 45 (2006) [12] R.D. Cortright, R.R. Davda, J.A. Dumesic, Nature 418 (2002) 964.
5161. [13] G.W. Huber, J.N. Chheda, C.J. Barrett, J.A. Dumesic, Science 308 (2005) 1446.
[4] C. Luo, S.A. Wang, H.C. Liu, Angewandte Chemie International Edition 46 (2007) [14] B. Kamm, Angewandte Chemie International Edition 46 (2007) 5056.
7636. [15] A. Onda, T. Ochi, K. Yanagisawa, Topics in Catalysis 52 (2009) 801.
[5] W.P. Deng, X.S. Tan, W.H. Fang, Q.H. Zhang, Y. Wang, Catalysis Letters 133 (2009) [16] J.F. Pang, A.Q. Wang, M.Y. Zheng, T. Zhang, Chemical Communications 46 (2010)
167. 6935.
[6] H. Kobayashi, Y. Ito, T. Komanoya, Y. Hosaka, P.L. Dhepe, K. Kasai, K. Hara, A. Fukuoka, [17] A. Onda, T. Ochi, K. Yanagisawa, Green Chemistry 10 (2008) 1033.
Green Chemistry 13 (2011) 326. [18] M. Hara, M. Okamura, A. Takagaki, M. Toda, J.N. Kondo, K. Domen, T. Tatsumi, S. Hayashi,
[7] H. Kobayashi, H. Matsuhashi, T. Komanoya, K. Hara, A. Fukuoka, Chemical Com- Chemistry of Materials 18 (2006) 3039.
munications 47 (2011) 2366.
[8] J. Geboers, S. Van de Vyver, K. Carpentier, K. de Blochouse, P. Jacobs, B. Sels, Chem-
ical Communications 46 (2010) 3577.

Vous aimerez peut-être aussi