Sie sind auf Seite 1von 10

ARTICLE IN PRESS

G Model
MOLCAA-8868; No. of Pages 10

Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Solid base supported metal catalysts for the oxidation and


hydrogenation of sugars
Anup Tathod, Tanushree Kane, E.S. Sanil, Paresh L. Dhepe ∗
Catalysis and Inorganic Chemistry Division, CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune 411008, India

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

Article history: Pt impregnated on ␥-Al2 O3 (acidic support) and hydrotalcite (basic support) catalysts were synthesized,
Received 28 April 2013 characterized and used in the oxidation and hydrogenation reactions of C5 and C6 sugars. In the absence
Received in revised form 21 August 2013 of homogeneous base, 83% yield for gluconic acid; an oxidation product of glucose can be achieved over
Accepted 11 September 2013
Pt/hydrotalcite (HT) catalyst at 50 ◦ C under atmospheric oxygen pressure. Similarly, 57% yield for xylonic
Available online xxx
acid, an oxidation product of xylose is also possible over Pt/HT catalyst. Hydrogenation of glucose con-
ducted using Pt/␥-Al2 O3 + HT catalytic system showed 68% sugar alcohols (sorbitol + mannitol) formation.
Keywords:
The 82% yield for C5 sugar alcohols (xylitol + arabitol) was obtained by subjecting xylose to hydrogena-
Solid base
Supported metal catalysts
tion over Pt/␥-Al2 O3 + HT at 60 ◦ C. UV analysis helped to establish the fact that under alkaline conditions
Oxidation sugars prefer to remain in open chain form in the solution and thus exposes CHO group which further
Hydrogenation undergoes oxidation and hydrogenation reactions to yield acids and alcohols.
Sugars © 2013 Elsevier B.V. All rights reserved.

1. Introduction water soluble cleansing agent, an ingredient for concrete and as an


intermediate in the food and pharmaceutical industries [13]. Cal-
The requirement for alternate resources for obtaining chem- cium gluconate is used as a detoxifying agent after exposure to HF
icals which are presently derived from non-renewable (in short [14]. Gluconic acid has an annual estimated market of 6 × 104 ton
cycle), fossil feedstocks have attracted researchers across the world and on industrial scale is currently prepared by fermentation of glu-
to work on the renewable resource, viz., biomass. Among biomass cose by Aspergillus niger [10,15]. However, fermentation method
non-edible to humans, lignocellulosic materials (cellulose, hemi- has many drawbacks such as slow reaction rate and disposal of
celluloses, and lignin), which can be derived from agricultural dead microbes and excretory substances of microbes [16,17]. To
wastes have dominated the research domain for chemical synthe- overcome these drawbacks, replacement of enzymes with hetero-
sis. In an acidic medium, cellulose undergoes hydrolysis reaction geneous catalysts is a viable option.
to yield glucose which upon further dehydrocyclization gives 5- There have been few reports on the aerobic oxidation of glucose
hydroxymethyl furfural (HMF). Similarly, hemicelluloses (xylans) to gluconic acid carried over supported transition metals catalysts
can yield xylose, arabinose and furfural under the acidic conditions. including Pd, Pt, Rh and Au supported on various supports like TiO2 ,
Once these sugars and furans are obtained, those can act as plat- Al2 O3 and activated carbon (AC) [8,10,11,18–27]. Among those,
form chemicals to synthesize value added chemicals by undergoing gold catalysts have been extensively studied by various research
oxidation (acids), hydrogenation (sugar alcohols), isomerization, groups [10,18,19,23–27]. It is claimed that gold nanoparticles (NP)
hydrogenolysis (glycols), and reforming (alkanes, hydrogen) reac- supported on activated carbon show good catalytic activity in the
tions. oxidation of glucose to gluconic acid done at 50 ◦ C [22]. Another
Glucose, derived from cellulose is used as a food additive report claims better turnover frequency (TOF) values per surface
but it can also be converted to many useful chemicals such as Au atom of 45 s−1 at 50 ◦ C, pH 9.0 and 56 s−1 at 50 ◦ C, pH 9.5 in
ethanol (fermentation) [1], sorbitol and mannitol (hydrogenation) the case of gold catalyst supported on ZrO2 [28]. In most of these
[2,3], 5-hydroxymethyl furfural (dehydrocyclization) [4], hydro- studies metals suffer from over oxidation and thus they eventually
gen (hydrogenolysis/cracking/reforming) [5–7] and gluconic and deactivate [22–24,27]. Study on particle size effect shows that rate
glucaric acid (oxidation) [8–12]. Gluconic acid has many diverse of reaction is inversely proportional to Au size indicating that reac-
industrial applications such as, biodegradable chelating agent, tion is structure-sensitive. In Au/C catalyst, Au with 1.9 nm particle
size gives maximum yield [29]. It is also observed that these metals
undergo sintering during the reaction and thus drop in activity is
∗ Corresponding author. Tel.: +91 20 25902024; fax: +91 20 25902633. reported [30]. Moreover, in all these reactions, homogeneous base
E-mail address: pl.dhepe@ncl.res.in (P.L. Dhepe). (1–20 equiv.) is used to maintain the reaction solution pH between

1381-1169/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.molcata.2013.09.014

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

2 A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx

disadvantages associated with nickel based catalysts by synthe-


sizing Fe, Mo, Sn, Cr promoted nickel, and nickel supported on
ZrO2 , TiO2 and ZrO2 /TiO2 catalysts [3,43]. Hydrogenation of glu-
cose over supported metal catalysts has been extensively studied
by the researchers in the last few decades [2,46–50]. Ru cata-
lysts are expansively used for the glucose hydrogenation, but the
catalysts are observed to be deactivating during the continuous
use [51]. Typically, neutral (AC, SiO2 ) [2,46,47] or mildly acidic
(Al2 O3 ) [3] supports have been used in the study. Enzymatic con-
version of glucose is another route for the production of sorbitol
but enzymes, soluble in water are difficult to separate after use and
hence intricate to recycle [52]. Xylose hydrogenation using sup-
ported metal catalysts and enzymes is well known [39,53–58]. It is
shown that hydrogenation of xylose is possible over Ru/C catalyst
[59]. Recently, it is reported that the reduced Ni/Cu/Al hydrotalcite
catalyst at 125 ◦ C and 30 bar H2 pressure can convert glucose to
yield ca.70% sorbitol [60].
In the current work we have studied glucose and xylose hydro-
genation using Pt supported on acidic, ␥-Al2 O3 and basic, HT
supports under mild reaction conditions (Scheme 1). Good yield
and selectivity has been achieved at optimum reaction conditions.
Even though solid base catalysts have been used in several
industries, extensive catalytic studies on them are scanty compared
Scheme 1. Conversion of sugars to corresponding acids and sugar alcohols. to the solid acid catalysts [61]. Application of homogeneous bases
in reactions such as isomerization, addition, alkylation and cycliza-
tion is well documented. But the homogeneous nature makes their
9 and 11. The increase in pH (>11) typically leads to degradation separation and recycling very tricky. Beside this, use of homoge-
reactions and hence it is required to charge the homogeneous base neous bases makes the overall process corrosive and most of the
in continuous mode to maintain the pH [22]. Additionally, handling times it needs to be added in the stoichiometric amount. On the
of corrosive homogeneous base is also a major problem. It is known other hand solid bases are easily separable from reaction mix-
that glucose oxidation carried out in a media other than alkaline is ture, safe to handle and can be recycled. In recent past, solid base
a very slow process. Under acidic conditions, AC supported Au–Pt catalysts such as, basic zeolites, hydrotalcites, apatites, chrysotile,
bimetallic catalysts show 64% conversion and turnover frequency sepiolite, basic metal oxides like MgO, CaO and mixed oxide like
(TOF) of 295 h−1 [8]. Most of the reports use high pressures of oxy- SiO2 -MgO, SiO2 -CaO, Al2 O3 -MgO, etc. have been used instead of
gen in the reactions. There are only few reports available in the homogeneous bases [61–67]. It is understood that in solid base cat-
literature on the oxidation of xylose to xylonic acid using heteroge- alysts like metal oxides, surface O2− species form the Lewis base site
neous catalysts [31]. Electro-catalytic oxidation of xylose to xylonic [61,64].
acid using Au and Pt electrodes in alkaline medium is known [32]. Hydrotalcites (HT) are anionic clays having a double layered
To avoid the use of homogeneous base and to develop a method structure with OH− and HCO3 − groups. They have the gen-
which is environmentally friendly and harmless, we have employed eral formula, [Me2+ 1−x Me3+ x (OH)2 ]x+ (An− )x/n ·mH2 O and naturally
Pt-supported on solid acid and basic supports. In this paper, we occurring hydrotalcite is hydroxycarbonate of magnesium and alu-
report the liquid phase aerobic oxidation of glucose to gluconic minum having formula Mg6 Al2 (OH)16 CO3 ·4H2 O. They are known
acid (Scheme 1) using Pt/hydrotalcites (HT). To compare the results to show catalytic activity in numerous base catalyzed reactions
we have also carried out reactions with Pt/␥-Al2 O3 catalysts with such as, aldol condensation, Knoevenagel and transesterification
the addition of homogeneous base (Na2 CO3 ) and solid base (HT). [68–70]. HT supported metal catalysts are applied in the oxidation
Oxidation of xylose to xylonic acid is also performed. reactions of alcohols [71]. This class of catalysts shows applications
In the US DOE report, sugar alcohols, sorbitol and xylitol in chemical synthesis, in isomerization reactions and as additive
obtained by hydrogenation of glucose and xylose, respectively are in polymer and in medicine [72–74]. Layered double hydroxides of
named among the 12 value added chemicals obtained from biomass Mg and Fe have been used in catalysis; water purification and anion
[33]. These sugar alcohols are low calorie sweeteners and find exchange [75–77]. Basic nature of hydrotalcites has been studied
application in oral hygiene products. Sorbitol is used as humectant extensively to use them in few other reactions [78,79].
and utilization of xylitol is known in cosmetics and pharmaceu- In this work, our aim of using basic support like HT is to shun
tical industries. Moreover, sugar alcohols have non-diabetes and the use of homogeneous base in oxidation reactions and to make
anti-caries properties because of low amount of lactic acid forma- the overall process environmentally benign and safe. It is thought
tion [34,35]. Recently, exploitation of sorbitol for the production that properties pertaining to strong interaction between metal and
of hydrogen and C5, C6 hydrocarbons has been investigated [6,36]. basic support may help us in achieving higher activities in oxidation
Additionally, a spectrum of other functional chemicals such as, 1,4- and hydrogenation reactions.
sorbitan, sorbose, isosorbide, glycols, lactic acid, vitamin C etc. are
produced from sugar alcohols [37,38]. Due to widespread applica-
tions of sugar alcohols demand for those is increasing day by day 2. Experimental
[39,40].
Sorbitol (annual production: 6.5 × 105 ton) and xylitol (annual 2.1. Materials
production: 2.4 × 104 ton) are commercially prepared by hydro-
genation of glucose and xylose using (Raney) nickel catalyst Magnesium nitrate hexahydrate (Mg(NO3 )2 ·6H2 O) was pur-
[41,42]. However, catalyst deactivation is a major problem in chased from Merck, India (99%), Aluminum nitrate nonahydrate
these reactions [43–45]. Efforts have been taken to rise above the (Al(NO3 )3 ·9H2 O) was purchased from Thomas Baker, Germany.

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx 3

Tetraamine platinum nitrate (99.99%) was purchased from Alfa 2.5. Catalytic reactions and analysis
Aesar, UK. Sodium carbonate (Na2 CO3 ) (99.5%), sodium hydroxide
(NaOH) (98%) and xylose were purchased from Loba Chemicals, 2.5.1. Glucose and xylose oxidation reactions and analysis
India. Sorbitol and levulinic acid were purchased from Aldrich Glucose and xylose oxidation reactions were carried out in a
Chemicals, USA. Glucose, arabitol, arabinose, xylitol, mannitol, batch mode reactor (Parr autoclave, USA) with a teflon beaker. The
gluconic acid, glycerol, ethylene glycol, 1,2-propanediol were pur- reactor was charged with 3.9 mmol of glucose or 4.7 mmol of xylose
chased from S.D. Fine Chemicals Limited, India. All the chemicals in 30 mL water. Subsequently supported metal catalyst was added.
were used as received. While reactions were conducted using Pt/␥-Al2 O3 catalyst, either
0.2 g of Na2 CO3 or 0.3 g HT was added in the reaction mixture. In
this work during the course of reaction (intermediate addition) to
2.2. Synthesis of hydrotalcite maintain the pH of the solution, no base was added. The reactor
was flushed 3 times with oxygen and then environment of oxygen
Hydrotalcite was synthesized by co-precipitation method. was maintained (atmospheric oxygen pressure). The autoclave was
NaOH was used to maintain the pH, while Na2 CO3 was used as a then heated to desired temperature under slow stirring (100 rpm).
carbonate source. For the synthesis of catalyst by co-precipitation After attaining the reaction temperature, stirring was increased to
method, an aqueous solution ‘A’ (37.5 mL) of Mg(NO3 )·6H2 O 700 rpm Reaction mixture was analyzed using HPLC (Shimadzu,
(0.279 mol) and Al(NO3 )3 ·9H2 O (0.093 mol) were slowly added into Japan) equipped with HC-75 H+ (9 ␮m, 7.8 mm × 305 mm) column
aqueous solution ‘B’ (37.5 mL) of NaOH (0.4375 mol) and Na2 CO3 while succinic acid (0.5 mmol) is used as an eluent (0.5 mL/min).
(0.1125 mol) under vigorous stirring over a period of 2 h. Solution The refractive index detector (model no. RID-6A) was used for the
pH was maintained between 8 and 10. The white precipitate thus detection and calibration of compounds. The calibration of com-
obtained was kept in autoclave at 60 ◦ C for 16 h. Precipitate was pounds was done using commercially available standards.
then washed with deionised water until pH becomes neutral. Solid
thus obtained was then dried at 80 ◦ C for 18 h. Sample was calcined
2.5.2. Glucose and xylose hydrogenation reactions and analysis
at 550 ◦ C in the presence of air for 8 h. Mg/Al ratio in hydrotalcite
Glucose and xylose hydrogenation reactions were carried out in
obtained was 3.
a batch reactor (Amar equipments, India). The reactor was charged
with 0.83 mmol of glucose or 1.00 mmol of xylose in 35 mL water.
2.3. Preparation of supported platinum catalysts During reactions catalyst/substrate ratio was kept constant at 1:2
(wt/wt). When reactions were conducted using HT, 0.075 g of HT
Before the synthesis of supported metal catalysts, supports were was added. Experiments were conducted at temperatures ranging
evacuated at 150 ◦ C for 6 h. An aqueous solution of platinum precur- between 60 ◦ C and 190 ◦ C under 8–24 bar of hydrogen pressure.
sor, tetrammine platinum nitrate (quantity used corresponding to The reactions were conducted for 2–6 h. Reaction mixture was ana-
3.5 wt% loading on support) was added to the support suspended in lyzed using HPLC (Agilent, USA) equipped with HC-75 Pb++ (9 ␮m,
water. The solution was stirred for 16 h at room temperature. Next, 7.8 mm × 300 mm) column. Water was used as an eluent, at the flow
water was removed by rotary evaporator and powder obtained was rate of 0.6 mL/min, and RID (model no. 1260 infinity) was used to
dried in oven at 60 ◦ C for 16 h. The powder was later dried under detect the compounds. The products which are poorly separable by
vacuum at 150 ◦ C for 6 h. This dry powder was subjected for calci- using HC-75 Pb++ column, were analyzed using HPLC (Shimadzu,
nation in air flow (20 mL/min) and was reduced in hydrogen flow Japan) equipped with HC-75 H+ (9 ␮m, 7.8 mm × 305 mm) col-
(20 mL/min) at 400 ◦ C for 2 h each. umn. Succinic acid (0.5 mmol) is used as an eluent (0.5 mL/min).
The refractive index detector (Model no. RID-6A) was used for the
detection and calibration of compounds. The calibration of all the
2.4. Characterization of catalysts compounds was done using commercially available standards.
Calculations of selectivity and yield (%),
The powder X-ray diffraction (XRD) patterns of all samples were  mole of gluconic acid formed (HPLC) 
recorded on a Philips Powder XRD, operated at Cu K␣ radiation Selectivity (%) = × 100
( = 1.540598 Å) as a primary beam. The patterns were recorded mole of glucose converted (HPLC)
with a scanning rate of 2◦ /min. Nitrogen sorption analysis at
low temperature was performed using Dona Quantachrome Nova  mole of product formed (HPLC) 
4200e instrument. The surface areas of catalysts were calculated by Yield (%) = × 100
theoretical mole of product
the BET method. The temperature programmed desorption (TPD) of
CO2 was done using the Micromeritics AutoChem 2910 instrument.
Prior to measurement, samples were activated at 350 ◦ C for 15 min 3. Results and discussion
The adsorption of CO2 was done at 50 ◦ C and then the sample was
kept at 100 ◦ C under He flow. Next, sample was heated from 100 ◦ C 3.1. Catalysis
to 600 ◦ C with a ramping rate of 4 ◦ C/min. The TPD-NH3 analysis
was done using Micrometrics Autochem-2910 model, instrument; 3.1.1. Results
USA. Catalyst was activated at a temperature of 500 ◦ C in He gas 3.1.1.1. Oxidation reactions. Oxidation of glucose was carried out
flow (25 mL/min). NH3 adsorption (30 mL/min) was done at 100 ◦ C over Pt supported on ␥-Al2 O3 and HT catalysts in water medium
and desorption was started from 100 ◦ C to 823 ◦ C at the rate of at 50 ◦ C under atmospheric oxygen pressure (Table 1). The Pt/HT
10 ◦ C/min. Metal contents in the prepared catalysts were ana- and Pt/␥-Al2 O3 + Na2 CO3 catalysts gave similar oxidation results
lyzed by inductive coupled plasma atomic emission spectroscopy (entry 1 and 2). Pt/HT catalytic system without addition of any
(ICP-AES) using SPECTRO ARCOS Germany, FHS 12 instrument. homogeneous base gave good performance for the oxidation of
Transmission electron microscopy (TEM) was done using FEI Tec- glucose showing 83% gluconic acid yield after 12 h reaction time
nai TF-30 instrument. UV absorbance spectras were recorded using (entry 1). Along with gluconic acid ca. 8% fructose formation was
PerkinElmer spectrophotometer (model-Lambda 650). Tempera- also observed. Whereas in conventional system in which homoge-
ture programmed reduction (TPR) was done using Micromeritics neous base, Na2 CO3 was used along with the Pt/␥-Al2 O3 catalyst
AutoChem chemisorptions analyzer. 82% gluconic acid yield was observed (entry 2). In this reaction ca.

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

4 A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx

Table 1
Comparison of gluconic acid yield with different catalyst at various time intervals.

Entry Catalyst Gluconic acid yield (%) TOFa pH


(Glucose conversion (%))

6h 8h 10 h 12 h Before reaction After reaction

1 Pt/HT 73 76 81 (95) 83 (99) 0.252 9.31 7.96


2 Pt/␥-Al2 O3 + Na2 CO3 65 82 (98) 83 (99) – 5.278 11.06 9.36
3 Pt/␥-Al2 O3 + HT 55 (94) 56 (99) 55 54 5.485 8.76 7.02
4 HT Isomerization 9.24 9.16
5 ␥-Al2 O3 + Na2 CO3 Isomerization 10.95 10.84
6 Non catalytic – – – –

Reaction conditions: glucose 0.7 g, catalyst 0.3 g, Na2 CO3 0.2 g, HT 0.3 g, water 30 mL, 1 bar O2 at R.T., 50 ◦ C. During the course of reaction pH was not maintained.
a
Turn over frequency in min−1 . TOF is calculated at initial stage of reaction at almost similar conversion levels (±5% difference) in all the reactions.

5% fructose formation also was observed. Careful inspection reveals catalyst in the presence of homogeneous base, showed that increase
that actually after 6 h, higher gluconic acid yield was observed with in oxygen partial pressure gives poor FDCA yields [83]. This might
Pt/HT catalyst (73%) than Pt/␥-Al2 O3 + Na2 CO3 catalyst (65%). The be due to the presence of higher concentration of oxygen in water
Pt/␥-Al2 O3 + HT catalytic system afforded maximum gluconic acid at elevated pressures which causes oxidation of active metal [84].
yield of 56% after 8 h of reaction. In all these experiments, almost In this work we did reactions over Pt/HT with varying pressure
complete conversion of glucose was achieved. It is important to (1 bar, 5 bar) and we achieved marginally less gluconic acid yield at
note here that in all the reactions besides desired products, we also 5 bar (51%, 24 h). The catalysts could be recycled effectively at least
observed formation of oxalic acid, malonic acid, lactic acid etc. in for 3 times with a loss of ca. 10% activity after each catalytic run.
minor amount (4–5%). Blank run (non catalytic, entry 6) did not This loss of activity might be due to the leaching of Pt in the solu-
show any activity under the reaction conditions. Bare HT could tion which was confirmed by ICP-OES analysis (please refer catalyst
not convert glucose to oxidation products; however, 20% yield of characterization section).
fructose was observed (entry 4), which is in line with the liter-
ature wherein it is reported that basic catalysts are capable of 3.1.1.2. Hydrogenation reactions. Hydrogenation of glucose was
isomerizing glucose [80–82]. Similarly, reactions done using only studied over Pt/␥-Al2 O3 , Pt/HT and Pt/␥-Al2 O3 + HT catalysts at
␥-Al2 O3 + Na2 CO3 (entry 5) did not yield any oxidation product 90 ◦ C using 16 bar hydrogen pressure. Amongst all the catalytic
but formation of fructose was observed (19%). These results clearly systems, Pt/␥-Al2 O3 + HT catalytic system gave the best result
indicate that not only basicity but metal also is necessary to obtain (Table 3). As observed, Pt/␥-Al2 O3 + HT showed 68% sugar alco-
oxidation products. hols yield with 92% conversion (entry 4). Under similar reaction
To check the stability of gluconic acid under the reaction con- conditions, Pt/HT (entry 3) showed 39% sugar alcohol yield (80%
ditions, catalytic runs with gluconic acid as substrate over HT, conversion) and the lowest activity was observed for Pt/␥-Al2 O3
␥-Al2 O3 , Pt/HT, Pt/␥-Al2 O3 + Na2 CO3 and Pt/␥-Al2 O3 + HT were car- catalyst (8.4% yield, 15% conversion, entry 2). In the absence of
ried out and we did not observe any conversions. This clearly any catalyst only 2% conversion with 0.5% sugar alcohol yields is
implies that gluconic acid is stable and does not participate in any observed (entry 1). Besides C6-sugar alcohols, formation of C5-
further reactions or undergoes any degradation reactions under our sugar alcohol (xylitol), ethylene glycol, 1,2-propanediol, glycerol,
reaction conditions. gluconic acid and levulinic acid is observed. The results clearly indi-
Table 2 summarizes the results for xylose oxidation reaction. cate that Pt/␥-Al2 O3 + HT catalytic system gives maximum yield
As seen, maximum yield of 65% was attainable with Pt/␥- (68%) of C6-sugar alcohols compared to any other catalysts.
Al2 O3 + Na2 CO3 catalytic system (entry 2). On the other hand, The results imply that change in support affects the activity of
Pt/␥-Al2 O3 + HT (entry 3) showed maximum yield of 53% and Pt/HT catalyst and hence reactions with only supports were carried out. It
catalyst showed 57% yield (entry 1). Xylose oxidation carried out is interesting to note here that without catalyst († ESI Table S1, entry
in the absence of catalyst did not show any conversions (entry 1) and with ␥-Al2 O3 († ESI Table S1, entry 2) as a catalyst maximum
6). Along with oxidation products, arabinose, lactic acid, tartaric of 4% conversion is possible after 6 h. However, when basic cata-
acid formation and some other unidentified products were also lyst is introduced the conversion increased to 80% († ESI Table S1,
observed. Formation of the minor products (ca. 5%) arises due to entry 3) yielding 39% fructose and 16% sugar alcohols. As explained
isomerization reaction of xylose when reactions were done using earlier (Section 3.1.1.1), isomerization of glucose is possible under
bare HT and ␥-Al2 O3 + Na2 CO3 catalytic systems at 50 ◦ C. the alkaline medium. It is also evidenced that the HT in hydroxide
Our earlier work on the oxidation of 5-hydroxymethyl fur- form has the most basic properties which may help achieve higher
fural to yield 2,5-furandicarboxylic acid (FDCA) using Pt/␥-Al2 O3 isomerization activity [85]. Reaction with ␥-Al2 O3 + HT (entry 4)

Table 2
Comparison of xylonic acid yield with different catalysts at various time intervals.

Entry Catalyst Xylonic acid yield (%) TOFa pH


(Xylose conversion (%))

6h 12 h 18 h 24 h Before reaction After reaction

1 Pt/HT 24 – 52 57 (90) 0.124 9.28 9.01


2 Pt/␥-Al2 O3 + Na2 CO3 27 30 54 65 (84) 1.119 11.02 9.87
3 Pt/␥-Al2 O3 + HT 18 20 21 53 (86) 0.726 8.62 8.09
4 HT Isomerization 9.24 9.36
5 ␥-Al2 O3 + Na2 CO3 Isomerization 10.93 10.80
6 Non catalytic – – – –

Reaction conditions: xylose 0.7 g, catalyst 0.3 g, Na2 CO3 0.2 g, HT 0.3 g, water 30 mL, 1 bar O2 at R.T., 50 ◦ C. During the course of reaction pH was not maintained.
a
Turn over frequency in min−1 . TOF is calculated at initial stage of reaction at almost similar conversion levels (±5% difference) in all the reactions.

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx 5

Table 3
Effect of different catalysts on the hydrogenation of glucose.

Entry Catalyst Time (h) Conversion (%) Product yield (%) TOFa pH

Sugar alcohols Fructose Glycols Gluconic Before After


acid reaction reaction

Sorbitol Mannitol Xylitol

1 Without catalyst 2 – – – – – – – – – –
4 2 0.5 – – – –
6 4 1.5 0.5 – – –

2 Pt/␥-Al2 O3 2 8 5.7 – – – – 0.91 5.51 5.69


4 15 8.4 – – – 1.5 –
6 25 9.8 – 1 – 3.5 –

3 Pt/HT 2 60 12 13.3 2 15 3.7 5 0.98 9.29 9.64


4 80 19.4 19.6 4.5 16 6.5 5.5
6 85 26 19.5 5 14 9 5.5

4 Pt/␥-Al2 O3 +HT 2 80 42 16 4 3 6.5 4 10.21 8.42 9.19


4 92 54 14 4 – 7 5
6 98 53 12 4 – 8 5

Reaction conditions: glucose 0.15 g, catalyst 0.075 g, HT 0.075 g, water 35 mL, 16 bar H2 at R.T., 90 ◦ C.
a
Turn over frequency in min−1 . TOF is calculated at initial stage of reaction at almost similar conversion levels (±5% difference) in all the reactions.

also showed higher glucose conversions (72%) with 42% fructose in 2 h of reaction time and 92% conversion with 68% yield of C6-
and 18.6% sugar alcohol formation. Contrary to the recent report sugar alcohols could be obtained in 4 h (Table 3, entry 4). From the
wherein it is shown that alkaline medium (NaOH) yields less sor- figure it is clear that we could achieve very high carbon balance
bitol compared to acidic or neutral medium [60], we observed that since there is typically only 5–7% difference between conversion
better yields are possible in alkaline medium (solid base). and total product yields.
Fig. 1 depicts the effect of temperature on the catalytic activity It is known from the literature that hydrogen solubility follows
over Pt/␥-Al2 O3 + HT catalyst. As seen, increase in temperature to Henry’s law which means that solubility of hydrogen in water is
130 ◦ C does not necessarily increase the sugar alcohols yield even directly proportional to pressure and temperature [84]. To under-
though conversion was increased to 99% in 2 h. This in fact lowers stand how these factors would play role in our catalytic system we
the yield for C6-sugar alcohols formation (56%) at the expense of altered the pressures. The increase in pressure from 16 to 24 bar did
increase in the yield for glycols (1,2-propane diol, glycerol, ethyl- not alter the catalytic results (16 bar: 68% yield with 92% conver-
ene glycol) and other products like levulinic acid and xylitol (39%). sion; 24 bar: 67% yield with 93% conversion, 4 h). However, if 8 bar
On the other hand decrease in temperature to 60 ◦ C reduced the hydrogen pressure is used then activity was decreased (54% yield
conversion levels to 28% with 21% yield of C6-sugar alcohols. In with 93% conversion, 8 h).
view of above results, 90 ◦ C reaction temperature is optimum to Hydrogenation of xylose is studied over catalysts, Pt/␥-Al2 O3 ,
obtain higher conversions (80%) and 58% C6-sugar alcohols yield Pt/HT and Pt/␥-Al2 O3 + HT at 60 ◦ C with 16 bar hydrogen pressure
(Table 4). In these reactions also Pt/␥-Al2 O3 + HT catalytic system
showed the best activity (82% yield with 99% conversion) among
all the catalysts evaluated. It can be seen that highest TOF was
observed for Pt/␥-Al2 O3 + HT system.
The effect of temperature was studied and it was found out that
60 and 90 ◦ C reaction temperatures are the best (Fig. 2). Although
At 60 ◦ C conversion is less (82%) than other temperatures, the selec-
tivity for sugar alcohols is highest, i.e. 87.8%. Maximum 82% yield
of xylitol + arabitol could be obtained from xylose at 60 ◦ C within
4 h.
Effect of supports on the yields of sugar alcohols has been stud-
ied by conducting reactions with bare supports († ESI Table S2).
Similar to glucose hydrogenation, xylose hydrogenation has
been studied by applying various H2 pressures. Yield of sugar alco-
hols obtained after complete conversion, at any pressure, seems
to be similar (ca. 80%). But the noteworthy point is that the time
required for total conversion of xylose is different. Xylose conver-
sion was 95% at 24 bar (80% yield of sugar alcohols), 82% at 16 bar
(72% yield) and 63% at 8 bar (58% yield) after 2 h. Complete conver-
sion with 82% yield occurs after 4 h at higher pressure but it takes
6 h for complete conversion at 8 bar
Recycle study has been done for Pt/␥-Al2 O3 + HT catalytic sys-
tem for glucose and xylose hydrogenation. In case of glucose
hydrogenation, activity decreases by 20% after 1st cycle and 12%
after 2nd cycle. However, in xylose hydrogenation, catalyst shows
good recyclability up to 3rd cycle with negligible decrease in the
Fig. 1. Results for the conversion of glucose to sugar alcohols at various reaction
temperatures using Pt/␥-Al2 O3 + HT. Reaction conditions: glucose 0.15 g, Pt/␥-Al2 O3 yield after each run (1st run: 82%; 2nd run: 80%; 3rd run: 74%).
0.075 g, HT 0.075 g, water 35 mL, 16 bar H2 at R.T., 2 h. These results can be correlated to the fact that during reactions Pt

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

6 A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx

Table 4
Effect of different catalysts on the hydrogenation of xylose.

Entry Catalyst Time (h) Conversion (%) Product yield (%) TOFa pH

Sugar alcohols Glycols Xylonic Before After


acid reaction reaction

Xylitol Arabitol

1 Without catalyst 2 – – – – – – – –
4 – – – – – – – –
6 2 0.5 – – – – – –

2 Pt/␥-Al2 O3 2 4 3.3 – – – 0.38 5.31 5.47


4 6 3 – 1 – – –
6 9 4.5 – 1.5 –

3 Pt/HT 2 41.5 36 6 – – 0.42 9.20 9.57


4 59 50 6 1.5 – – –
6 72 55 4 3.5 2 – –

4 Pt/␥-Al2 O3 + HT 2 82 68 4 4 2 13.01 8.31 9.29


4 99 79 3 6 2.5 – –
6 99 78 3 6 2.5 – –

Reaction conditions: xylose 0.15 g, catalyst 0.075 g, HT 0.075 g, water 35 mL, 16 bar H2 at R.T., 60 ◦ C.
a
Turn over frequency in min−1 . TOF is calculated at initial stage of reaction at almost similar conversion levels (±5% difference) in all the reactions.

is not leached out in the solution (refer catalyst characterization adsorption of sorbitol on HT and ␥-Al2 O3 . The results presented in
section). Table S3 († ESI) however shows that there is no considerable differ-
ence in the adsorption of sorbitol on either HT or ␥-Al2 O3 . This may
3.1.2. Discussion imply that to obtained enhanced activity over Pt/HT catalyst com-
It is well documented that solid base can suppress the crack- pared to Pt/␥-Al2 O3 catalyst weak or strong adsorption of product
ing reaction, which generally occur on acidic catalysts (␥-Al2 O3 ) may not be the reason.
[86]. It is suggested that over solid bases no strong adsorption of It is known from the literature that at the surface of a basic sup-
complexes (through ‘O’ of the product) is possible which means port typically, oxide ions form ion pairs with low coordinated metal
that the products formed will be readily removed from the sur- ions [87]. Usually, hydrogen molecule is split heterolytically over
face to avoid undergoing further reactions (glycols, coke). This can supported metal (Pt) to give rise to hydrogen ions (H+ and H− ).
improve the yield for sugar alcohols and acids derived from C5 In case of base support, there is a possibility of retention of the
and C6 sugars (glucose and xylose) by suppressing further reac- molecular identity of H atoms during a reaction as is reported ear-
tions of acids/alcohols and formation of coke. This may indicate the lier [87]. It is possible that hydrogen ions are immobilized on the
reason for the better results on Pt/HT compared with Pt/␥-Al2 O3 set of active sites (O2− – metal cation pair) and that these sites are
catalysts. To prove this we conducted the experiments to check the separated from each other and hence migration and recombination
of hydrogen ions to form hydrogen molecule is not possible. This
ultimately may give rise to rich environment of hydrogen ions on
the support. This pool of hydrogen ions then may be available for
the hydrogenation reaction thus increasing the activity of a catalyst.
The other possibility for increased activity in case of Pt sup-
ported on basic support is that the basic support can give rise
to electron-rich metallic particles. In fact it is reported that elec-
tron rich Pd particles are possibly formed when MgO is used as
a support [88]. A similar phenomenon is also reported for Pt par-
ticles supported on hydrotalcites [89,90]. In yet another report it
is evidenced that since Pt particle size is almost same in all the
tested catalysts, it is the electronic ligand effect produced by basic
support that enhances the hydrogenation activity [91]. The strong
metal-support interaction (SMSI) is first reported for Pt/MgO cat-
alyst and it was showed that higher metal dispersion achieved by
SMSI enhances the hydrogenation activity [92]. In case of our cat-
alysts, better hydrogenation activity with Pt/HT may arise due to
SMSI in Pt/HT compared to Pt/␥-Al2 O3 catalyst as we observed that
over HT support Pt is well dispersed to give rise to average parti-
cle size of 1.9 nm while in case of Pt supported on ␥-Al2 O3 average
particle size of Pt was 20–30 nm (TEM study, refer Section 3.2). In
oxidation reactions metal in zero oxidation state is most active [93]
and as discussed above when Pt is present on basic support (HT,
MgO) it may have more electron rich environment. This may be the
reason for higher oxidation activity for Pt/HT catalyst. Moreover, Pt
particle size decreases if there is a strong metal support interaction
and thus in Pt/HT catalyst particle size of Pt was observed to be
Fig. 2. Results for the conversion of xylose to sugar alcohols, at various reaction
temperatures using Pt/␥-Al2 O3 + HT. Reaction conditions: xylose 0.15 g, Pt/␥-Al2 O3 much less than Pt supported on ␥-Al2 O3 . This again will improve
0.075 g, HT 0.075 g, water 35 mL, 16 bar H2 at R.T., 2 h. the oxidation activity on Pt/HT catalyst. An attempt was made to

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx 7

Fig. 3. Proposed changes in the structures of glucose under alkaline conditions.

carefully look at the Pt(4f) core level using XPS as any electronic difference in activity/selectictivity for Pt/␥-Al2 O3 , Pt/HT and Pt/␥-
perturbation associated with Pt can result in changes in the binding Al2 O3 + HT catalysts. Careful observation of the data from Table 3
energy values Figure S1 († ESI). Unfortunately, this was not success- reveals that when Pt/␥-Al2 O3 is used (entry 2) no fructose forma-
ful as Pt (4f) binding energy values (71 eV) falls almost at the same tion is observed. However, when Pt/HT catalyst is used (entry 3),
energy for Al (2p) (73 eV) making it extremely difficult to get any fructose (isomerization product of substrate, glucose) is obtained in
meaningful binding energy shifts. However, we believe that from high quantity (14%, 6 h). In comparison over Pt/␥-Al2 O3 + HT cata-
earlier reports Pt might present in more electron rich state on HT lyst (entry 4, 3%, 2 h) fructose formation was seen. As shown in Table
[89–92]. Besides this we performed TPR studies to know the hydro- S1 († ESI) when bare HT is used for the reaction (entry 3) almost
gen consumption by both Pt/HT and Pt/␥-Al2 O3 catalyst. The data is 39% fructose formation was possible. Compared to this if bare ␥-
plotted in Figure S2 († ESI). It is evident from the figure that in case Al2 O3 is used in the reaction no fructose formation was seen († ESI
of Pt/HT catalyst higher amount of hydrogen (22.9 cm3 /g) is con- Table S1, entry 2). Considering this, it is obvious to expect that if a
sumed as compared to Pt/␥-Al2 O3 catalyst (18.8 cm3 /g). This may basic support is used to impregnate Pt then higher concentrations
imply that Pt present on HT is in more electron rich environment; of mannitol should form in the reaction (via fructose formation
however, it is essential to prove this point beyond doubt. and its further hydrogenation). Now, again looking at sugar alco-
There is a possibility of catalytic transfer hydrogenation of alde- hols formation data from Table 3, with Pt/HT catalyst sorbitol is
hydes while solid base is used as a support (catalyst). However, this formed with 26% yield while mannitol (a hydrogenation product of
possibility is ruled out in our study as with only HT as a catalyst (in fructose) has 19.5% yield (entry 3). Whereas with Pt/␥-Al2 O3 + HT
absence of Pt) minor sugar alcohols formation was observed († ESI system, 53% sorbitol yield is obtained with only 12% mannitol yield.
Tables S1 and S2). Moreover, when Pt/␥-Al2 O3 catalyst is used (without addition of
The best hydrogenation activity for Pt/␥-Al2 O3 + HT catalytic any base) formation of mannitol was not observed. These results
system can be attributed to the fact that presence of HT may influ- indicate that higher the fructose formation, higher the mannitol
ence the isomerization of glucose to make it available in open chain yield and lower is the sorbitol yield (Table 3, entry 3 and 4). We
form. It is proved that concentration of open-chain glucose in water would like to bring to notice that pH of the reaction solution in case
at equilibrium is very small. Dissolving glucose in alkaline medium of Pt/HT catalyst was observed to be 9.3 while with Pt/␥-Al2 O3 + HT
yields fructose by undergoing keto-enol tautomerizations reactions catalyst it was 8.4 and with Pt/␥-Al2 O3 it was 5.5 (Table 3). This dif-
as known by Lobry de Bruyn-Albeda van Ekenstein transformation. ference in pH makes the variation in products yield, since higher
The enolate ion thus formed as an intermediate is present in an pH enhances the concentration of glucose in the enolate form (open
open-chain form (in water) and this higher concentration of ion is chain) which ultimately leads to fructose formation.
available for hydrogenation reactions. It is not possible to obtain The above explanation about presence of glucose in higher con-
similar concentration of glucose in an open chain form under neu- centration in open chain form under alkaline conditions also holds
tral and acidic reaction conditions and hence it is suggested that true for oxidation reactions. In oxidation reactions, gluconic acid
higher yields are possible when Pt/␥-Al2 O3 + HT catalyst is used can be formed in high yields if higher concentration of glucose in
for the reaction (Table 3). It is interesting to point out here the open chain form is available in solution. Moreover in Pt/HT catalytic

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

8 A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx

1.0 out additional studies on using NaOH instead of Na2 CO3 as a base.
With the addition of NaOH with same mole as that like Na2 CO3 ,
a-Glucose+Na 2CO3 we observed additional peak appearing at max = 310.5 nm, which
0.8 267.8 b-Glucose+NaOH can be assigned to enediol anion as reported earlier [94]. Along with
a c-Glucose this peak we also observed peak at max = 270 nm ascribed to CHO
group. To understand why after addition of NaOH additional peak
Absorbance (%)

0.6
appears, we checked the pH of the solution and found out that with
the addition of NaOH, pH was 13.84 while after addition of Na2 CO3 ,
0.4 pH of the solution was 11.48. It is hence understood that with high
270 b pH values (due to NaOH) glucose is converted from cyclic form to
open chain form to yield, enolate ion and subsequently gives rise
0.2 to enediol formation. We could not perform the experiments with
266.2 c the addition of HT since it is a solid base and after adding the same
it is difficult to do UV analysis.
0.0

3.2. Characterizations
225 250 275 300 325 350 375 400
Wavelenth (nm) Fig. 5 shows the XRD patterns of fresh and spent catalysts used in
hydrogenation reactions. It can be observed that most of the peaks
Fig. 4. UV absorption spectra of glucose.
are retained in the spent catalyst. However, careful observation
suggests that in Pt/HT catalyst Pt peaks are very small indicating
system, reaction solution pH is about 9.3 (Table 1) which means that Pt particle size may be very small and Pt might be well dis-
that it falls in the exact range of pH 8–10 at which maximum oxi- persed on the support. The peaks appeared at 35.23◦ , 43.34◦ and
dation activity is reported to be possible [28]. At the same time 62.63◦ correspond to the lattice plane (1 1 1), (2 0 0), and (2 2 0) are
Pt/␥-Al2 O3 + HT system shows pH of 8.7 (Table 1) which is on the characteristic peak for the hydrotalcite (JCPDF file No. 45-0946).
lower marginal side of best pH range to obtain higher oxidation On the contrary in Pt/␥-Al2 O3 catalyst, sharp peaks for Pt could
results and hence shows slightly less activity than Pt/HT catalyst. be seen at characteristic peak positions. Diffraction through lattice
As discussed above glucose may undergo Lobry de Bruyn- plane (1 1 1) gives peak at 39.70◦ , similarly peak at 46.20◦ , 67.22◦ ,
Alberda van Ekenstein transformation (Fig. 3) under alkaline 81.22◦ and 85.67◦ correspond to lattice plan (2 0 0), (2 2 0), (3 1 1)
medium to yield enolate ions, enediol and fructose formation. and (2 2 2). These sharp peaks imply that Pt has agglomerated to
To prove that the higher concentrations of open chain form glu- form big particles. The additional peak at 24.5◦ in spent catalyst
cose is possible under alkaline conditions we performed UV studies. can originate from absorbed carbon species.
From the literature it is very well known that glucose is UV active The TEM images and particle size distribution of fresh Pt/HT cat-
if it is present in open chain form due to exposure of CHO group. alyst (Fig. 6) indicate that Pt nanoparticles of average size of 1.9 nm
Based on this fact we performed UV of glucose (0.233 g) dissolved in are highly dispersed on HT support. On the other hand, in case of
only water (10 mL) under neutral conditions (pH = 6.7). The peak at Pt/␥-Al2 O3 catalyst, Pt nanoparticles with average size of 20–30 nm
max = 266.2 nm was visible as a small hump (Fig. 4). However, after are present. This information points out that the Pt has a strong
addition of mineral base (Na2 CO3 , 0.066 g) in the glucose solution interaction with basic support (HT) which helps to have higher
(glucose, 0.233 g in 10 mL water: same glucose concentration solu- dispersions and stabilization of metal particles (Section 3.1.2). The
tion as was used for oxidation reaction) the peak at max = 267.8 nm oxidation reactions are structure sensitive [29], and thus smaller
was increased many folds. This proves that under alkaline condi- particle size of Pt obtained over Pt/HT catalysts are naturally more
tions, glucose is present in open chain form. Moreover, we carried active than the Pt/␥-Al2 O3 catalyst.

Pt
HT (111)
(200)
Pt Pt
HT
(200) (220) Spent
(220) Pt/γ-Al2 O3
Spent Pt
HT Pt/HT (311)
intensity (a.u.)

(222)
Intensity (a.u)

Fresh
Fresh
Pt/γ-Al2 O3
Pt/HT

γ-Al2 O3
HT

10 20 30 40 50 60 70 80 90 10 20 30 40 50 60 70 80 90
o o
2θ 2θ

Fig. 5. XRD patterns of fresh and spent catalysts used in hydrogenation reaction.

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx 9

Fig. 6. TEM images of fresh and spent Pt/␥-Al2 O3 and Pt/HT catalysts.

Table 5
Summary on catalyst characterizations.

Entry Catalyst Surface area, m2 /g Basicity, mmol/ga Acidity, mmol/gb pHc

1 HT 212 0.88 – 9.24


2 Pt/HT-Fresh 159 0.81 – 9.30
3 Pt/HT-Spentd 73 1.02 – –
4 ␥-Al2 O3 154 – 0.39 –
5 Pt/␥-Al2 O3 -Fresh 174 – 0.35 5.52
6 Pt/␥-Al2 O3 -Spentd 132 – – –
7 Pt/␥-Al2 O3 + HT – – – 8.50
a
Basicity calculated from TPD-CO2 .
b
Acidity calculated from TPD-NH3 .
c
pH was observed by suspending 0.075 g catalyst in 30 mL water.
d
Spent catalyst from glucose oxidation reaction.

Table 5 summarizes the results on the characterization of fresh from oxidation reaction showed loss of only 0.03% of Pt in the
and spent catalysts. It can be noticed that surface area of bare solution in the case of Pt/␥-Al2 O3 and 0.018% in the case of Pt/HT
HT decreases to 159 m2 /g after loading of Pt when used in oxida- catalyst.
tion reaction. However, after the oxidation reaction, surface area
decreases to 73 m2 /g. This implies that catalyst may undergo some
structural changes during the reaction. In Pt/␥-Al2 O3 -spent catalyst 4. Conclusion
marginal decrease in surface area was seen.
The temperature programmed desorption study done with car- Catalytic oxidation and hydrogenation over Pt supported on
bon dioxide as a probe molecule gave 0.88 mmol/g as basicity value basic supports can offer a convenient way to advance the valoriza-
for bare HT catalyst (Table 5). The basicity remains almost constant tion of glucose and xylose to yield acids and alcohols. The oxidation
after Pt loading (0.81 mmol/g). The TPD-NH3 was done to check the carried out at 50 ◦ C is a feasible way to achieve as high as 83% glu-
acidity of ␥-Al2 O3 and Pt/␥-Al2 O3 catalysts (Table 5). The pH val- conic acid yield by suppressing the side reactions. Sugar alcohols
ues for Pt/HT, Pt/␥-Al2 O3 are 9.30 and 5.52, respectively (Table 5). yield of 68% could be achieved with Pt/␥-Al2 O3 catalyst suspended
However addition of HT to Pt/␥-Al2 O3 changes the pH to 8.5. This with basic support (HT). The basicity of support might be helpful
is natural since HT has basicity (pH 9.24). in stabilizing metal particles to give higher dispersions and make
The possibility of leaching of Pt during reactions was studied them electron rich which helps in attaining higher yields. The alka-
by ICP-OES technique. In case of Pt/␥-Al2 O3 catalysts 3.3% and in line medium is useful to build higher concentration of glucose in
Pt/HT 3.2% Pt loading in fresh catalysts was found which is in line open chain form which readily gets converted to corresponding
with the actual loading (3.5%). The spent catalyst used for hydro- acids and alcohols. Hence a combined effect of metal as well as
genation reactions shows similar concentration of Pt implying that acidic and basic sites renders these multifunctional catalysts ideal
there is no leaching of Pt in the solution. This result is corroborated for both hydrogenation and oxidation of substrates like glucose and
by not observing any Pt in the reaction solution. Spent catalyst xylose.

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014
ARTICLE IN PRESS
G Model
MOLCAA-8868; No. of Pages 10

10 A. Tathod et al. / Journal of Molecular Catalysis A: Chemical xxx (2013) xxx–xxx

Acknowledgment [45] R. Albert, A. Strätz, G. Vollheim, Chem. Ing. Tech. 52 (1980) 582–587.
[46] B.W. Hoffer, E. Crezee, P.R.M. Mooijman, A.D. van Langeveld, F. Kapteijn, J.A.
Moulijn, Catal. Today 79–80 (2003) 35–41.
Author, A. T. thanks UGC for senior research fellowship. [47] B. Kusserow, S. Schimpf, P. Claus, Adv. Synth. Catal. 345 (2003) 289–299.
[48] K.V. Gorp, E. Boerman, C.V. Cavenaghi, P.H. Berben, Catal. Today 52 (1999)
Appendix A. Supplementary data 349–361.
[49] J. Wisniak, R. Simon, Ind. Eng. Chem. Prod. Res. Dev. 18 (1979) 50–57.
[50] A. Perrard, P. Gallezot, J.P. Joly, R. Durand, C. Baljou, B. Coq, P. Trens, Appl. Catal.,
Supplementary data associated with this article can be A 331 (2007) 100–104.
found, in the online version, at http://dx.doi.org/10.1016/ [51] B.J. Arena, Appl. Catal., A 87 (1992) 219–229.
[52] M. Makkee, A.P.G. Kieboom, H. van Bekkum, Starch 37 (1985) 232–241.
j.molcata.2013.09.014. [53] J.H. Kim, K.C. Han, Y.H. Koh, Y.W. Ryu, J.H. Seo, J. Ind. Microbiol. Biotechnol. 29
(2002) 16–19.
[54] D.K. Oh, S.Y. Kim, J.H. Kim, Biotechnol. Bioeng. 58 (1998) 440–444.
References [55] J. Zhang, A. Geng, C. Yao, Y. Lu, Q. Li, Bioresour. Technol. 105 (2011) 134–141.
[56] M.G.A. Felipe, M. Vitolo, I.M. Mancilha, S.S. Silva, Biomass Bioenergy 13 (1997)
[1] Y. Lin, S. Tanaka, Appl. Microbiol. Biotechnol. 69 (2006) 627–642. 11–14.
[2] P. Gallezot, N. Nicolaus, G. Fleche, P. Fuertes, A. Perrard, J. Catal. 180 (1998) [57] S. Kumar, S. Singh, I. Mishra, D. Adhikari, J. Ind. Microbiol. Biotechnol. 36 (2009)
51–55. 1483–1489.
[3] R. Geyer, P. Kraak, A. Pachulski, R. Schödel, Chem. Ing. Tech. 84 (2011) 513–516. [58] J.P. Mikkola, T. Salmi, A. Villela, H. Vainio, P. Maki Arvela, A. Kalantar, T. Ollon-
[4] J.N. Chheda, Y. Roman-Leshkov, J.A. Dumesic, Green Chem. 9 (2007) 342–350. qvist, J. Vayrynen, R. Sjoholm, Braz. J. Chem. Eng. 20 (2003) 263–271.
[5] R.D. Cortright, R.R. Davda, J.A. Dumesic, Nature 418 (2002) 964–967. [59] H.M. Baudel, C.A.M.D. Abreu, C.Z. Zaror, J. Chem. Technol. Biotechnol. 80 (2005)
[6] R.R. Davda, J.A. Dumesic, Chem. Commun. (2004) 36–37. 230–233.
[7] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, Appl. Catal., [60] J. Zhang, S. Wu, Y. Liu, B. Li, Catal. Commun. 35 (2013) 23–26.
B 56 (2005) 171–186. [61] H. Hattori, Chem. Rev. 95 (1995) 537–558.
[8] M. Comotti, C.D. Pina, M. Rossi, J. Mol. Catal. A: Chem. 251 (2006) 89–92. [62] C.B. Dartt, M.E. Davis, Catal. Today 19 (1994) 151–186.
[9] M. Wenkin, R. Touillaux, P. Ruiz, B. Delmon, M. Devillers, Appl. Catal., A 148 [63] K. Ebitani, K. Motokura, K. Mori, T. Mizugaki, K. Kaneda, J. Org. Chem. 71 (2006)
(1996) 181–199. 5440–5447.
[10] Y. Önal, S. Schimpf, P. Claus, J. Catal. 223 (2004) 122–133. [64] Y. Ono, T. Baba, Catal. Today 38 (1997) 321–337.
[11] A. Abbadi, H. van Bekkum, J. Mol. Catal. A: Chem. 97 (1995) 111–118. [65] Y. Duan, D. Xia, Y. Xiang, X. Zhang, J. Colloid Interface Sci. 281 (2005) 197–200.
[12] S. Karski, T. Paryjczak, I. Witonnska, Kinet. Catal. 44 (2003) 618–622. [66] Z. Gao, J. Zhou, F. Cui, Y. Zhu, Z. Hua, J. Shi, Dalton Trans. 39 (2010)
[13] S. Ramchandran, Food Technol. Biotechnol. 44 (2006) 185–195. 11132–11135.
[14] K. Matsuno, Occup. Med. 46 (1996) 313–317. [67] B. Veldurthy, J.M. Clacens, F. Figueras, Adv. Synth. Catal. 347 (2005) 767–771.
[15] J.Z. Liu, L.P. Weng, Q.L. Zhang, H. Xu, L.N. Ji, Biochem. Eng. J. 14 (2003) 137–141. [68] J.M. Fraile, J.I. Garcia, J.A. Mayoral, F. Figueras, Tetrahedron Lett. 37 (1996)
[16] P. Beltrame, M. Comotti, C.D. Pina, M. Rossi, J. Catal. 228 (2004) 282–287. 5995–5996.
[17] J. Bao, K. Furumoto, K. Fukunaga, K. Nakao, Biochem. Eng. J. 8 (2001) 91–102. [69] K. Motokura, N. Fujita, K. Mori, T. Mizugaki, K. Ebitani, K. Kaneda, J. Am. Chem.
[18] C. Baatz, B. Thielecke, Y. Pru␤e, Appl. Catal., B 70 (2007) 653–660. Soc. 127 (2005) 9674–9675.
[19] C. Baatz, U. Pru␤e, J. Catal. 249 (2007) 34–40. [70] M. Lakshmi Kantam, B.M. Choudary, C. Venkat Reddy, K. Koteswara Rao, F.
[20] M. Wenkin, P. Ruiz, B. Delmon, M. Devillers, J. Mol. Catal. A: Chem. 180 (2002) Figueras, Chem. Commun. (1998) 1033–1034.
141–159. [71] P. Liu, Y. Guan, R.A.V. Santen, C. Li, E.J.M. Hensen, Chem. Commun. 47 (2013)
[21] S. Hermans, M. Devillers, Appl. Catal., A 235 (2002) 253–264. 11540–11542.
[22] S. Biella, L. Prati, M. Rossi, J. Catal. 206 (2002) 242–247. [72] L.B. Kunde, S.M. Gade, V.S. Kalyani, S.P. Gupte, Catal. Commun. 10 (2009)
[23] M. Comotti, C. Della Pina, E. Falletta, M. Rossi, Adv. Synth. Catal. 348 (2006) 1881–1888.
313–316. [73] D.G. Evans, X. Duan, Chem. Commun. 0 (2006) 485–496.
[24] M. Comotti, C. Della Pina, R. Matarrese, M. Rossi, Angew. Chem. 116 (2004) [74] C.M. Jinesh, C.A. Antonyraj, S. Kannan, Appl. Clay Sci. 48 (2010) 243–249.
5936–5939. [75] P.S. Kumbhar, J. Sanchez-Valente, J.M.M. Millet, F.O. Figueras, J. Catal. 191 (2000)
[25] N. Thielecke, K.D. Vorlop, U. Pru␤e, Catal. Today 122 (2007) 266–269. 467–473.
[26] M. Comotti, C. Della Pina, E. Falletta, M. Rossi, J. Catal. 244 (2006) 122–125. [76] R. Chitrakar, S. Tezuka, A. Sonoda, K. Sakane, T. Hirotsu, Ind. Eng. Chem. Res. 47
[27] P. Beltrame, M. Comotti, C. Della Pina, M. Rossi, Appl. Catal., A 297 (2006) 1–7. (2008) 4905–4908.
[28] T. Ishida, N. Kinoshita, H. Okatsu, T. Akita, T. Takei, M. Haruta, Angew. Chem. [77] W. Meng, F. Li, D.G. Evans, X. Duan, Mater. Res. Bull. 39 (2004) 1185–1193.
120 (2008) 9405–9408. [78] D.P. Debecker, E.M. Gaigneaux, G. Busca, Chem. Eur. J. 15 (2009) 3920–3935.
[29] H. Okatsu, N. Kinoshita, T. Akita, T. Ishida, M. Haruta, Appl. Catal., A 369 (2009) [79] V. Choudhary, R. Jha, P. Choudhari, J. Chem. Sci. 117 (2005) 635–639.
8–14. [80] C. Moreau, R. Durand, A. Roux, D. Tichit, Appl. Catal., A 193 (2000) 257–264.
[30] I. Nikov, K. Paev, Catal. Today 24 (1995) 41–47. [81] S. Lima, A.S. Dias, Z. Lin, P. Brandao, P. Ferreira, M. Pillinger, J. Rocha, V. Calvino-
[31] A. Mirescu, U. Pru␤e, Appl. Catal., B 70 (2007) 644–652. Casilda, A.A. Valente, Appl. Catal., A 339 (2008) 21–27.
[32] A.T. Governo, L. Proenca, P. Parpot, M.I.S. Lopes, I.T.E. Fonseca, Electrochim. Acta [82] R.O.L. Souza, D.P. Fabiano, C. Feche, F. Rataboul, D. Cardoso, N. Essayem, Catal.
49 (2004) 1535–1545. Today 195 (2012) 114–119.
[33] US Department of Energy, Energy Efficiency and Renewable Energy, Top Value [83] R. Sahu, P.L. Dhepe, 2013 (submitted for publication).
Added Chemicals From Biomass, Vol. 1: Results of Screening for Potential [84] H.A. Pary, C.E. Schweickert, B.H. Minnich, Ind. Eng. Chem. 44 (1952) 1146–1151.
Candidates from Sugars and Synthesis Gas, 2004 http://eereweb.ee.doe.gov/ [85] C. Moreau, J. Lecomte, A. Roux, Catal. Commun. 7 (2006) 941–944.
biomass/pdfs/35523.pdf [86] F. King, G.J. Kelly, Catal. Today 73 (2002) 75–81.
[34] J.P. Mikkola, T. Salmi, Chem. Eng. Sci. 54 (1999) 1583–1588. [87] S. Coluccia, A.J. Tench, Proceeding of the 7th International Congress on Catalysis,
[35] J.P. Mikkola, T. Salmi, Catal. Today 64 (2001) 271–277. Tokyo, 1980, p. 1160.
[36] L. He, D. Chen, ChemSusChem 5 (2011) 587–595. [88] M. Kappers, C. Dossi, R. Psaro, S. Recchia, A. Fusi, Catal. Lett. 39 (1996) 183–189.
[37] J. Sun, H. Liu, Green Chem. 13 (2011) 135–142. [89] F. Humblot, F. Lepeltier, J.P. Candy, J. Corker, O. Clause, F. Bayard, J.M. Basset, J.
[38] P. De Wulf, W. Soetaert, E.J. Vandamme, Biotechnol. Bioeng. 69 (2000) 339–343. Am. Chem. Soc. 120 (1998) 137–146.
[39] M. Yadav, D.K. Mishra, J.S. Hwang, Appl. Catal., A 425–426 (2012) 110–116. [90] Z. Gandao, B. Coq, L.C. de Menorval, D. Tichit, Appl. Catal. A 147 (1996) 395–406.
[40] J. Zhang, A. Geng, C. Yao, Y. Lu, Q. Li, Bioresour. Technol. 105 (2013) 134–141. [91] S. Recchia, C. Dossi, N. Poli, A. Fusi, L. Sordelli, R. Psaro, J. Catal. 184 (1999) 1–4.
[41] F.W. Lichtenthaler, Acc. Chem. Res. 35 (2002) 728–737. [92] J. Adamiec, R.M.J. Fiederow, S.E. Wanke, 74th Annual A.I.Ch.E. Meeting, vol. 59,
[42] T. GranstrÖm, K. Izumori, M. Leisola, Appl. Microbiol. Biotechnol. 74 (2007) New Orleans, Paper 13A, Fiche, 1981.
273–276. [93] J.M.H. Dirkx, H.S. van der Baan, J. Catal. 67 (1981) 1–13.
[43] P. Gallezot, P.J. Cerino, B. Blanc, G. Fleche, P. Fuertes, J. Catal. 146 (1994) 93–102. [94] G. de Wit, C. de Haan, A.P.G. Kieboom, H. van Bekkum, Carbohydr. Res. 86 (1980)
[44] M. Herskowitz, Chem. Eng. Sci. 40 (1985) 1309–1311. 33–41.

Please cite this article in press as: A. Tathod, et al., J. Mol. Catal. A: Chem. (2013), http://dx.doi.org/10.1016/j.molcata.2013.09.014

Das könnte Ihnen auch gefallen