You are on page 1of 11

Research Article

Received: 29 July 2010

Revised: 14 October 2010

Accepted: 16 November 2010

Published online in Wiley Online Library: 6 January 2011

(wileyonlinelibrary.com) DOI 10.1002/jctb.2565

Sugar hydrogenation over a Ru/C catalyst


Kordas,
b
Vctor A. Sifontes Herrera,a Oluwamuyiwa Oladele,a Krisztian
a Jyri-Pekka Mikkola,a,c Dmitry Yu. Murzina and Tapio Salmia

Kari Eranen,
Abstract
BACKGROUND: In recent years, exploitation of renewable resources has gained considerable attention. In this respect, polyols
derived from the hydrogenation of sugar molecules are versatile molecules with a variety of uses, such as low-caloric sweeteners.
The hydrogenation of D-maltose, D-galactose, L-rhamnose and L-arabinose was carried out on a finely dispersed Ru/activated
carbon catalyst with the objective of studying the kinetics of the production of the corresponding polyols. The reactions were
carried out in a stirred tank reactor at temperatures ranging from 90 to 130 C and hydrogen pressures from 40 to 60 bar.
RESULTS: Sugar conversions up to 100% were achieved. Some by-product formation affecting the quality of the selectivity
was also observed at elevated operating conditions. The catalyst was characterized by scanning electron microcopy (SEM),
transmission electron microscopy (TEM), inductively coupled plasma optical emission spectrometry (ICP-OES) and nitrogen
physisorption. Kinetic models based on the Langmuir Hinshelwood assumptions were proposed for the reactions and a
nonlinear regression was performed to obtain the numerical values of the kinetic parameters.
CONCLUSIONS: The kinetic models predicted well the sugar hydrogenation process and the kinetic parameters were established.
The model can be used to predict the behaviour of batchwise operating slurry reactors.
c 2011 Society of Chemical Industry

Keywords: D-maltose; D-galactose; L-rhamnose; L-arabinose; hydrogenation; ruthenium catalyst; reduction; kinetic modeling

NOTATION
ccalculatedi
cexperimentali
cmean
cmodeli
S
SOH
A1
A2
Ea1
Ea2
k1
k2
KH 2
KS
KSOH

658

P H2

R
concentration calculated by the model
during parameter estimation [%(w/w)]
experimental concentration [%(w/w)]
mean concentration for all the experimental values [%(w/w)]
concentration calculated by the model
[%(w/w)]
sugar molecule
polyol molecule
pre-exponential factor of the sugar-topolyol reaction [s1 ]
pre-exponential factor of the maltitol-tosorbitol reaction [s1 ]
activation energy of the sugar-to-polyol
reaction [J mol1 ]
activation energy of the maltitol-to-sorbitol
reaction [J mol1 ]
kinetic constant of the sugar-to-polyol
reaction [s1 ]
kinetic constant of the maltitol-to-sorbitol
reaction [s1 ]
hydrogen adsorption equilibrium constant
[bar1 ]
sugar adsorption equilibrium constant
[%(w/w)1 ]
sugar alcohol adsorption equilibrium constant [%(w/w)1 ]
hydrogen pressure [bar]

J Chem Technol Biotechnol 2011; 86: 658668

r1
S
SOH

H2
S
SOH
cat

universal gas constant [8.314472 JK1


mol1 ]
reaction rate of the sugar hydrogenation
reaction [%(w/w)/s]
sugar molecule concentration [%(w/w)]
polyol molecule concentration [%(w/w)]
Greek letters
catalyst free site
adsorbed hydrogen molecule
adsorbed sugar species
adsorbed polyol species
catalyst bulk density [g/mL]

INTRODUCTION
It is becoming increasingly evident that the current trends in
industrial chemistry point towards the optimized exploitation
of renewable raw materials. As such, carbohydrates constitute
three-quarters of the available renewable biomass1 and are thus

Correspondence to: Tapio Salmi, Laboratory of Industrial Chemistry and


Reaction Engineering Process Chemistry Centre, bo Akademi University FI20500 Turku/bo, Finland. E-mail: tapio.salmi@abo.fi.

a LaboratoryofIndustrial ChemistryandReaction Engineering,ProcessChemistry


Centre, bo Akademi University FI-20500 Turku/bo, Finland
b Microelectronics and Materials Physics Laboratories, University of Oulu FI90570, Oulu, Finland
c Ume University, Chemistry Department, Chemical and Biochemical Centre,
SE-90187 Ume, Sweden

www.soci.org

c 2011 Society of Chemical Industry




Sugar hydrogenation over a Ru/C catalyst

www.soci.org

J Chem Technol Biotechnol 2011; 86: 658668

in monoliths and on activated carbon cloths.3,16 Furthermore, it


is possible to carry out sugar hydrogenation in loop reactors,
where the liquid, containing the catalyst, is recirculated through
an external loop system.16 Since the sugar molecules are large
and the solubility of hydrogen is limited, the diffusion limitation
inside the porous catalyst layer becomes severe in porous catalyst
layers, as has been demonstrated experimentally and by numerical
simulations.26,27 Therefore, only egg-shell catalyst pellets and
structured catalysts can be considered as feasible alternatives to
slurry technology.
The basis of any kind of technology development is the intrinsic
reaction kinetics, including studies of the product selectivity.
The present work is devoted to the hydrogenation kinetics of
some monosaccharides and disaccharides of industrial potential:
L-arabinose, D-galactose, D-maltose and L-rhamnose on a carbonsupported ruthenium catalyst. The reactants and main reaction
products are illustrated in Fig. 1.

EXPERIMENTAL SECTION
Experimental setup
The conventional production technology of sugar alcohols is
based on the use of batchwise operating slurry reactors: finely
dispersed, supported or sponge metal catalyst (catalyst particles
<0.1 mm) are immersed in a batch of an aqueous sugar solution,
to which hydrogen is continuously added so that the pressure is
kept constant. Among the reported operating conditions, it can be
found that hydrogen pressure is typically kept at 30180 bar and
the temperature ranges from 80 to 150 C.16,21 Hydrogen pressure
is one of the limiting factors of the process; therefore alternative
solvents with better hydrogen solubilities, such as ethanol, have
been proposed.28
Taking the above conditions, the reactions were performed
in a laboratory-scale batch reactor (Parr 4561, 300 mL) equipped
with baffles, a gas entrainment impeller, a sampling line with a
sintered filter (7 m), a cooling coil, a heating jacket, a temperature
and stirring rate controller (Parr 4843), a pressure controller and
microprocessor (Brooks 5866 and Brooks 0154 respectively), and a
bubbling chamber for pre-treating the sugar solutions (Fig. 2).
Kinetic experiments
A series of experiments were carried out with the purpose of
studying the kinetic behaviour of the hydrogenation reactions of
four selected sugars (D-maltose, D-galactose, L-arabinose and Lrhamnose; Danisco Sweeteners, Kantvik, Finland). For each sugar,
a set of experiments was performed under varying conditions of
pressure (40, 50 and 60 bar) and temperature (90, 105, 120 and
130 C) in the presence of a supported Ru/active carbon catalyst
with a loading of 0.91%(w/w) (relative to the dry sugar mass).
A given mass of catalyst was loaded in the vessel and flushed
with nitrogen (99.999%, AGA) to purge oxygen. At the same
time, the sugar was diluted in deionized water. Concentrated
solutions are prone to cause mass transport limitation,24 thus it
was convenient to work with sugar solutions with concentrations
of 10%(w/w) and, in the unique case of D-galactose, 4%(w/w)
(due to the lower solubility of D-galactitol). The sugar solutions
were then bubbled with nitrogen and, immediately afterwards,
with hydrogen. In the next step, the solution was injected into
the vessel, nitrogen was purged with hydrogen, and the system
was finally adjusted to the desired operating conditions. Each
experiment was carried out at a stirring speed of 1800 rpm for 3 h,

c 2011 Society of Chemical Industry




wileyonlinelibrary.com/jctb

659

becoming a key target in the field of green chemistry. However,


not only should naturally occurring carbohydrates become an
important industrial feedstock but also, in order to guarantee
a smooth transition to a biomass-based industry, the products
derived thereof should be developed having alternative raw
materials in mind, as opposed to seeking routes to reproduce
exactly the same chemicals as obtained from other technologies,
e.g. the petrochemical industry.2
Following this strategy, naturally occuring polysaccharides, such
as hemicelluloses, are starting molecules for the derivatization
to simpler monosaccharides and disaccharides, which, in turn,
act as substrates for the production of value added molecules.
Notable among these are the sugar alcohols. These compounds
are obtained through reduction of the carbonyl group in the
sugar molecule either by means of chemical agents (such as
sodium borohydride) or by molecular hydrogen in the presence
of a homogeneous or heterogeneous catalyst. The catalytic route
is preferred for the industrial hydrogenation of sugars to sugar
alcohols, where heterogeneous catalysts based on Ni, Pd, Pt or Ru
can be used.3,4
As an additional motivation, sugar alcohols are versatile
molecules that have found a wide variety of applications. Higher
molecular weight polyols (chains longer than four carbon atoms)
could be used for the manufacture of polyesters, alkyd resins
and polyurethanes.3 They have also been used as intermediates
in the production of pharmaceuticals and starting molecules
for the synthesis of ligands.5 13 However, their most common
use is associated with the fact that polyols possess a sweet
taste, contain fewer calories, and produce a lower glycemic
response. Accordingly, their use as sweeteners is quite desirable
and straightforward.14,15
In addition to the aforementioned, the general publics
awareness of how dietary choices have an impact on the general
quality of life, further explains why research on health-promoting
food ingredients is becoming an important and exciting topic.
This has been further intensified as the segment of the population
with special dietary requirements, such as diabetic patients, is
increasing. As proof of the evolving importance of this topic, the
global production of sorbitol has increased from 700 000 tons per
year in 20073 to over 1 000 000 tons per year in 2009.16
The selective catalytic hydrogenation of naturally occurring
sugar molecules to their corresponding sugar alcohols is an
environmentally friendly route to the production of alternative
sweeteners. Solid metal-based catalysts can be applied in an
aqueous or alcoholic environment, circumventing the use of
stoichiometric reducing agents and the subsequent production of
inorganic salts as waste materials.
Although these reactions are at the core of many important
industrial processes, the availability of literature regarding catalytic
sugar hydrogenation is relatively scarce. However, work on sugar
hydrogenation has been carried out by some authors with a
particular focus on glucose, fructose, xylose and lactose.17 24 The
most common solid catalysts used for sugar hydrogenation to
sugar alcohols are based on nickel, such as sponge nickel (Raney
nickel), but recently the use of ruthenium has received attention,
since Ru has shown good activity and excellent selectivity in the
hydrogenation of sugar molecules.3,20,22,25 Thus, it is expected that
Ru will replace Ni as a sugar hydrogenation catalyst, particularly
because of the toxic properties of Ni and the rapid deactivation of
sponge Ni.
In recent years, sugar hydrogenation in continuous operation
has been demonstrated, for instance, for glucose hydrogenation

www.soci.org

HO
O

HO

VA Sifontes Herrera et al.

OH
H2

OH

HO

OH

D-maltose

D-maltitol
O
OH
HO
HO

H2

OH
D-sorbitol

OH
HO
HO
OH

OH

D-galactose

OH

D-galactitol
O

HO
H2
HO
HO

HO

HO

HO
OH
OH

OH

HO
HO

OH
OH

OH

OH

L-arabinose

OH

OH

OH

HO

OH
OH

HO

OH

OH

H2

OH
HO

OH

OH OH

HO

L-arabitol

H2

OH
OH

HO
HO

HO
HO

L-rhamnose

L-rhamnitol

Figure 1. Hydrogenation of D-maltose, D-galactose, L-arabinose and L-rhamnose.

Hydrogen

since it has been previously shown that similar systems required


a stirring speed of 15001600 rpm to ensure operation in the
kinetically controlled regime.17,18,22,23 Samples were withdrawn
from the system at different intervals in order to perform the
concentration and pH measurements.
Off-line concentration analysis was conducted using a highperformance liquid chromatograph (HP 1100 Series LC) equipped
with a Biorad Aminex HPX-87C carbohydrate column and coupled
to a refractive index detector (HP 1047A). The method employed
consisted of a 2 L sample injection and a mobile phase flow
of 0.400 mL min1 (aqueous solution of CaSO4 1.2 mmol L1 )
at 80 C. The use of CaSO4 as an inorganic modifier has been
suggested to help regenerate the stationary phase calcium ions
that might have been removed by interaction with ions derived
from aldonic acids.29
For some selected samples, a subsequent study to determine
the nature of some unidentified components was carried out by
using a combined gas chromatographymass spectrometer (HP
6890/5973) equipped with a HP-1 column. To achieve this, an
aliquot of 10 L of the sample was diluted in acetone and dried
first under a flow of nitrogen for 45 min and then placed in a
vacuum oven for 30 min at 40 C. The next step was to silylate the
samples using pyridine, HMDS and TMCS so as to make the sugars
more volatile and ready for the GC/MS analysis.

Nitrogen

PT

PC

Sugar solution
1

Sample

2
TC

TT

Cooling Water

Figure 2. Reaction system: (1) bubbling chamber, (2) reaction vessel.

660
wileyonlinelibrary.com/jctb

Catalyst characterization
Nitrogen adsorption. The surface area of the catalyst was
determined by nitrogen physisorption (99.999%, AGA) using a
Sorptomatic 1900 by Carlo Erba Instruments. First, the samples
were outgassed under vacuum at 150 C for 3 h and the
adsorption/desorption steps were carried at 77 K, using liquid
nitrogen as coolant. The data obtained were interpreted with the
BrunauerEmmettTeller (BET) isotherm.
Inductively coupled plasma optical emission spectrometry (ICPOES). The ruthenium content of the catalyst was determined with

c 2011 Society of Chemical Industry




J Chem Technol Biotechnol 2011; 86: 658668

Sugar hydrogenation over a Ru/C catalyst

www.soci.org

Table 1. Catalyst characterization results


Property
Catalyst
Surface area
Ruthenium content
Ruthenium particle size,
determined by TEM
(Fig. 4)
Ruthenium leaching
Particle shape
Particle average size
Particle average
thickness

Value
Ruthenium over active carbon
690 m2 g1
2.5%(w/w)
23 nm (mono-dispersed particles)
<0.9 mg L1
Flakes
10 m 10 m
12 m

ICP-OES (Optima 5300DV). In addition, metal leaching from the


catalyst was studied on a set of samples taken at the end of the
selected kinetic runs.
Scanning electron microscopy (SEM). Catalyst images as well
as elemental analyses were obtained with a Leo Gemini 1530
scanning electron microscope equipped with a ThermoNORAN
Vantage X-ray detector for EDXA analysis. The images were taken
using the secondary electron and backscattered electron detectors
at 15 kV, and the in-lens secondary electron detector at 2.70 kV.
Transmission electron microscopy (TEM) and X-ray diffraction
(XRD). An energy-filtered transmission electron microscope (EFTEM
LEO 912 OMEGA, 120 kV) and a scanning electron microscope
(SEM/EDS, Jeol JSM-6400) to study the microstructure of the
support as well as the distribution, the size and the concentration
of the catalyst. The crystal structure and the average size of catalyst
were determined by using X-ray diffraction (Siemens D5000 with
CuK radiation).

CATALYST CHARACTERIZATION RESULTS


A fresh commercial Ru/activated carbon catalyst was employed
for the kinetic experiments and it consisted of a powder made up
of small flakes with a cross-section of 10 10 m and an average
thickness of 12 m. BET analysis gave a total surface area of
690 m2 g1 .
The active metal content was determined to be 2.5%(w/w) by
ICP-OES. TEM studies revealed that ruthenium was present as
particles with diameters ranging from 2 to 3 nm.
After conducting each experiment, the contents of the reactor
were stored and a number of samples were withdrawn and filtered,
and the solution was analysed using ICP-OES in order to discover
traces of leached ruthenium. The analysis showed that the metal
content in the solution was below the detection limit of the
instrument (0.9 mg L1 ), which confirms that leaching, if any, was
negligible.
The characterization results are summarized in Table 1. Some
SEM and TEM images of the catalyst are provided in Figs 3 and 4.

QUALITATIVE KINETICS

J Chem Technol Biotechnol 2011; 86: 658668

Pressure effect
To reveal the influence of the hydrogen pressure on the
hydrogenation kinetics, several experiments were made with each
sugar at 40, 50 and 60 bar.
As Fig. 5 shows, an increase in the operating pressure results
in higher reaction rates. However, it can be observed that the
influence of pressure for any given sugar is not highly significant,
and D-galactose exhibits an almost invariant behaviour, i.e. close to
zero order with respect to hydrogen pressure. This is in agreement
with the hydrogenation behaviour of xylose over a range of
different catalysts, in which the kinetics become independent of
the operating pressure at values exceeding 40 bar.17
Temperature effect
It was possible to draw conclusions about the influence of
temperature on the hydrogenation kinetics (Fig. 6). As can be
observed, the reaction temperature has a significant effect on
hydrogenation, with the rates becoming considerably faster as the
temperature increases. This remains true for the four sugars. The
apparent activation energies were in the range 5684 kJ mol1
(Table 3). The orders of magnitude of these activation energies
are in accordance with values reported in the literature for the
hydrogenation of D-glucose.4,24,25,30
Products and by-products
Typically, the selectivity towards the sugar alcohol is high on
both sponge Ni and Ru catalysts, exceeding 95% under optimal
conditions. Such data has been reported for the hydrogenation of
glucose, xylose and lactose.18,22,24,28,31 For some special cases, such
as hydrogenation of fructose to maltitol, the product selectivity
is a serious issue, the highest selectivities are limited to about
6070% on Cu-based catalysts.32
It was possible to reach complete conversion of the sugar
molecules to yield the desired sugar alcohols as predominant
products. This qualitative observation is based primarily on direct
inspection of the chromatograms. Detailed information on the
by-products formed is still required.
By-product yields were observed to be dependent on the
operating conditions: the more severe the conditions (higher
pressures and temperatures), the more diverse and abundant the
by-products, amounting to ca 1012% of the total product yield
under the more severe conditions for D-maltose and L-rhamnose
and to ca 57% for L-arabinose and D-galactose under the same
severe conditions. However, in the majority of conditions, the
by-product yield <1%.
Moreover, by-product formation was very much dependent on
the sugar in question: for L-arabinose and D-galactose, a relatively
negligible by-product yield was observed, whereas in the case of
D-maltose and L-rhamnose, the by-product formation was more
important, though still significantly smaller than the main (desired)
product.

c 2011 Society of Chemical Industry




wileyonlinelibrary.com/jctb

661

The hydrogenation behaviour of the four selected sugars was


similar in terms of the pressure and temperature influence. As can
be seen from the data obtained (Figs 5 and 6), the effect of the
hydrogen pressure was minor, whereas the temperature effect
was very profound.

As will be described in more detail in the next section, the


sugar hydrogenation proceeds via a LangmuirHinshelwood
mechanism and exhibits a concentration-dependent reaction
order, which ranges from zero order at high concentrations to first
order at a low concentrations. However, the hydrogen pressure
dependence was slight, which could explain the absence of a
maximum in the reaction rate (as would be expected for a reaction
proceeding via this type of mechanism).

www.soci.org

VA Sifontes Herrera et al.

Figure 3. SEM images of the Ru/C catalyst at different scales: (a) 250, (b) 1000 and (c) 50000.

Figure 4. TEM images of the Ru/C catalyst.

In the case of D-maltose, the most relevant by-product, which


was readily identified during HPLC analysis, was D-sorbitol and
the complete reaction path starting from D-maltose is displayed
in Fig. 1.
Variation of pH
The pH remained acidic during experiments on the four sugars,
with a maximum observed variation of ca 2 units. Additionally,
it was noted that as the conversion progressed, the pH value
decreased until a minimum was reached, threafter it increased up
to approximately the starting value. The lowest pH value coincided
with complete conversion (Figure 7).
Consequently, this behaviour is similar to that of the reaction
rate meaning that it is heavily influenced by the temperature but
rather insensitive to the hydrogen pressure. This is true for all
four sugars and suggests that the change in pH is mostly due to
the chemical transformations occurring in the reaction media and
may point to the formation of sugar acids such as aldonic acids,
which could then be either decomposed or hydrogenated, thus
restoring the pH to the initial values.
The ideal pH for sugar hydrogenation should not exceed 6.5
because otherwise it will either lead to isomerization of the starting
sugars or promote a Canizzaro type disproportionation.21,33,34
During the studied hydrogenations, the pH never increased
beyond a value of 6, which helped in avoiding the by-product
formation.

KINETIC MODELING

662

When it comes to understanding the sugar hydrogenation


mechanism, the question about which of the sugar tautomers are
involved in the hydrogenation reaction remains to be answered, as
there have been differing opinions regarding this matter. Catalytic
hydrogenation of D-fructose has been suggested to proceed via
a closed ring formation19,20,35 with some authors proposing the

wileyonlinelibrary.com/jctb

formation of -fructofuranose.36 However, there is also some more


general information suggesting that hydrogenation of sugars
proceeds via an open-chain formation.37
Furthermore, there have been several attempts to find a
mechanistically sound expression that describes the sugar hydrogenation. For the most common cases of D-glucose, D-lactose
and D-xylose, it is possible to find studies and reviews that offer
different ways of treating the reaction kinetics.4,17,24,25,30 32,38,39
In the majority of these studies, the surface reactions are claimed
to proceed via a LangmuirHinshelwood mechanism. However,
there are differing opinions regarding the way hydrogen is adsorbed on the surface (molecularly or atomically) and regarding
the occupation of active sites by hydrogen and the sugar molecule:
(a) sugar and hydrogen compete for adsorption on the same type
of sites or (b) sugar and hydrogen adsorb on different sites due to
structural differences. In this respect, it is also worth mentioning
that, by coupling quantum calculations and classic kinetic studies, it is possible to determine the amount of sites required by
each molecule in order to adsorb and react on an active metal
cluster.36,40,41
Based on the abovementioned arguments, some rival hypotheses were compared in the derivation of the rate equation for
the hydrogenation of L-arabinose, D-galactose, D-maltose and
L-rhamnose. Such models consisted of LangmuirHinshelwood
with non-competitive (Equation (1)) and competitive adsorption
(Equation (2)) of reactants, and a semi-empirical model, Equation (3). Molecular adsorption of hydrogen was assumed in all
of the models. Details about the general methods employed
for the parameter estimation of the aforementioned models are
presented in this section later on.
r1 =
r1 =
r1 =

(1 + KH2

k KS KH2 [S] PH2


PH2 ) (1 + KS [S] + KSOH [SOH ])
k KS KH2 [S] PH2

(1 + KH2 PH2 + KS [S] + KSOH [SOH ])2


k KS KH2 [S] PH2
(1 + KH2 PH2 + KS [S] + KSOH [SOH ])

(1)
(2)
(3)

The coefficients of determination for the models tested are


presented in Table 2. The coefficient of determination, calculated
from Equation (4), compares the model performance with respect
to the variance of all experimental points. The coefficients of
determination vary between 0 and 100%. A good deterministic
model describing the reaction kinetics should have a high
coefficient of determination, typically exceeding 9597%. At first
glance, all the models were comparable (Table 2). The description
of the simple first-order model is poorer than that of the other
ones, particularly because the experimentally observed rates were

c 2011 Society of Chemical Industry




J Chem Technol Biotechnol 2011; 86: 658668

Sugar hydrogenation over a Ru/C catalyst

40 bar

50 bar

(b)

60 bar

40 bar

100 %

100 %

80 %

80 %
Conversion [%]

Conversion [%]

(a)

www.soci.org

60 %

40 %

60 bar

60 %

40 %

20 %

20 %

0%

0%
0

8 10 12 14 16 18 20 22

(c)

40 bar

50 bar

8 10 12 14 16 18 20 22

Adjusted Time [gmin]

Corrected Time [gmin]


(d)

60 bar

40 bar

100 %

100 %

80 %

80 %
Conversion [%]

Conversion [%]

50 bar

60 %

40 %

50 bar

60 bar

60 %

40 %

20 %

20 %

0%

0%
0

Corrected Time [gmin]

8 10 12 14 16 18 20 22

Adjusted Time [gmin]

Figure 5. Effect of pressure on hydrogenation rates at 120 C for (a) L-arabinose, (b) D-maltose, (c) D-galactose and (d) L-rhamnose.

Table 2. Coefficients of determination for the models studied

L-arabinose
D-galactose
D-maltose
L-rhamnose

Competitive
L-H

Non-competitive
L-H

Semi-empirical

98.67
97.34
98.58
94.15

95.20
96.78
98.55
94.23

97.99
97.27
98.47
94.00

not of first order with respect to hydrogen, but of a lower order.


n

(cexperimentali cmodeli )2

i=1
100%
1

(4)
R2 =

n


2
(cexperimentali cmean )

side reactions was a more difficult task. In the case of D-maltose, the
activation energy for D-maltitol cleavage to yield D-sorbitol was
rather poorly identified. Despite these shortcomings, the model
can satisfactorily describe the experimental data.
Based on the previous evaluation, we propose a LangmuirHinshelwood type of mechanism that assumes competitive
adsorption of any of the four sugars studied and molecular hydrogen. The mass balances follow a simple reaction network since, as
was noted in the previous section, the reactions either proceeded
without any appreciable by-product yield or the by-products
were not quantified in a manner that would justify any detailed
modeling.
In agreement with the conventional assumptions, the reactions
on the catalyst surface are illustrated as follows:
The adsorptiondesorption quasi-equilibria for the involved
molecules are expressed by:

i=1

KSOH
KH 2

(5)
(6)
(7)

The balance of the catalyst sites is necessary to relate all the


species adsorbed on the surface:
+ S + SOH + H2 = 1

c 2011 Society of Chemical Industry




(8)

wileyonlinelibrary.com/jctb

663

A sensitivity analysis was performed by evaluating the impact of


each parameter on the objective function. This analysis revealed
that not only had the competitive LangmuirHinshelwood model
(Equation (2)) a relatively better agreement with the experimental
data, but also it was shown that the majority of the estimated
parameters had an important contribution to the overall model
performance, thus leading to a better identifiability.
In general, the parameters of the best models were well
identified. However, determination of the rate parameters for the

J Chem Technol Biotechnol 2011; 86: 658668

S
S
SOH
=
SOH
H2
=
H2

KS =

www.soci.org
90 C

105 C

120 C

130 C

90 C

(b)

100 %

100 %

80 %

80 %
Conversion [%]

Conversion [%]

(a)

VA Sifontes Herrera et al.

60 %

40 %

105 C

60 %

40 %

0%

0%
0

8 10 12 14 16 18 20 22

Adjusted Time [gmin]


90 C

105 C

120 C

8 10 12 14 16 18 20 22

Adjusted Time [gmin]


130 C

90 C

(d)

100 %

100 %

80 %

80 %
Conversion [%]

Conversion [%]

130 C

20 %

20 %

(c)

120 C

60 %

40 %

105 C

120 C

130 C

60 %

40 %

20 %

20 %

0%

0%
0

Adjusted Time [gmin]

10

15

20

Adjusted Time [gmin]

Figure 6. Effect of temperature on the hydrogenation rates at 60 bar for (a) L-arabinose, (b) D-maltose, (c) D-galactose and (d) L-rhamnose.
Maltose conversion

pH
3.8

80 %
3.7
60 %
3.6

40 %
20 %

pH

Conversion [%]

100 %

3.5

0%

3.4
0

20

40

60

80 100 120 140 160 180


Time [min]

Figure 7. Maltose hydrogenation at 130 C and 60 bar.

KS

1. q + S

KH 2

2. q + H2
3. qH2 + qS
4. qSOH

KSOH

in Fig. 8), and employing Equations (5)(8), the rate equation


becomes (for any sugar molecule S and its corresponding sugar
alcohol SOH ):

qS
qH2
K

r1 = k S H2 = k Ks [S] KH2 PH2 2 =


k KS KH2 [S] PH2
=
(1 + KH2 PH2 + KS [S] + KSOH [SOH ])2

qSOH + q

SOH + q

where the rate constant, ki , is obtained with the following Arrhenius


equation:

Figure 8. Surface reaction steps.

ki = Ai e

664

Assuming that the rate determining step is the surface reaction


between adsorbed species of the sugar and hydrogen (step 3

wileyonlinelibrary.com/jctb

(9)

Eai
RT

(10)

By examining Equation (9) and the values presented in Table 3,


it is possible to explain the lower influence of the hydrogen

c 2011 Society of Chemical Industry




J Chem Technol Biotechnol 2011; 86: 658668

Sugar hydrogenation over a Ru/C catalyst

www.soci.org

Table 3. Regressed model parameters (the units are given in Notation)


Parameter

L-arabinose
D-galactose
D-maltose
L-rhamnose

D-galactose

Ea1

A1

KS

KSOH

KH2

Ea2

A2

57200
3.6%
60500
2.4%
83600
2.2%
56800
3.6%

43.81
100%
20.35
100%
24.26
100%
11.51
100%

17.70
24.6%
157.00
100%
2.18
96.3%
101.00
100%

5.60
29.8%
61.00
100%
0.47
100%
62.70
100%

0.01
89.9%
0.01
100%
0.01
30.0%
0.05
100%

4340.00
100%

4.49
100%

L-arabinose

D-maltose

molecules. Nonetheless, it is also possible to note that the preexponential factors diverge in orders of magnitude among the
different sugars, which could, a priori, indicate a certain level of
correlation between parameters.

L-rhamnose

100 %

80 %
Conversion [%]

Q=

n

(cexperimentali ccalculatedi )2

(13)

i=1

60 %

40 %

20 %

0%
0

20

40

60

80 100 120 140 160 180 200


Time [min]

Figure 9. Hydrogenation of the four sugars at 105 C and 50 bar.

pressure, represented by a low adsorption constant (KH2 ). Owing


to the aforementioned weak dependence, no maximum of the
reaction rate, as a function of the operative pressure, was observed
in the range studied.
The mass balances of the sugar and sugar alcohol derived from
the reaction path shown in Fig. 1 takes the following form:
d[S]
= r1 cat
dt
d[SOH ]
= r1 cat
dt

(11)
(12)

J Chem Technol Biotechnol 2011; 86: 658668

CONCLUSIONS
The hydrogenation of D-maltose, D-galactose, L-arabinose and
L-rhamnose was successfully performed on a carbon-supported Ru
catalyst in a pressurized slurry reactor. Kinetic data were obtained
under conditions in which intrinsic kinetics prevailed and the
diffusion resistances were suppressed (small catalyst particles, low
catalyst concentrations, high stirring speeds). The influence of
hydrogen pressure and the reaction temperature as well as the pH
response were established. A kinetic model has been proposed
that can describe the experimental data. The model parameters
were established and the model can be used to predict the
behaviour of slurry reactors. Thus the model may be used for
scale-up purposes.

ACKNOWLEDGEMENTS
This work is part of the activities at bo Akademi Process
Chemistry Centre (PCC) within the Finnish Centre of Excellence
Programmes (2000-2005, 2006-2011) by the Academy of Finland.
Mr Sten Lindholm (Laboratory of Analytic Chemistry, bo Akademi

c 2011 Society of Chemical Industry




wileyonlinelibrary.com/jctb

665

The parameter estimation was carried out with the numerical


software ModEst42 by choosing the sum of residual squares,
Equation (13), as the objective function and then minimizing it
with a Simplex algorithm. This resulted in a very good description
of the hydrogenation of the four sugars. The parameters are shown
in Table 3. To better understand the results presented,
Figure 9 compares the hydrogenation of the sugars studied.
From here, it is possible to realize that the disaccharide, indeed,
has the highest activation energy and thus is the molecule that
exhibits a slower rate, while L-rhamnose and L-arabinose, which
present the lowest activation barriers, are the fastest reacting

The estimation results shown in Fig. 10 demonstrate how the


model describes the behaviour of the L-arabinose hydrogenation,
for different hydrogen pressures and temperatures; the modeling
of D-galactose, displayed in Figure 11, was also successful.
The cases of L-rhamnose and D-maltose were somewhat
different owing to the existence of unidentified by-products,
which had some negative impacts on the quality of the model as
shown in Figs 12 and 13, respectively.
As was noted earlier, during the hydrogenation experiments
with D-maltose, it was possible to detect the presence of D-sorbitol.
Consequently, the modeling was carried out as described above
but following the overall reaction route depicted in Fig. 1, which
includes the formation of D-sorbitol through cleavage of the Dmaltitol molecule and its subsequent reduction. The adsorption
equilibrium for D-sorbitol was neglected.

www.soci.org

(a)

L-arabinose (real)

L-arabitol (real)

L-arabinose (model)

L-arabitol (model)

VA Sifontes Herrera et al.

(b)

L-arabinose (real)

L-arabitol (real)

L-arabinose (model)

L-arabitol (model)

10 %
10 %
Concentration [wt.%]

Concentration [wt.%]

8%
8%
6%
4%

6%

4%

2%

2%

0%

0%
0

20

40

60

80 100 120 140 160 180

20

40

60

Time [min]

80 100 120 140 160 180


Time [min]

Figure 10. Modelling results for L-arabinose hydrogenation at different operating conditions: (a) 105 C and 50 bar and (b) 120 C and 60 bar.

D-galactose (real)

D-galactitol (real)

D-galactose (model)

D-galactitol (model)

(b)

4.5 %

4.5 %

4.0 %

4.0 %

3.5 %

3.5 %

Concetration [wt.%]

Concetration [wt.%]

(a)

3.0 %
2.5 %
2.0 %
1.5 %

D-galactitol (real)

D-galactose (model)

D-galactitol (model)

3.0 %
2.5 %
2.0 %
1.5 %

1.0 %

1.0 %

0.5 %

0.5 %

0.0 %

D-galactose (real)

0.0 %
0

20

40

60

80 100 120 140 160 180

20

40

Time [min]

60

80 100 120 140 160 180

Time [min]

Figure 11. Modeling results for D-galactose hydrogenation at different operating conditions: (a) 105 C and 50 bar and (b) 120 C and 60 bar.

(a)

L-rhamnose (real)

L-rhamnitol (real)

L-rhamnose (model)

L-rhamnitol (model)

(b)

L-rhamnose (real)

L-rhamnitol (real)

L-rhamnose (model)

L-rhamnitol (model)

10 %
10 %
Concentration [wt.%]

Concentration [wt.%]

8%
8%
6%
4%
2%

6%

4%

2%
0%
0%
0

20 40 60 80 100 120 140 160 180 200


Time [min]

20

40

60

80 100 120 140 160 180


Time [min]

666

Figure 12. Modeling results for L-rhamnose hydrogenation at different operating conditions: (a) 105 C and 50 bar and (b) 120 C and 60 bar.

wileyonlinelibrary.com/jctb

c 2011 Society of Chemical Industry




J Chem Technol Biotechnol 2011; 86: 658668

Sugar hydrogenation over a Ru/C catalyst

www.soci.org

D-malt. (real)

D-maltit. (real)

D-sorbit. (real)

D-malt. (real)

D-maltit. (real)

D-sorbit. (real)

D-malt. (mod.)

D-maltit. (mod.)

D-sorbit. (mod.)

D-malt. (mod.)

D-maltit. (mod.)

D-sorbit. (mod.)

8%

8%
Concentration [wt.%]

(b) 10 %

Concentration [wt.%]

(a) 10 %

6%

4%

2%

6%

4%

2%

0%

0%
0

20

40

60

80 100 120 140 160 180

20

40

60

Time [min]

80 100 120 140 160 180


Time [min]

Figure 13. Modeling results for D-maltose hydrogenation at different operating conditions: (a) 105 C and 50 bar and (b) 120 C and 60 bar.

University) is gratefully acknowledged for providing his expertise in


the ICP-OES analysis. Mr Linus Silvander (Laboratory of Inorganic
Chemistry, bo Akademi University) is also acknowledged for
his contribution with SEM-EDXA imaging and analysis. Financial
support from the Graduate School in Chemical Engineering to
Vctor A. Sifontes Herrera is gratefully acknowledged.

REFERENCES

J Chem Technol Biotechnol 2011; 86: 658668

c 2011 Society of Chemical Industry




wileyonlinelibrary.com/jctb

667

1 Lichtenthaler F and Peters S, Carbohydrates as green raw materials for


the chemical industry. C R Chim 7:6590 (2004).
2 Lichtenthaler F, Towards improving the utility of ketoses as organic
raw materials. Carbohydr Res 313:6989 (1998).
3 Corma A, Iborra S and Velty A, Chemical routes for the transformation
of biomass into chemicals. Chem Rev 107:24112502 (2007).
4 Dechamp N, Gamez A, Perrard A and Gallezot P, Kinetics of glucose
hydrogenation in a trickle-bed reactor. CatalToday 24:2934 (1995).
5 Cunningham ML and Walker CB, Maltitol-rich solutions and their
manufacture. Patent USA 20040224058 (2004).
6 Niimi M, Hario Y, Ishii Y, Kataura K and Kato K, Preparation of maltose
and maltitol from starch. Patent Japan 02042997 (1990).
7 Darsow G, Preparation of epimer-free sugar alcohols. European Patent
423525 (1991).
8 Magara M, Shimazu K, Fuse M, Kataura K, Osada J, Kato K, et al,
Manufacture of maltitol with high purity. Patent Japan
01093597 (1989).
9 Kondo J, Miyamoto T and Asano M, Maltitol from maltose. Patent
Japan 53119812 (1978).
10 Maisin A, Lefevre A, Wauters M and Germain A, Reactivation of a
catalyst containing platinum group metals for hydrogenation of
sugars. Patent Belgium 882279 (1980).
11 Vanoppen D, Maas-Brunner, M, Kammel Ulrich and Arndt J-D,
Preparation and use of silicon dioxide-supported ruthenium
catalysts for saccharide hydrogenation. Patent Germany
10128205 (2002).
12 Benessere V, Del Litto R, De Roma A and Ruffo F, Carbohydrates as
building blocks of privileged ligands. Coord Chem Rev 254:390401
(2010).
13 Helmchen G and Murmann C, Preparation of optically active diphosphine ligands as asymmetric hydrogenation cocatalyst. European
Patent 885897 (1998).
14 Duffy V and Sigman-Grant M, Position of the american dietetic association: use of nutritive and nonnutritive sweeteners. J Am Dietetic
Assoc 104:(2004).
15 Mitchell H, Sweeteners and Sugar Alternatives in Food Technology.
Blackwell Publishers, Oxford and Ames, Iowa (2006).
16 Eisenbeis C, Guettel R, Kunz U and Turek T, Monolith loop reactor for
hydrogenation of glucose. Catal Today 147S:S342S346 (2009).

17 Wisniak J, Hershkowitz M and Stein S, Hydrogenation of xylose over


platinum group catalysts. Ind Eng Chem Prod Res Dev 13:232236
(1974).
18 Wisniak J, Hershkow M, Leibowit R and Stein S, Hydrogenation of
xylose to xylitol. Ind Eng Chem Prod Res Dev 13:7579 (1974).
19 Heinen A, Peters J and van Bekkum H, Hydrogenation of fructose on
Ru/C catalysts. Carbohydr Res 328:449457 (2000).
20 Makkee M, Kieboom A and van Bekkum H, Hydrogenation of Dfructose and D-fructose D-glucose mixtures. Carbohydr Res
138:225236 (1985).
21 Chen B, Dingerdissen U, Krauter J, Rotgerink H, Mobus K, Ostgard D,
et al, New developments in hydrogenation catalysis particularly
in synthesis of fine and intermediate chemicals. Appl Catal A
280:1746 (2005).
22 Kuusisto J, Mikkola J-P, Sparv M, Warn J, Heikkila H, Perala R, et al,
Hydrogenation of lactose over sponge nickel catalysts kinetics
and modeling. Ind Eng Chem Res 45:59005910 (2006).

23 Mikkola J-P, Sjoholm


R, Salmi T and Maki-Arvela P, Xylose hydrogenation: kinetic and NMR studies of the reaction mechanisms, in
European Symposium on Catalysis in Multiphase Reactors, Toulouse,
France. Elsevier Science Bv, pp. 7381 (1998).
24 Crezee E, Hoffer B, Berger R, Makkee M, Kapteijn F and Moulijn J,
Three-phase hydrogenation of D-glucose over a carbon supported
ruthenium catalyst-mass transfer and kinetics. Appl Catal A
251:117 (2003).
25 Wisniak J and Simon R, Hydrogenation of glucose, fructose, and their
mixtures. Ind Eng Chem Prod Res Dev 18:5057 (1979).
26 Salmi T, Kuusisto J, Warn J and Mikkola J-P, Alternative sweeteners:
la dolce vita. Chim Ind 88:9096 (2006).
27 Salmi T, Murzin D, Warna J, Mikkola J, Aumo J and Kuusisto J, Modeling and optimization of complex three-phase hydrogenations
and isomerizations under mass-transfer limitation and catalyst
deactivation, in Catalysis of Organic Reactions, ed. by Schmidt SR.
CRC Press, Florida, USA pp. 187196 (2007).
28 Mikkola J, Salmi T and Sjoholm R, Effects of solvent polarity on the
hydrogenation of xylose. J Chem Technol Biotechnol 76:90100
(2001).
29 Simms P, Hicks K, Haines R, Hotchkiss A and Osman S, Separation
of lactose, lactobionic acid and lactobionolactone by highperformance liquid-chromatography. J Chromatogr A 667:6773
(1994).
30 Brahme PH and Doraiswamy LK, Modeling of a slurry reaction hydrogenation of glucose on raney-nickel. Ind Eng Chem
Process Des Dev 15:130137 (1976).
31 Kuusisto J, Mikkola J, Sparv M, Warn J, Karhu H and Salmi T, Kinetics
of the catalytic hydrogenation of D-lactose on a carbon supported
ruthenium catalyst. Chem Eng J 139:6977 (2008).
32 Kuusisto J, Mikkola J, Casal P, Karhu H, Varyrynen J and Salmi T, Kinetics
of the catalytic hydrogenation of D-fructose over a CuO-ZnO
catalyst. Chem Eng J 115:93102 (2005).

www.soci.org
33 Nishimura S, Handbook of Heterogeneous Catalytic Hydrogenation for
Organic Synthesis. John Wiley, New York (2001).
34 Albert R, Stratz A and Vollheim G, Catalysis of organic reactions, in
Chemical Industries, ed. by Moser WR. Marcel Dekker, New York
(1981).
35 Ruddlesden JF, Stewart A, Thompson DJ and Whelan R, Diastereoselective control in ketose hydrogenations with supported copper
and nickel-catalysts. Faraday Discuss 72:397411 (1981).

36 Ostgard DJ, Duprez V, Berweiler M, Roder


S and Tacke T, Science
and Technology in Catalysis 2006: Proceedings of the 5th Tokyo
Conference on Advanced Catalytic Science and Ttechnology, Tokyo,
July 2328, 2006, in Studies in Surface and Catalysis, ed. by Eguchi K,
Machida M and Yamanaka I. Elsevier Science Ltd, Amsterdam (2007).
37 Collins P and Ferrier RJ, Monosaccharides: Their Chemistry and Their
Roles in Natural Products. Wiley & Sons, Chichester and New York
(1995).

VA Sifontes Herrera et al.


38 Turek F, Chakrabarti R, Lange R, Geike R and Flock W, On the experimental-study and scale-up of 3-phase catalytic reactors hydrogenation of glucose on nickel-catalyst. Chem Eng Sci 38:275283
(1983).

39 Mikkola J, Salmi T and Sjoholm


R, Modelling of kinetics and mass
transfer in the hydrogenation of xylose over Raney nickel catalyst.
J Chem Technol Biotechnol 74:655662 (1999).
40 Castoldi M, Camara L, Monteiro R, Constantino A, Camacho L,
Carneiro J, et al, Experimental and theoretical studies on glucose
hydrogenation to produce sorbitol. React Kinet Catal Lett
91:341352 (2007).
41 Bernas H, Taskinen A, Warn J and Murzin D, Describing the inverse
dependence of hydrogen pressure by multi-site adsorption of the
reactant: hydrogenolysis of hydroxymatairesinol on a Pd/C catalyst.
J Mol Catal A: Chem 306:3339 (2009).
42 Haario H, ModEst 6.0 A Users Guide. ProfMath, Helsinki (2001).

668
wileyonlinelibrary.com/jctb

c 2011 Society of Chemical Industry




J Chem Technol Biotechnol 2011; 86: 658668

You might also like