Anda di halaman 1dari 6

3574 Ind. Eng. Chem. Res.

2007, 46, 3574-3579

Hydrogen Production from Glucose Using Ru/Al2O3 Catalyst in Supercritical


Water
Adam J. Byrd, K. K. Pant,, and Ram B. Gupta*,
Department of Chemical Engineering, Auburn UniVersity, Auburn, Alabama 36849-5127, and Department of
Chemical Engineering, Indian Institute of Technology, Delhi, India 110016

Glucose, as a model biomass compound, was catalytically reformed in supercritical water to produce hydrogen.
The reforming experiments were conducted in a continuous tubular reactor with and without Ru/Al2O3 catalyst
at short residence times. The addition of catalyst significantly enhanced the overall conversion and hydrogen
yield, and reduced methane formation. The gaseous products contained mainly hydrogen, carbon dioxide,
methane, and a small amount of carbon monoxide. The effects of experimental conditions such as temperature,
reaction time, and concentration of glucose in the feed on formation of hydrogen product were investigated.
Experimental hydrogen yields as high as 12 mol of H2/mol of glucose were obtained, which is the stoichiometric
limit. The gas yield was sensitive to temperature, residence time, and feed concentration. High yield of H2
with low CO and CH4 yields were obtained at high reaction temperature and low glucose concentrations. Tar
formation was observed at high glucose concentrations (>5 wt %). The catalytic conversion of glucose with
ruthenium catalyst in supercritical water is an effective method for hydrogen production directly at a high
pressure, which can be extended to other biomass materials. A reaction mechanism for catalytic reforming in
supercritical water is also discussed.

Introduction presence of a metal catalyst (platinum, nickel, etc.) or an alkali


catalyst.17-27 Cortright et al.17 conducted glucose gasification
Due to the environmental impact of burning fossil fuels,
at 265 C and 56 bar in water, using a platinum catalyst, and
hydrogen is now considered an excellent replacement. Hydrogen
obtained high yields of CO2 and H2. Minowa et al.19 investigated
produced from biomass leads to low carbon dioxide emission
because CO2 produced from gasification is fixed by photosyn- cellulose gasification in near-critical water at 350 C and 165
thesis during biomass growth. Thermochemical gasification of bar with a reduced nickel catalyst and reported that 70% of the
biomass has been identified as a possible way to produce carbon could be gasified.
renewable hydrogen. Several researchers have investigated One of the main problems encountered during the biomass
hydrogen production from methanol, ethanol, and biomass gasification is reactor plugging. Yu et al.13 reported that the
materials in subcritical and supercritical water.1-8 Compared gasification of biomass is effective only at low concentrations
to other biomass thermochemical reforming processes, super- of biomass, as at high concentrations polymerization of the
critical water reforming has a high gasification efficiency and decomposition products occurs. Kruse et al.8,10,11 observed that
operates at a lower temperature.9 Supercritical water (SCW) the hydrogen yield can be significantly increased by adding 5
possesses properties very different from that of liquid water at wt % KOH in SCW. Watanabe et al.14,15 observed a 2-fold
ambient conditions. The dielectric constant of SCW is much increase in the gasification efficiency of glucose and cellulose
less than that of ambient water and hydrogen bonding is much in a batch reactor by adding zirconia catalyst. Similarly nickel-
weaker. Therefore SCW behaves like an organic solvent and is based catalysts have also been tested. Elliott et al.22 conducted
completely miscible with organic materials. Thus with SCW it a gasification of p-cresol in water at 350 C and 200 bar using
is possible to conduct reactions with organic compounds in a various types of base and noble catalysts and reported that nickel
single fluid phase which would otherwise occur in a multiphase and ruthenium were active for the reaction. Again using a batch
system under conventional conditions.1,5,10,11 Heterogeneous reactor, Osada et al.23 reported formation of a high amount of
reactions can also be performed in supercritical fluids. The high methane in the presence of ruthenium catalyst during the
diffusivity of SCW can significantly enhance mass transfer. gasification of cellulose and lignin in SCW. In a recent study,
SCW can reduce coke formation on the catalyst as it is a good Lu et al.9 studied the gasification of cellulose in the presence
solvent for the intermediate coke precursors.6 Hence, gasification of metal catalysts including CeO2, Pd/C, and Ru/C in a batch
of biomass in SCW has many advantages including high reactor, and observed that the maximum hydrogen yield was
gasification efficiency and a high yield of hydrogen.1,9 In obtained with Ru/C catalyst.
addition, the product hydrogen is obtained directly at a high Most of the above studies were done in batch mode, in which
pressure; hence further compression is not needed. Several the biomass/water/catalyst is loaded in a small steel tubular
attempts have been made in recent years for the gasification of
reactor and then sealed and placed in an oven. After the reaction,
biomass model compounds and biomass in supercritical water.8-16
the mixture is quenched and analyzed. Typical reaction time
Cellulose, glucose, and phenolic compounds (e.g., phenol,
varied from minutes to hours. In our recent work on methanol
guaiacol) can be gasified in subcritical water and SCW in the
reforming, it was observed that high reaction time leads to the
secondary reaction of methane formation.1 To limit the methane
* To whom correspondence should be addressed. Tel.: (334) 844-
2013. Fax: (334) 844-2063. E-mail: Gupta@auburn.edu. formation, the reaction time needs to be limited to the order of

Auburn University. seconds. This work examines the reforming of glucose in SCW
in a continuous reactor with a short reaction time (on the order
Indian Institute of Technology.

10.1021/ie070241g CCC: $37.00 2007 American Chemical Society


Published on Web 04/24/2007
Ind. Eng. Chem. Res., Vol. 46, No. 11, 2007 3575

Figure 1. Schematic diagram of the experimental apparatus.

of seconds). Ru/Al2O3 was selected as a catalyst. The effects (Omega FMA-1600). A six-port injection valve (Valco) having
of reaction time, temperature, and feed concentrations were a 100 L sample loop was used for the online sample injection.
studied. The product gas composition was measured using a gas
chromatograph (GC; SRI 8610C) equipped with a thermal
Experimental Section conductivity detector and 60-80 mesh Carboxen-1000 carbon
molecular sieve column (Supelco) having dimensions of 15 ft
Materials. All the chemicals used were of high purity 1/8 in. Nitrogen was used as the carrier gas. The gas
(99.9%) and of analytical grade. Glucose and catalyst (5 wt % chromatograph was calibrated using a standard gas mixture of
Ru supported on amorphous Al2O3) were procured from Aldrich known composition. The liquid product coming out of the phase
and used without any further treatment. The surface area and separator was collected at regular time intervals. The total
pore volume of the catalyst are 100 m2/g and 0.30 cm3/g,
organic carbon (TOC) content of the liquid was analyzed using
respectively. Distilled deionized water was used.
a TOC analyzer (Shimadzu TOC-VCSN). The flow meter (FMA-
Apparatus. The reforming of glucose was carried out in a
1600) was used in the H2 mode, and it generated 30 readings
tubular reactor (0.5 m long, 0.25 in. o.d. and 0.12 in. i.d.) made
of the volumetric flow rate per second based the built-in
of Inconel 600 (Microgroup) having a composition of 73% Ni,
properties of pure H2. These instantaneous volumetric flow rates
18% Cr, and 9% Fe. Initially a few experiments were also
were acquired on a computer via an RS-232 port and corrected
carried out in an empty bed reactor to study the effect of the
for pressure and temperature. The average volumetric flow rate
reactor wall. For the catalytic experiments, a known mass of
was determined by totalizing the flow for a period of 15 min.
the catalyst (2.0 g) was packed into the reactor tube. The catalyst
This average gas flow rate which corresponds to pure H2 was
was retained in the reactor by placing stainless steel frits with
corrected to the actual gas coming out of the phase separator
a pore diameter of 0.5 m (Valco) at either end. The same
charge of catalyst was used for all of the experiments. Glucose by estimating the viscosity of the gas mixture by Wilkes
feedstock was prepared by dissolving a known mass of glucose semiempirical formula28 using the gas composition obtained
in water, and from the feed tank it was pumped to the reactor from the GC analysis. This flow measurement method was
using an HPLC pump (Waters 590). The schematic diagram of checked and confirmed for accuracy by flowing the calibration
the experimental setup is shown in Figure 1. The reactor and gas consisting of H2, CO, CH4, and CO2 of known composition
preheater assembly were placed inside a tubular furnace at several flow rates and measuring the actual flow by a soap-
equipped with a temperature controller (Thermolyne 21100). bubble flow meter. All the experiments were carried out under
The reactor temperature at the exit of the furnace was monitored isothermal conditions.
using a type-K thermocouple. The ends of the tubular furnace Experimental Procedure. First, distilled water was pumped
were covered properly to avoid heat loss and achieve a uniform through the system and pressurized to the desired pressure by
temperature. The gaseous products exiting the reactor were adjusting the back-pressure regulator. After achieving a steady
cooled using a water-cooled double pipe heat exchanger made pressure, the tubular furnace was switched on to heat the reactor.
of SS 316 tubing. The reactor pressure was constantly monitored After a steady exit temperature was achieved, glucose solution
by a pressure gauge. The pressure of the stream was reduced to was introduced into the reactor. The steady-state condition was
ambient by means of a back-pressure regulator (Straval marked by a constant temperature at the exit of the reactor. The
BP00201T-SS). The gas-liquid mixture was separated in a glass gas analysis was done at least three times to get a constant gas
phase separator having gastight valves to prevent the escape of composition. The gas flow was totalized for a period of 15 min,
gases. The gas flow rate was measured using a gas flow meter and the average gas flow rate was calculated using the method
3576 Ind. Eng. Chem. Res., Vol. 46, No. 11, 2007

Table 1. Product Distribution during Thermal (Empty Bed) and


Catalytic Reforming of Glucose in Supercritical Watera
typical product composition (mol %)
H2 CO CH4 CO2 H2 yield
empty bed 54.5 5.6 8.5 31.4 6.5
catalyst (Ru/Al2O3) 68.9 0.1 1.3 29.8 12
a T, 700 C; P, 248 bar; 1 wt % glucose; empty bed residence time, 2 s;

catalytic W/FAo, 15.9 g of catalysts/g-mol of glucose.

discussed before. After completion of each experiment, the feed


was switched back to distilled water to flush the reactor.
All the experiments were conducted at a pressure of 248 bar Figure 2. Hydrogen yield during empty bed reforming of glucose (T,
700 C; P, 248 bar; glucose concentration, 1 wt %).
(3643 psi) and at temperatures ranging from 700 to 800 C.
Experiments were also carried out in an empty bed reactor with
glucose feedstock in supercritical water to study the influence
of the reactor wall. The effect of glucose feed concentration
was studied by feeding glucose having a concentration ranging
between 1 and 5 wt % calculated at the entrance of the reactor.
The residence time in the catalyst bed was varied from between
1 and 6 s.
The molar flow rates of the product gases were calculated
based on the volumetric gas flow rate and dry gas composition
obtained from the GC. The carbon content of the liquid stream Figure 3. Effect of residence time on product gas yields (T, 700 C; P,
was calculated knowing the TOC value of the liquid. All the 248 bar; glucose concentration, 1 wt %; 2.0 g of Ru/Al2O3 catalyst).
measurements were made in triplicate. The accuracy of the run
was checked by calculating the overall carbon balance for the Al2O3 catalyst) of glucose, the product gas contained ap-
system. The error in the overall carbon balance was found to proximately 54 mol % hydrogen along with 1.7% CO, 34.1%
be less than 10%. Scattering in the data of the totalized gas CO2, and 10.2% CH4. Due to the presence of catalyst, the
flow rate measured by the flow meter was less than 1%. The maximum mole percent hydrogen obtained in the product
error in the dry gas composition obtained by the GC analysis gas was significantly increased to the experimental value of
was typically less than 2%. The overall error in the calculation 69 mol %. The yield of major products for thermal and catalytic
of the gas yields due to the errors introduced by the individual gasification of glucose is given in Figures 2 and 3, respectively.
analysis techniques and experimental error was found to be less In the catalytic experiments, there was only a negligible amount
than 5%. of CO in the product, and the methane concentration was also
significantly reduced. Here glucose undergoes dehydrogenation
on the metal surface to give adsorbed intermediates before the
Results and Discussion
cleavage of C-C or C-O bonds. Subsequent cleavage of C-C
Effect of Ru/Al2O3 Catalyst. To study the effect of the Ru/ bonds leads to formation of CO and H2. CO reacts with water
Al2O3 catalyst on hydrogen yield for reforming of glucose in to form CO2 and H2 by the water gas shift reaction. Carbon
supercritical water, catalytic experiments were compared to the monoxide was present only in trace amounts, probably because
empty reactor experiments (without adding Ru/Al2O3 catalyst) the Ru catalyst promotes the water gas shift reaction to form
under identical conditions. The gas yield is defined as the moles carbon dioxide and hydrogen from carbon monoxide and water.
of product gases divided by the moles of glucose fed to the The shift reaction is initiated through interaction of CO with
reactor. Typical product distributions are shown in Table 1 for OH-, which are formed by ionic dissociation of supercritical
experiments with and without Ru/Al2O3 catalyst at 700 C with water on the metal surface, and forming the formate ion which
1 wt % glucose feed. The hydrogen yield increased from 7 to then decomposes in to CO2 and hydride anion. The hydride
12 (moles of hydrogen formed per moles of glucose fed) upon anion further interacts with water, forming H2 and OH- by
adding the catalyst. There was also a significant reduction in electron transfer.25
carbon monoxide and methane yields in the presence of the
catalyst. The main products of the reaction were hydrogen, OH- + CO T HCOO- T H- + CO2 (1)
methane, carbon dioxide, and carbon monoxide. Small amounts
of higher molecular weight carbon compounds such as phenols H- + H2O T H2 + OH- (2)
and aldehydes could be seen in the liquid product in a few
experiments as analyzed by HPLC. Typical values of TOC in Effect of Residence Time. Effect of residence time inside a
the liquid product were approximately 150 ppm. Although the catalyst bed was studied by varying the contact time from 1.6
detailed properties of the catalyst, such as catalyst life and strong to 6. 2 s. The residence time was calculated as the void volume
metal-support interaction, have not been reported for Ru/Al2O3 in the catalyst bed (mL) divided by the volumetric flow rate
catalyst, the probable reason for higher gasification performance (mL/s) at experimental conditions. The corresponding space time
is that the intermediate agents formed during glucose decom- (mass of catalyst in bed/molar flow rate of glucose) varied from
position, such as aldehydes and phenols, were gasified. 7.8 to 229 g of catalysts/g-mol of glucose in the feed. The
The gasification of hydrocarbons in supercritical water results are shown in Figures 3 and 4 for 1 and 5 wt % glucose
proceeds via several complex reactions such as pyrolysis, in the feed, respectively. At 700 C and 1 wt % glucose
hydrolysis, steam reforming, water gas shift, and methanation. concentration, increasing residence time from 1 to 4 s increased
The product distribution mainly depends upon the relative extent the hydrogen yield approximately from 11 to 12. The increase
of various reactions. During thermal gasification (without Ru/ was less significant at high glucose concentration. For the empty
Ind. Eng. Chem. Res., Vol. 46, No. 11, 2007 3577

Figure 7. Effect of glucose concentration on product yields (T, 700 C; P,


Figure 4. Effect of residence time on total organic carbon in liquid phase 248 bar; residence time, 2 s, 2 g of Ru/Al2O3 catalyst).
(T, 700 C; P, 248 bar; residence time, 2 s).
showed incomplete conversion of glucose to gaseous products,
as witnessed by a strong hydrocarbon smell in the liquid product,
which was further confirmed by high TOC values. The solubility
of these intermediate tarry compounds in supercritical water
plays a significant role in their ability to be decomposed to CO,
and then further react with water to produce H2 gas.20,21
At low reaction temperatures, H2 and CO2 readily react to
form methane and water;17 hence hydrogen and carbon dioxide
yields increase with increase in temperature while the methane
yield decreases. In the present study, however, the presence of
catalyst combined with low residence times suppresses the
formation of methane even at relatively low reaction temper-
ature. The low carbon monoxide yield indicates that the water
Figure 5. Effect of residence time on product gas yields (T, 700 C; P,
gas shift reaction approaches completion. The high water excess
248 bar; glucose concentration, 5 wt %; 2 g of Ru/Al2O3 catalyst).
leads to a preference for the formation of hydrogen and carbon
dioxide instead of carbon monoxide. In addition, the intermedi-
ate products such as acids and aldehydes which may otherwise
form in the conventional reactions are not found in supercritical
water reforming in appreciable concentrations.24-27 As compared
with a typical temperature required (800-1000 C) in conven-
tional biomass gasification, supercritical water gasification can
be carried out at a lower temperature,9,27 which is consistent
with our observations using Ru/Al2O3 catalyst.
Effect of Glucose Concentration. Figure 7 shows the effect
of glucose feed concentration on product gas yields at 700 C.
The hydrogen yield dropped by 17% as the glucose concentra-
tion in feed was increased from 1 to 5 wt %. Also, a decrease
in the carbon dioxide yield and a small increase in methane
and carbon monoxide yield were observed on increasing glucose
concentration. The trends of experimental yield of hydrogen,
Figure 6. Effect of temperature on product yields (P, 248 bar; glucose carbon dioxide, and carbon monoxide are similar to those
concentration, 4 wt %; residence time, 2 s; 2 g of Ru/Al2O3 catalyst). reported by other researchers.29-31 Further increase in the
glucose concentration results in formation of heavier molecular
bed experiments, higher residence time for reaction is required
weight hydrocarbons and coke, which often plugs the reactor.
to achieve complete conversion (Figure 2). The effect of
However, the heavy molecules and coke can be reduced by
increasing residence time on total organic carbon in the liquid
increasing temperature in SCW.
product for catalytic and empty bed experiments is shown in
Reaction Mechanism. As discussed earlier, biomass gasifi-
Figure 5. Increasing the residence time in the empty bed runs
cation proceeds via several complex reactions, such as pyrolysis,
significantly reduced the total organic carbon in the liquid,
hydrolysis, steam reforming, water gas shift reaction, and
indicating that these were further converted to CO2 and H2 at
methanation.17.21,31 For glucose reforming in supercritical water,
higher residence times. In the catalytic experiments, the low
a maximum theoretical yield of hydrogen can be calculated by
TOC values correspond to near-complete conversion of carbon
converting all the feed carbon to carbon dioxide with water as
to gaseous products. It should be noted that the residence times
in these experiments are much lower than those in the previous C6H12O6 + 6H2O f 6CO2 + 12H2 (3)
batch experiments.23
Effect of Temperature. Reactor temperature influences the In the overall reaction scheme, CO2 also undergoes hydrogena-
gasification rate and hydrogen formation in biomass reactions. tion reaction to form CO and CH4 as
To study the influence, experiments were carried out at
temperatures ranging between 700 and 800 C. The residence CO2 + H2 T CO + H2O (4)
time inside the catalyst bed was 2 s, while the glucose feed
concentration was 4 wt %. As shown in Figure 6, hydrogen and
yield was near the theoretical maximum of 12 mol of H2/mol
of glucose fed. Additional experiments carried out at 650 C CO + 3H2 T CH4 + H2O (5)
3578 Ind. Eng. Chem. Res., Vol. 46, No. 11, 2007

However, mechanistically glucose reforming via reaction in- The above equation was solved using a nonlinear regression
termediates can be shown as follows:25 technique on the rate of hydrogen production at different glucose
concentrations at a constant pressure and temperature. The value
glucose f acids/aldehydes f gases (6) of rate constant (kR) at 700 C was calculated to be 0.33 (
0.03 s-1 with 95% confidence level. The data could not be
With the presence of ruthenium catalyst in supercritical water,
compared as no kinetic results have been reported on catalytic
extremely high gasification rate and a high hydrogen yield were
supercritical water-glucose reforming in a tubular reactor.
obtained. Also, negligible coke was formed on the catalyst
However, the lowest rate of hydrogen production in the present
surface for up to 4 wt % glucose feed, which indicates that all
study is more than 4 104 mol/gh, which is significantly
the solid products such as chars and aldehydes are decomposed
higher than the maximum value of hydrogen production rate
by the supercritical water in the presence of catalyst. The
reported for glucose reforming by enzymatic route (7 102
dominant reactions in the gas phase are those for water gas shift
mol/gh) and also comparable with the values reported for
and methanation reaction.
catalytic reforming of glucose over Pt/Al2O3 catalyst (3 103
The presence of methane and some traces of liquid hydro-
at 225 C with 10 wt % glucose).17
carbon indicate that reforming of glucose in supercritical water
occurs via the above reaction intermediates. Reforming of these
intermediates by water gas shift reaction is highly likely since Conclusions
water is in high excess. Although the exact role of ruthenium The supercritical water reforming of biomass materials has
is not established in supercritical water reforming, it may be been found to be an effective technique for the production of
assumed that some complex is formed as a result of the reaction high-pressure hydrogen with short residence times. The presence
between adsorbed glucose and supercritical water. The formation of 5 wt % Ru/Al2O3 catalyst significantly enhanced the
of adsorbed intermediates by the decomposition of this complex conversion and hydrogen yield from glucose, while significantly
is assumed to be the rate-determining step. The adsorbed reducing the coke and other heavier liquid product formation.
intermediates further react with water in the gas phase to give The product gases mainly consisted of hydrogen and carbon
CO2 and H2. Assuming S denotes an active site on the catalyst dioxide, with a small amount of methane and carbon monoxide.
and A and B denote glucose and water, respectively, the Hydrogen yields approaching the stoichiometric limit of 12 mol
overall reaction steps may be represented as follows:32 of H2/mol of glucose could be achieved in the present study.
k1,k-1 Hydrogen production is influenced by temperature, residence
A + S 798 AS (7) time, and glucose concentration. The high yield of hydrogen
was obtained at high reactor temperature, low residence time,
k2
AS + B 98 ABS (8) and low glucose concentration. There were no heavy liquid
products (tar) in the liquid effluent and negligible loss in
k3 k4 catalytic activity for up to 5 wt % glucose in the feed. However,
ABS 98 intermediates 98 CO2 + H2 (9) the hydrogen yield decreased as the concentration of glucose
in the feed increased beyond 5 wt %.
where eq 7 denotes the reversible adsorption of glucose on the
catalyst surface, and eq 8 represents the reaction of adsorbed
glucose molecule with water to form a complex molecule ABS. Acknowledgment
Assuming the steady-state hypothesis for the intermediate This research was partially supported by the U.S. Department
complex ABS and AS, the dependence of the rate (r) on reactant of Energy through the Office of Fossil Energy (FE), National
concentration (c) can be expressed as Energy Technology Laboratory (NETL), under DOE Contract
No. DE-FC26-05424.56 via the Consortium of Fossil Fuel
k1k2cAcB Science.
r) (10)
k-1 + k1cA + k2cB + (k1k2cAcB/k3)
Literature Cited
Since in our experiments, water concentration is significantly
(1) Gadhe, J. B.; Gupta, R. B. Hydrogen production by methanol
high, cB is assumed constant and the above expression is reforming in supercritical water: suppression of methane formation. Ind.
simplified as Eng. Chem. Res. 2005, 44, 4577-4585.
(2) Sanjay, P.; Pant, K. K. Production of hydrogen with low carbon
kRcA monoxide formation via catalytic steam reforming of methanol. J. Fuel Cell
r) (11)
1 + bcA Sci. Technol. 2006, 3, 369.
(3) Matsumura, Y.; Xu, X.; Antal, M. J. Gasification Characteristics of
an Activated Carbon in Supercritical Water. Carbon 1997, 35, 819.
where kR and b are lumped parameters defined as (4) Sahoo, D. R.; Vajpai, S.; Patel, S.; Pant, K. K. Kinetic modeling of
steam reforming of ethanol for the production of hydrogen over Co/Al2O3
k1k2cBo catalyst. Chem. Eng. J. 2007, 125, 139-147.
kR ) (12)
k-1 + bcBo (5) Xu, X.; Matsumura, Y.; Stenberg, J.; Antal, M. J., Jr. Carbon-
catalyzed gasification of organic feedstock in supercritical water. Ind. Eng.
Chem. Res. 1996, 35, 2522.
and (6) Williams, P. T.; Onwudili, J. Subcritical and supercritical water
gasification of cellulose, starch, glucose, and biomass waste. Energy Fuels
k1 + (k1k2cBo/k3) 2006, 20, 1259.
b) (13)
k-1 + k2cBo (7) Antal, M. J., Jr.; Allen, S. G.; Schulman, D.; Xu, X. D.; Divilio, R.
J. Biomass gasification in supercritical water. Ind. Eng. Chem. Res. 2000,
39, 4040.
When (bcA) , 1, eq 11 reduces to (8) Kruse, A.; Meier, D.; Rimbrecht, P.; Schacht, M. Gasification of
pyrocatechol in supercritical water in the presence of potassium hydroxide.
r ) kRcA (14) Ind. Eng. Chem. Res. 2000, 39, 4842.
Ind. Eng. Chem. Res., Vol. 46, No. 11, 2007 3579

(9) Lu, Y. J.; Guo, L. J.; Ji, C. M.; Zhang, X. M.; Hao, X. H.; Yan, Q. (22) Elliott, D. C.; Sealock, L. J.; Backer, E. G. Chemical processing
H. Hydrogen production by biomass gasification in supercritical water: A in high pressure aqueous environment: development of catalyst for
parametric study. Int. J. Hydrogen Energy 2006, 31, 822-831. gasification. Ind. Eng. Chem. Res. 1993, 32, 1542-1548.
(10) Kruse, A.; Gawlik, A. Biomass conversion in water at 330-410 C (23) Osada, M.; Sato, T.; Watanabe, M.; Adschiri, T.; Arai, K. Low
and 30-50 MPa. Identification of key compounds for indicating different temperature catalytic gasification of lignin and cellulose with a ruthenium
chemical reaction pathways. Ind. Eng. Chem. Res. 2003, 42, 267. catalyst in supercritical water. Energy Fuels 2004, 18, 327-333.
(11) Kruse, A.; Henningsen, T.; Sinag, A.; Pfeiffer, J. Biomass gasifica- (24) Saisu, M.; Sato, T.; Watanabe, M.; Adschiri, T.; Arai, K. Conversion
tion in supercritical water: Influence of the dry matter content and the of lignin with supercritical water-phenol mixtures. Energy Fuels 2003, 17,
formation of phenol. Ind. Eng. Chem. Res. 2003, 42, 3711. 922.
(12) Yoshida, T.; Oshima, Y. Partial Oxidative and catalytic biomass (25) Penninger, J. M. L.; Rep, M. Reforming of aqueous wood pyrolysis
gasification in supercritical water: A promising Flow reactor system. Ind. condensate in supercritical water. Int. J. Hydrogen Energy 2006, 31, 1597-
Eng. Chem. Res. 2004, 43, 4097. 1606.
(13) Yu, D.; Aihara, M.; Antal, M. J., Jr. Hydrogen production by steam (26) Courson, C.; Makaga, E.; Petit, C.; Kiennemann, A. Development
reforming glucose in supercritical water. Energy Fuels 1993, 7, 574. of Ni catalysts for gas production from biomass gasification. Reactivity in
(14) Watanabe, M.; Inomata, H.; Arai, K. Catalytic hydrogen generation steam- and dry-reforming. Catal. Today 2000, 63, 427.
from biomass (glucose and cellulose) with ZrO2 in supercritical water. (27) McKendry, P. Energy production form bio mass: Conversion
Biomass Bioenergy 2002, 22, 405. technologies (Part 2). Bioresour. Technol. 2002, 83, 47-54.
(15) Watanabe, M.; Inomata, H.; Osada, M.; Sato, T.; Adschiri, T.; Arai, (28) Wilke, C. R. A viscosity equation for gas mixtures. J. Chem. Phys.
K. Catalytic effects of NaOH and ZrO2 for partial oxidative gasification of 1950, 18, 517-19.
n-hexadecane and lignin in supercritical water. Fuel 2003, 82, 545. (29) Sinag, A.; Kruse, A.; Schwarzkopf, V. Key compounds of the
(16) Hologate, H. R.; Meyer, J. C.; Tester, W. J. Glucose hydrolysis hydropyrolysis of glucose in supercritical water in the presence of K2CO3.
and oxidation in supercritical water. AIChE J. 1995, 41, 637. Ind. Eng. Chem. Res. 2003, 42, 3516.
(17) Cortright, R. D.; Davda, R. R.; Dumesic, J. A. Hydrogen from (30) Yan, Q.; Guo, L.; Lu, Y. Thermodynamic analysis of hydrogen
catalytic reforming of biomass-derived hydrocarbons in liquid water. Nature production from biomass gasification in supercritical water. Energy ConVers.
2002, 418, 964-967. Manage. 2006, 47, 1515-1528.
(18) Minowa, T.; Ogi, T.; Yokoyama, S. Hydrogen production from (31) Hashaikeh, R.; Butler, S.; Kozinski, J. A. Selective promotion of
wet cellulose by low temperature gasification using a reduced nickel catalyst. catalytic reactions during biomass gasification to hydrogen. Energy Fuels
Chem. Lett. 1995, 286, 937. 2006, 20, 2743-2747.
(19) Minowa, T.; Ogi, T.; Yokoyama, S. Effect of pressure on low- (32) Vaidya, P. V.; Rodrigues, A. E. Kinetics of steam reforming of
temperature gasification of wet cellulose into methane using reduced nickel ethanol over a Ru/Al2O3 catalyst. Ind. Eng. Chem. Res. 2006, 45, 6614-
catalyst and sodium carbonate. Chem. Lett. 1995, 280, 285. 6618.
(20) Yoshida, T.; Matsumura, Y. Gasification of Cellulose, Xylan, and
Lignin Mixtures in Supercritical Water. Ind. Eng. Chem. Res. 2001, 40, ReceiVed for reView February 13, 2007
5469. ReVised manuscript receiVed March 16, 2007
(21) Yoshida, T.; Oshima, Y.; Matsumura, Y. Gasification of biomass Accepted March 22, 2007
model compounds and real biomass in supercritical water. Biomass
Bioenergy 2004, 26, 71. IE070241G

Anda mungkin juga menyukai