You are on page 1of 9

1018 Ind. Eng. Chem. Res.

2010, 49, 1018–1026

Kinetics of Hydrogen Uptake and Release from Heteroaromatic Compounds for


Hydrogen Storage
Farnaz Sotoodeh and Kevin J. Smith*
Department of Chemical & Biological Engineering, UniVersity of British Columbia, 2360 East Mall,
VancouVer, BC, Canada V6T 1Z3

The kinetics of hydrogen uptake by N-ethylcarbazole and carbazole, and the kinetics of H2 release from their
hydrogenated products, is reported. The hydrogenation of N-ethylcarbazole at 130-150 °C on Ru was well
described by a network of first-order stepwise hydrogenation reactions. The hydrogenation of N-ethylcarbazole
was significantly faster than the hydrogenation of carbazole, and in both cases, >95% selectivity to the
completely hydrogenated products, dodecahydro-N-ethylcarbazole and dodecahydrocarbazole, was achieved.
The dehydrogenation of dodecahydro-N-ethylcarbazole at 101 kPa and 150-170 °C, proceeded to 100%
conversion of the reactant over a Pd catalyst within 1 h. However, only 69% of the stored H2 was recovered
due to a low selectivity to N-ethylcarbazole. In the case of dodecahydrocarbazole dehydrogenation at 101
kPa and 170 °C over the same Pd catalyst, 28% H2 recovery was obtained due to the strong adsorption of the
product carbazole on the catalyst surface that also resulted in a slower rate of dodecahydrocarbazole
dehydrogenation compared to dodecahydro-N-ethylcarbazole.

1. Introduction present work, the complete hydrogenation-dehydrogenation


cycle needed for hydrogen storage and release has been
One approach to hydrogen storage for fuel-cell vehicles is to
investigated for both carbazole and N-ethylcarbazole. The
use organic compounds that have a high capacity to bind
kinetics of the hydrogenation reactions needed for hydrogen
hydrogen covalently. Storage and release of the hydrogen is
storage on N-ethylcarbazole and carbazole, and the kinetics of
achieved by hydrogenation and dehydrogenation of the organic
the dehydrogenation reactions needed for hydrogen release from
compounds, and these reactions are done in the presence of a
the corresponding hydrogenation products, are reported. N-
catalyst. For transportation applications, the H2 storage (hydro-
Ethylcarbazole was hydrogenated over a supported Ru catalyst
genation) would be done off-board the vehicle. Also, the
in an autoclave batch reactor at 7 MPa and 130-150 °C. The
candidate compounds must have high hydrogen release rates,
hydrogenated compound, dodecahydro-N-ethylcarbazole, was
more than 5.5 wt % H2 storage capacity (calculated as the mass
then dehydrogenated over a supported Pd catalyst at 150-170
of H2 liberated divided by the mass of hydrogenated compound),
°C and 101 kPa. Kinetic models of the hydrogenation and
melting points below -40 °C and boiling points above 200 °C.1,2
dehydrogenation reactions, as well as the activation energies
Organic heteroaromatic compounds such as carbazole have
attracted attention because they meet the storage capacity targets of the reactions, are reported. For comparison, carbazole was
and the boiling point criteria. hydrogenated at 7 MPa and 170 °C, using the same catalyst
Although some results have been reported for H2 storage on and experimental setup used for the hydrogenation of N-
heteroaromatic compounds, limited data are available on the ethylcarbazole. The hydrogenated product, dodecahydrocarba-
kinetics of the reactions involved in H2 uptake and release.3-7 zole, was subsequently dehydrogenated at 170 °C and 101 kPa
Carbazole and N-ethylcarbazole, when hydrogenated to dodecahy- over a supported Pd catalyst.
drocarbazole and dodecahydro-N-ethylcarbazole, have hydrogen
storage capacities of 6.7 and 5.8 wt %,5,8,9 respectively. They 2. Experimental Section
are, therefore, potential candidates for hydrogen storage ap- 2.1. Catalyst Preparation and Characterization. A 5 wt
plications. However, the hydrogenation-dehydrogenation kinet- % Pd on silica catalyst was prepared by wet impregnation and
ics of these compounds are not well established, especially at used as the dehydrogenation catalyst. Details of the catalyst
the low temperature (<200 °C) conditions required for on-board preparation method have been reported previously by the
dehydrogenation reactions. In addition, although platinum group authors.10 A commercial 5 wt % Ru on alumina catalyst
metals are widely used for hydrogenation and dehydrogenation (Aldrich, 5 wt % Ru, reduced), received in a reduced state, was
reactions, the influence of the heteroatom present in these used as the hydrogenation catalyst and was rereduced prior to
compounds on the reaction kinetics needs to be understood, use. Both catalysts were characterized to determine BET surface
given that strong heteroatom adsorption can lead to a loss in area, pore volume, and pore size using a Micromeritics ASAP
catalyst activity. The N heteroatom associated with N-ethylcar- 2020 analyzer. A Micromeritics Autochem II 2920 was used
bazole and dodecahydro-N-ethylcarbazole is bound to an ethyl for temperature programmed reduction (TPR) analysis of the
group, whereas in the corresponding carbazole compounds, the catalysts and to determine the metal dispersion by CO pulse
N heteroatom is free to interact with the catalyst surface through chemisorption.10
the lone pair of electrons associated with the N. 2.2. Catalyst Activity. The hydrogenation of N-ethylcarba-
In previous work, hydrogen recovery from dodecahydro-N-
zole was carried out at 130, 140, and 150 °C and a H2 pressure
ethylcarbazole was demonstrated over a Pd catalyst.10 In the
of 7 MPa, in a 300 cm3 autoclave batch reactor with continuous
* To whom correspondence should be addressed. E-mail: kjs@ monitoring and control of stirrer speed, temperature, and
interchange.ubc.ca. pressure. Approximately 100 cm3 of a 6 wt % solution of
10.1021/ie9007002  2010 American Chemical Society
Published on Web 09/16/2009
Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010 1019
Table 1. Catalyst Properties
catalyst BET, m2/g pore volume, cm3/g average pore size, nm degree of reduction, % metal dispersion, %
Pd/silica (calcined) 311.5 0.9 9.6
Pd/silica (reduced) 325.6 1.1 9.5 78.0 4.2
Ru/alumina (reduced) 76.5 0.2 8.4 3.3a 6.2
a
The commercial catalyst was supplied in a prereduced state.

N-ethylcarbazole (Aldrich, 97%) in decalin (Sigma-Aldrich, stirred, and heated to the desired temperature. Helium, at a flow
98%) and 1 g of the 5 wt % reduced Ru/alumina catalyst, with of 170 cm3(STP) · min-1, was used as a carrier gas for continuous
a particle size <150 µm, were loaded into the reactor. The reactor removal of the produced H2. The exit gas composition was
was first purged in a 55 cm3(STP) · min-1 flow of N2 for about continuously monitored using a quadrupole mass spectrometer
1 h and then heated to the desired temperature while purging (SRC Residual Gas Analyzer, 200 amu) so that the produced
with ultra high purity (UHP) hydrogen at a 55 cm3(STP) · min-1
H2 could be quantified. Liquid samples were also withdrawn
flow and low pressure (<35 kPa) to ensure that no reaction
from the reactor for analysis by gas chromatography using a
occurred. After the desired temperature was reached, the reactor
was pressurized to 7 MPa, corresponding to more than 2.5 times 14-A Shimadzu gas chromatograph equipped with a flame
the amount needed for complete hydrogenation of the N- ionization detector (FID) and an AT-5 25 M × 0.53 mm
ethylcarbazole reactant, while stirring at a speed of 600 rpm. capillary column. Product identities were also confirmed by the
The H2 pressure was monitored continuously during the experi- Shimadzu QP-2010S GC/MS. The product from the hydrogena-
ment. Small volumes (<0.5 cm3) of liquid sample were removed tion of carbazole was also dehydrogenated in the same reactor
from the reactor periodically for analysis by gas chromatogra- at 101 kPa and 170 °C. In this case, 0.17 g of the reduced 5 wt
phy. Hydrogenation of carbazole was carried out using the same % Pd/silica was used as the dehydrogenation catalyst. The same
autoclave reactor at 7 MPa and 150 °C. In this case, a 1 wt % experimental procedure was used for both dehydrogenations.
solution of carbazole (Aldrich, 95%) in decalin (Sigma-Aldrich,
98%) and 1 g of the 5 wt % reduced Ru/alumina catalyst with The absence of internal and external mass transfer effects on
a particle size <150 µm were loaded into the reactor. The same the measured reaction kinetics was confirmed experimentally
start-up and sampling procedure used for the hydrogenation of and by theoretical analysis. For the hydrogenation reactions,
N-ethylcarbazole was used for the hydrogenation of carbazole. the gas-to-liquid mass transfer rates, estimated according to
However, due to the extremely low solubility of carbazole in Zievernek et al.,11 were at least 2 orders of magnitude higher
decalin at room temperature, liquid samples were only taken at than the observed reaction rate. Furthermore, repeating the
the end of the experiment and the identity of the final product reactions at stirrer speeds of 600 and 1000 rpm showed that, in
was confirmed by GC/MS analysis with a Shimadzu QP-2010S
both cases, the measured reaction rates were within (10% of
GC/MS using a Restek RTX5 30 M × 0.25 mm capillary
the average values. Internal mass transfer effects were minimal
column.
The recovered product from the hydrogenation of the N- because of the small catalyst particle size and confirmed by the
ethylcarbazole was subsequently used as a reactant for the calculated Weisz-Prater criterion of Φ < 1.0 at the highest
dehydrogenation experiments. These reactions were carried out reaction temperature. Similarly, both internal and external mass
in a 50 cm3 glass flask reactor using 0.3 g of 5 wt % Pd/silica transfer effects were shown to be minimal during the dehydro-
at 150, 160, and 170 °C and 101 kPa. The reactor was sealed, genation reactions.

Figure 1. Hydrogenation reaction pathway of N-ethylcarbazole obtained from GC/MS analysis: (a) N-ethylcarbazole, (b) tetrahydro-N-ethylcarbazole, (c)
octahydro-N-ethylcarbazole, (d) hexahydro-N-ethylcarbazole, (e) decahydro-N-ethylcarbazole, (f) dodecahydro-N-ethylcarbazole.
1020 Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010
Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010 1021

Figure 2. (A) Product distribution over time for N-ethylcarbazole hydrogenation at 150 °C: (a) N-ethylcarbazole, (b) tetrahydro-N-ethylcarbazole, (c) octahydro-
N-ethylcarbazole, (d) hexahydro-N-ethylcarbazole, (e) decahydro-N-ethylcarbazole, (f) dodecahydro-N-ethylcarbazole. Note the different scale of each plot.
(B) Product distribution over time for N-ethylcarbazole hydrogenation at 140 °C: (a) N-ethylcarbazole, (b) tetrahydro-N-ethylcarbazole, (c) octahydro-N-
ethylcarbazole, (d) hexahydro-N-ethylcarbazole, (e) decahydro-N-ethylcarbazole, (f) dodecahydro-N-ethylcarbazole. Note the different scale of each plot.
(C) Product distribution over time for N-ethylcarbazole hydrogenation at 130 °C: (a) N-ethylcarbazole, (b) tetrahydro-N-ethylcarbazole, (c) octahydro-N-
ethylcarbazole, (d) hexahydro-N-ethylcarbazole, (e) decahydro-N-ethylcarbazole, (f) dodecahydro-N-ethylcarbazole. Note the different scale of each plot.
3. Results and Discussion showed that N-ethylcarbazole hydrogenation proceeded through
the stepwise saturation of double bonds, leading to the parallel
3.1. Catalyst Properties. Table 1 summarizes the charac-
production of tetrahydro-N-ethylcarbazole and octahydro-N-
terization data for both catalysts of the present study. BET
ethylcarbazole as the primary reaction products. Rapid con-
surface area, pore volume, and average pore size were measured
sumption of tetrahydro-N-ethylcarbazole and subsequent pro-
for the 5 wt % Pd/silica after calcination and reduction and for
duction of hexahydro-N-ethylcarbazole suggested that the
the 5 wt % reduced Ru/alumina catalyst.
reaction proceeded to the production of hexahydro-N-ethylcar-
Temperature programmed reduction (TPR) of the calcined
bazole from tetrahydro-N-ethylcarbazole. The product distribu-
Pd/silica catalyst precursor showed a 78% degree of reduction
tion showed slower consumption of hexahydro-N-ethylcarbazole
of PdO to Pd. The commercial Ru/alumina was received in a
reduced state, and the low degree of reduction for this catalyst compared to tetrahydro-N-ethylcarbazole. Octahydro-N-ethyl-
reflects the stability of the reduced Ru in air. CO pulse carbazole, one of the main intermediates, can undergo saturation
chemisorption data were used to determine that the Pd dispersion of the double bonds through two parallel paths to produce
on the reduced Pd/silica catalyst was 4.2%, whereas the Ru decahydro-N-ethylcarbazole, as well as dodecahydro-N-ethyl-
dispersion was 6.2% on the rereduced Ru/alumina catalyst. Even carbazole, the major final product. The reaction completed with
though the BET area of the Pd/silica dehydrogenation catalyst more than 95% selectivity to the desired dodecahydro-N-
was about 4 times that of the Ru/alumina hydrogenation catalyst, ethylcarbazole, and less than 5% selectivity to the product
both catalysts had similar metal dispersions. octahydro-N-ethylcarbazole. The total amount of hydrogen
3.2. Catalyst Activity and Kinetics. Hydrogenation. When consumed during hydrogenation corresponded to approximately
N-ethylcarbazole was hydrogenated at an H2 pressure of 7 MPa 5.3 wt % of the reactant. This is very close to the storage
and temperature of 150 °C, the conversion of N-ethylcarbazole criterion of Department of Energy (DOE) and confirms that
proceeded to 98% after 1 h, with an initial rate of H2 N-ethylcarbazole is a promising compound for hydrogen storage.
consumption of 58.4 mM · min-1 · gcat-1. Analysis of the liquid The experimental product distribution data measured over
samples by GC/MS confirmed the production of partially and time at three reaction temperatures were used to model the
completely hydrogenated compounds during the reaction, as reaction kinetics. The kinetic model based on the reaction
represented in Figure 1. The reaction pathway in Figure 1 was network shown in Figure 1 assumed that reactions were first-
suggested from the product distributions measured over time order in the concentration of the heteroaromatic organic
at three temperatures as shown in Figure 2a-c. These data compounds and zero-order in hydrogen. The parameters (rate
1022 Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010

Table 2. First-Order Rate Constants and Activation Energies for the Hydrogenation of N-Ethylcarbazole over Ru/Alumina Catalyst
temperature, °C k1, h-1 k2, h-1 k3, h-1 k4, h-1 k5, h-1 k6, h-1 k7, h-1
130 0.021 ( 0.012a 0.444 ( 0.018 0.840 ( 0.462 0.006 ( 0.006 1.140 ( 0.060
140 0.042 ( 0.018 1.020 ( 0.018 2.034 ( 0.600 0.606 ( 0.672 0.012 ( 0.012 2.964 ( 0.528 0.048 ( 0.024
150 0.096 ( 0.012 1.764 ( 0.018 5.646 ( 1.188 2.538 ( 0.414 0.024 ( 0.006 3.300 ( 0.024 0.252 ( 0.084
activation 107.6 ( 7.0 97.9 ( 10.3 134.9 ( 7.5 208.0 (R2 ) 1) 98.2 ( 1.4 75.8 ( 33.6 240.9 (R2 ) 1)
energy, kJ/mol (R2 ) 0.99) (R2 ) 0.98) (R2 ) 0.99) (R2 ) 0.99) (R2 ) 0.83)
a
Standard error of estimated parameters.

Table 3. Initial Reaction Rates, Rate Constants, and Turnover Frequencies (TOFs) for Hydrogenation of N-Ethylcarbazole and Carbazole at
150 °C over Ru/Alumina Catalyst
reactant catalyst/reactant ratio, g · g-1 initial rate,a mM · min-1 · gcat-1 rate constant,b k, h-1 TOF,c s-1 reaction time,d h
N-ethylcarbazole 0.17 58.36 2.952 ( 0.122 0.223 1.16
carbazole 1.00 18.44 0.318 ( 0.034 0.108 13.33
a
Reaction rate with respect to H2 consumption. b
With respect to N-ethylcarbazole and carbazole consumption. c
Molecule reactant · s-1 · Ru surface
atom-1. d Reaction time for 100% conversion.

constants) of the kinetic model were estimated by minimizing compared to the other products. The estimated rate constants
an objective function, defined as the sum of squares of the for the conversion of tetrahydro-N-ethylcarbazole (k3) and
difference between the experimental component concentrations hexahydro-N-ethylcarbazole (k4) were consistent with the ob-
and those calculated by the kinetic model, using a weighted served slower consumption of hexahydro-N-ethylcarbazole
least-squares method. The objective function was minimized compared to the more rapid consumption of tetrahydro-N-
using a Nelder-Mead simplex (direct search) method. At each ethylcarbazole in the first 10 min of the reaction (Figure 2a).
iteration on the parameter values, the concentration of each Comparing the apparent activation energies of Table 2, it is
compound as a function of time was calculated by numerical seen that the production of octahydro-N-ethylcarbazole from
integration of the five ordinary differential equations that hexahydro-N-ethylcarbazole (reaction 4) and the production of
described the time rate-of-change of the hydrogenated product dodecahydro-N-ethylcarbazole from decahydro-N-ethylcarbazole
concentrations. In addition, the mole balance constraint was used (reaction 7) have the highest activation energies, in agreement
to calculate the reactant concentration at each time step. with the experimental observation of low consumption rates of
Although reaction orders from 0 to 2 in increments of 0.5 were hexahydro-N-ethylcarbazole and decahydro-N-ethylcarbazole.
examined for all heteroaromatic organic compounds as well as Carbazole was hydrogenated using the same autoclave reactor
hydrogen, the assumed first-order kinetics for reactant and and catalyst to produce the reactant for the dehydrogenation
hydrogenated products and zero-order kinetics for hydrogen gave reaction. At a H2 pressure of 7 MPa and a temperature of 150
the best fit of the experimental data based on a minimum °C, 100% conversion of carbazole was achieved after about 13 h.
magnitude of the residual sum of squares. A zero-order reaction Due to the extremely low solubility of carbazole in decalin,
in hydrogen likely reflects the excess hydrogen at high pressure only a few liquid samples were withdrawn after a reaction time
present during the hydrogenation reactions, and the relatively of 10 h. Liquid sample analysis by GC/MS confirmed the
narrow range of H2 pressures investigated herein, that results production of dodecahydrocarbazole as the main product.
in an approximate constant concentration of H2 in the liquid The reaction proceeded with more than 97% selectivity to the
phase as the reaction proceeded. completely hydrogenated product, dodecahydrocarbazole, after
The activation energies for the reactions were determined by about 13 h. The initial rate of H2 consumption by carbazole
linear regression according to the Arrhenius equation. Individual was 18.44 mM · min-1 · gcat-1 at 150 °C. Since the intermediate
first-order rate constants estimated at 130, 140, and 150 °C, as products could not be monitored as a function of time in this
well as the activation energies are reported in Table 2. The case, the kinetics of carbazole consumption were described by
model results calculated using the obtained rate constants are a first-order rate equation in carbazole concentration and zero-
plotted against the experimental product concentrations in Figure order in H2 concentration. The same mathematical procedure
2a-c. The standard error of the parameter estimates reported used to model the kinetics of N-ethylcarbazole hydrogenation
in Table 2 are considered acceptable given that a simple power was used to model carbazole hydrogenation. Different reaction
law model with fixed exponents has been used to fit the kinetics orders for the organic heteroaromatic compounds as well as
of a complex reaction network over a relatively wide range of hydrogen were examined in the model. The chosen orders are
reaction temperatures. based on the best fit of experimental data.
The model results confirmed that the production of intermedi- A comparison between the initial rates, the first-order rate
ates and the final product were well described by the simplified constants, and turnover frequencies for the consumption of
first-order kinetic model shown in Figure 1. The data for the carbazole and N-ethylcarbazole at 150 °C is presented in Table
concentration of tetrahydro-N-ethylcarbazole showed the largest 3. N-ethylcarbazole hydrogenation showed a higher initial
difference between the measured and calculated concentrations, reaction rate and proceeded to 100% conversion in a signifi-
but the concentration of this compound was the lowest among cantly shorter reaction time with lower metal loading of the
all compounds by approximately 1 order of magnitude. Table catalyst, compared to carbazole. The slower hydrogenation rate
2 summarizes the obtained rate constants for the individual of carbazole compared to N-ethylcarbazole can be attributed to
reactions of the model (Figure 1), as well as the activation the strong adsorption of the carbazole molecule through the N
energies estimated from the rate constants. The rate constants heteroatom to the metal surface. The inhibiting effects of basic
for reaction 4 (k4) and reaction 7 (k7) of Figure 1, corresponding N in organic compounds on precious metal catalysts, due to
to the conversion of hexahydro-N-ethylcarbazole and decahydro- the unshared pair of electrons on N, is well-known.12 The steric
N-ethylcarbazole, were zero at 130 °C, confirming the higher hindrance associated with the ethyl group of N-ethylcarbazole
activation barrier for the consumption of these two intermediates prevents the adsorption of the molecule on the metal surface
Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010 1023

Figure 3. Product distribution over time for dehydrogenation of dodecahydro-N-ethylcarbazole obtained by GC/MS: (a) 150, (b) 160, and (c) 170 °C; (O)
experimental data, (solid line) fitted model. The rate constants are reported in Table 4.

through the N, as has been previously reported for similar trations and those calculated from the kinetic model. Individual
geometries.13 Therefore, although carbazole has a higher first-order rate constants were obtained at 150, 160, and 170
theoretical H2 storage capacity than N-ethylcarbazole, the °C and these are reported in Table 4, together with the apparent
inhibiting effect of the N atom slowed the hydrogenation activation energies of each reaction. The model results calculated
reaction and complete H2 saturation of carbazole was obtained using the obtained rate constants are plotted against the
in a much longer reaction time compared to N-ethylcarbazole, experimental product concentrations in Figure 3. Clearly, the
making carbazole a less practical candidate for H2 storage. simplified kinetic scheme describes the product distribution from
Dehydrogenation. The products from N-ethylcarbazole hy- dodecahydro-N-ethylcarbazole dehydrogenation on Pd/silica
drogenation in decalin at 150 °C, containing octahydro-N- satisfactorily. The kinetics of dodecahydro-N-ethylcarbazole
ethylcarbazole (3%) and dodecahydro-N-ethylcarbazole (97%), consumption was also well described by a first-order rate
were transferred to the 50 cm3 reactor for dehydrogenation at equation, confirmed by testing the model with different reaction
150, 160, and 170 °C over the 5 wt % reduced Pd/silica catalyst. orders, and the corresponding rate constants and activation
The product distribution over time at temperatures of 150, 160, energy estimated from the experimental data are provided in
and 170 °C is presented in Figure 3. Previous studies showed Table 4.10
that decalin dehydrogenation occurs at temperatures above 200 Table 4 summarizes the reaction rates and rate constants, as
°C,14-17 and our own tests confirmed that decalin dehydroge- well as the percent hydrogen recovered at each temperature.
nation did not occur at temperatures below 180 °C on the Pd/ One hundred percent of the reactant, dodecahydro-N-ethylcar-
silica catalyst. Consequently, it was possible to use decalin as bazole, was converted at the three temperatures, resulting in 2
a solvent for the dehydrogenation reactions. The GC/MS wt % H2 recovery after 2 h and 4 wt % H2 recovery after 17 h
analysis of liquid samples confirmed the production of tetrahy- at 170 °C. This is comparable to results reported for the
dro-N-ethylcarbazole and octahydro-N-ethylcarbazole at all three dehydrogenation of another candidate compound, 1,2,3,4-
temperatures, with a relatively slow conversion of octahydro- tetrahydrocarbazole at 140 °C over 5 wt % Pd/Al2O3, where
N-ethylcarbazole to tetrahydro-N-ethylcarbazole. However, no 81% of the reactant was converted after 27 h.18,19 However,
significant amounts of N-ethylcarbazole were produced. The although 100% conversion of dodecahydro-N-ethylcarbazole
experimental product distribution data measured over time at was achieved in the present study, only 69% of the total stored
three temperatures were used to model the dehydrogenation H2 was recovered because the dehydrogenation proceeded to
reaction kinetics based on the reaction network shown in Figure tetrahydro-N-ethylcarbazole and octahydro-N-ethylcarbazole,
4A. The kinetic model assumed first-order reactions in the and not to the completely dehydrogenated product, N-ethylcar-
concentration of dodecahydro-N-ethylcarbazole. A weighted bazole. The lack of selectvity to N-ethylcarbazole is likely
least-squares method was used to minimize the sum of squares related to the adsorption geometries associated with tetrahydro-
of the difference between the experimental component concen- N-ethylcarbazole and octahydro-N-ethylcarbazole.
1024 Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010

Figure 4. Comparison between the reaction pathways of (A) dodecahydro-N-ethylcarbazole dehydrogenation (a) dodecahydro-N-ethylcarbazole, (b) octahydro-
N-ethylcarbazole, (c) tetrahydro-N-ethylcarbazole; (B) dodecahydrocarbazole dehydrogenation (a) dodecahydrocarbazole, (b) octahydrocarbazole, (c)
tetrahydrocarbazole, (d) carbazole.

Table 4. Initial Reaction Rates, Rate Constants, Activation Energies, and H2 Recovery in Dehydrogenation of Dodecahydro-N-ethylcarbazole
over Pd/Silica Catalyst
initial reaction rate, H2 recovery
temperature, °C mM · min-1 · gcat-1 rate constant, k, h-1 rate constant, k1, h-1 rate constant, k2, h-1 after 17 h, wt %
150 3.00 0.096 ( 0.000a 0.072 ( 0.000 0.036 ( 0.000 2.10
160 8.51 0.264 ( 0.000 0.078 ( 0.002 0.162 ( 0.001 3.30
170 12.11 0.450 ( 0.050 0.162 ( 0.036 0.210 ( 0.037 4.00
activation energy, kJ/mol 126.8 ( 24.3 67.1 ( 30.1 144.3 ( 61.0
a
Standard error of mean of estimated parameters.

The product from the carbazole hydrogenation reaction, results, it was confirmed that the production of octahydrocar-
containing isomers of dodecahydrocarbazole, was also loaded bazole and tetrahydrocarbazole were well described by first-
into the 50 cm3 dehydrogenation reactor. Dehydrogenation was order kinetics. Previous studies on the dehydrogenation of
performed at 101 kPa and 170 °C, to provide a comparison with tetrahydrocarbazole10,18,19 indicated that the reaction was zero-
dodecahydro-N-ethylcarbazole dehydrogenation. Figure 4A and order with respect to tetrahydrocarbazole consumption. How-
B compares the suggested reaction pathway for the dehydro- ever, because of the poisoning of Pd by the strong adsorption
genation of dodecahydro-N-ethylcarbazole and dodecahydro- of the product carbazole on the Pd surface, the reaction rate
carbazole, respectively, obtained from the experimental product decreased, as shown in Figure 5d, where the dashed line shows
distribution data over time. Dodecahydrocarbazole conversion the predicted zero-order carbazole production and the experi-
proceeded through two parallel reactions leading to the produc- mental data show a decrease in reaction rate because of product
tion of octahydrocarbazole and tetrahydrocarbazole. Production catalyst poisoning. The first-order rate constants for the produc-
of small amounts of carbazole, after nearly 4 h, was confirmed tion of octahydrocarbazole and tetrahydrocarbazole, and the
by analyzing the liquid samples as well as by quantifying the zero-order rate constant for the production of carbazole obtained
produced H2. from the model fit to the experimental data, are summarized in
The experimental data obtained by the analysis of the liquid Table 5.
samples were used to model the dehydrogenation kinetics of The experimental data show that only 1.9 wt % H2 recovery
dodecahydrocarbazole. A first-order rate equation was fitted to was achieved from the dehydrogenation of dodecahydrocarba-
the experimental data obtained for the production of octahy- zole after 17 h, which is smaller than the amount of H2 recovered
drocarbazole and tetrahydrocarbazole, whereas a zero-order rate from the dehydrogenation of dodecahydro-N-ethylcarbazole at
equation was considered for the conversion of tetrahydrocar- 170 °C for the same reaction time. A comparison between the
bazole to the final product, carbazole, as previously reported.10,18,19 dehydrogenation of dodecahydro-N-ethylcarbazole and dodecahy-
Figure 5 illustrates the experimental data for the dehydrogenation drocarbazole at 170 °C is provided in Table 6. This comparison
of carbazole at 170 °C and the fitted model. From the obtained shows that 100% conversion of dodecahydro-N-ethylcarbazole
Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010 1025

Figure 5. Product distribution over time for dehydrogenation of dodecahydrocarbazole obtained by GC/MS at 170 °C: (a) dodecahydrocarbazole, (b)
octahydrocarbazole, (c) tetrahydrocarbazole, (d) carbazole; (O) experimental data, (solid line) fitted model. The dashed line in part d corresponds to the
predicted carbazole concentration without catalyst poisoning. The rate constants are reported in Table 5.

Table 5. Initial Reaction Rate and Rate Constants for Dehydrogenation of Dodecahydrocarbazole at 170 °C
initial reaction rate, mM · min-1 · gcat-1 rate constant,a k, h-1 k1, h-1 k2, h-1 k3, mM · h-1 · gcat-1 H2 recovery, wt %
4.13 0.042 ( 0.000b 0.024 ( 0.000 0.018 ( 0.001 0.0003 ( 0.0000 0.64,c 1.90d
a b c d
First-order rate constant for dodecahydrocarbazole consumption. Standard error of mean of estimated parameters. After 2 h. After 17 h.

Table 6. Comparison between the Initial Reaction Rates, Rate Constants, and H2 Recovery for the Dehydrogenation of
Dodecahydro-N-ethylcarbazole and Dodecahydrocarbazole at 170 °C over 5 wt % Pd/SiO2
reactant initial rate, mM · min-1 · gcat-1 rate constant,a k, h-1 conversion,b % H2 recovery,b,c wt %
dodecahydro-N-ethylcarbazole 12.00 0.450 ( 0.050d 100 4.00
dodecahydrocarbazole 4.13 0.042 ( 0.001 52.6 1.90
a
With respect to reactant consumption. b After 17 h. c
The ratio of H2 liberated in grams to the weight of the starting reactant in grams (×100).
d
Standard error of mean of estimated parameters.

was obtained after 17 h, while only 52.6% of the dodecahy- followed first-order kinetics. Dehydrogenation of the dodecahy-
drocarbazole was converted to the dehydrogenated products dro-N-ethylcarbazole was also faster than the dehydrogenation
during the specified reaction time. The lower reaction rate and of dodecahydrocarbazole over a Pd/silica catalyst at 150-170
rate constant for the dehydrogenation of dodecahydrocarbazole °C. However, only 69% of the stored hydrogen was recovered
compared to dodecahydro-N-ethylcarbazole is likely because of from dodecahydro-N-ethylcarbazole at 170 °C, due to a high
the strong adsorption of the dehydrogenated products to the selectivity to tetrahydro-N-ethylcarbazole and octahydro-N-
metal surface through the N atom. For the case of dodecahydro- ethylcarbazole. H2 recovery from dodecahydrocarbazole was
N-ethylcarbazole, the steric hindrance of the ethyl group prevents 28%. By comparing the obtained results from the dehydroge-
the molecules from being adsorbed through N, and thus, a higher nation reaction, it is concluded that N-ethylcarbazole and its
reaction rate was observed. hydrogenated product, dodecahydro-N-ethylcarbazole, are more
practical candidates for H2 storage and release than carbazole
4. Conclusion and dodecahydrocarbazole.
N-Ethylcarbazole hydrogenation was faster than the hydro-
genation of carbazole over a supported Ru catalyst in the
temperature range 130-150 °C, whereas the selectivity to the Acknowledgment
completely hydrogenated products, dodecahydro-N-ethylcarba-
zole and dodecahydrocarbazole was 95% and 97%, respectively. Funding for the present study from Natural Sciences and
Both N-ethylcarbazole and carbazole hydrogenation reactions EngineeringResearchCouncilofCanadaisgratefullyacknowledged.
1026 Ind. Eng. Chem. Res., Vol. 49, No. 3, 2010

Literature Cited (12) Hegedus, L.; Mathe, T. Hydrogenation of Pyrrole Derivatives Part
V. Poisoning Effect of Nitrogen on Precious Metal on Carbon Catalysts.
(1) Oelerch, W.; Klassen, T.; Bormann, R. Mg-based Hydrogen Storage Appl. Catal. A: Gen. 2002, 226, 319–322.
Materials with Improved Hydrogen Sorption. Mater. Trans. 2001, 42, 8. (13) Vargas, A.; Brgi, Th.; Baiker, A. Adsorption of cinchonidine on
(2) Satyapal, S.; Petrovic, J.; Read, C.; Thomas, G.; Ordaz, G.; The, platinum: a DFT insight in the mechanism of enantioselective hydrogenation
U. S. Department of Energy’s National Hydrogen Storage Project: Progress of activated ketones. J. Catal. 2004, 226, 69–82.
towards meeting hydrogen-powered vehicle requirements. Catal. Today (14) Hodoshima, Sh.; Arai, H.; Takaiwa, Sh.; Saito, Y. Catalytic Decalin
2007, 120, 246–256. Dehydrogenation/Naphthalene Hydrogenation Pair as a Hydrogen Source
(3) Pez, G.; Scott, A.; Cooper, A. C.; Cheng, H. Hydrogen Storage by for Fuel-cell Vehicle. Intl. J. Hydrogen Energy 2003, 28, 1255–1262.
ReVersible Hydrogenation of Pi-conjugated Substrates, US Patent 7,429,372, (15) Stull, O. R.; Westrum, E. F.; Sinke, G. C. The Chemical
2008. Thermodynamics of Organic Compounds; John Wiley & Sons Inc.: New
(4) Clot, E.; Eisenstein, O.; Crabtree, R. H. Computational structure- York, 1969; pp 15-20.
activity Relationships in H2 Storage: How Placement of N Atoms Affects (16) Newson, E.; Truong, T. B.; Hottinger, P.; Roth, F. V.; Schucan,
Release Temperatures in Organic Liquid Storage Materials. Chem. Commun. T. H. H. Optimization of Seasonal Energy Storage in Stationary Systems
2007, 22, 2231–2233. with Liquid Hydrogen Carriers, Decalin and Methylcyclohexane. In
(5) Material Matters. Hydrogen Storage Materials; Aldrich, 2007; Vol. Proceedings of the 12th World Hydrogen Energy Conference, Buenos Aires,
2. Argentina, June 21-25, 1998; pp 935-942.
(6) Pez, G.; Scott, A.; Cooper, A. C.; Cheng, H.; Wilhelm, F. C.; (17) Loutfy, R. O.; Veksler, E. M. Investigation of Hydrogen Storage
Abdourazak, A. H. Hydrogen Storage by ReVersible Hydrogenation of Pi- in Liquid Organichydrides. In Proceedings of the International Hydrogen
conjugated Substrates. US Patent 7,351,395, 2008. Energy Forum, Munich, Germany, Sept 11-15, 2000; pp 335-340.
(7) Zhao, H. Y.; Oyama, S. T.; Naeemi, E. D. Hydrogen storage using (18) Crawford, P.; Burch, R.; Hardacre, C.; Hindle, K. T.; Hu, P.; Kalirai,
heterocyclic compounds: The hydrogenation of 2-methylthiophene. Catal. B.; Rooney, D. W. Understanding the Dehydrogenation Mechanism of
Today, published online April 5, http://dx.doi.org/10.1016/j.cattod.2009.02.039. Tetrahydrocarbazole over Palladium Using a Combined Experimental and
(8) Eberle, U.; Felderhoff, M.; Schüth, F. Chemical and Physical Density Functional Theory Approach. J. Phys. Chem. 2007, 111, 6434–
Solutions for Hydrogen Storage. Angew. Chem. Int. Ed. 2009, 48, 6608– 6439.
6630. (19) Hindle, K. T.; Burch, R.; Crawford, P.; Hardacre, C.; Hu, P.; Kalirai,
(9) Pez, G.; Scott, A.; Cooper, A. C.; Cheng, H. Hydrogen Storage by B.; Rooney, D. W. Dramatic Liquid-phase Dehydrogenation Rate Enhance-
Reversible Hydrogenation of Pi-conjugated Substrates. US Patent 7,101,530, ments Using Gas-phase Hydrogen Acceptors. J. Catal. 2007, 251, 338–
2006. 344.
(10) Sotoodeh, F.; Zhao, L.; Smith, K. J. Kinetics of H2 recovery from
dodecahydro-N-ethylcarbazole over a supported Pd catalyst. Appl. Catal.
A: Gen. 2009, 362, 155–162. ReceiVed for reView April 30, 2009
(11) Zieverink, M. M. P.; Kreutzer, M. T.; Kapteijn, F.; Moulijn, J. A. ReVised manuscript receiVed August 20, 2009
Gas-Liquid Mass Transfer in Benchscale Stirred Tanks: Fluid Properties Accepted August 28, 2009
and Critical Impeller Speed for Gas Induction. Ind. Eng. Chem. Res. 2006,
45, 4574–4581. IE9007002

You might also like