c2ee22066k paper

Upload: carlos-a-ramos

Post on 09-Mar-2016

217 views

Category:

Documents


0 download

DESCRIPTION

otro papercito :)

TRANSCRIPT

  • Comparison of catalytic performance of supported ruthenium and rhodium forhydrogenation of 9-ethylcarbazole for hydrogen storage applications

    Katarzyna Morawa Eblagon, Kin Tam and Shik Chi Edman Tsang*

    Received 27th April 2012, Accepted 29th May 2012

    DOI: 10.1039/c2ee22066k

    The stepwise hydrogenation of 9-ethylcarbazole to 9-ethyl-perhydro

    hydrogenated intermediate(s) was studied over a number of support

    stan

    ion

    l an

    but

    e in

    ha

    moderate activity gave a higher selectivity to the fully hydrogenated product under comparable

    conditions. It was also found that the presence of a hydrophilic support such as alumina or rutile can

    storage and supply to polymer

    on board of a vehicle has

    most of the LOH systems are

    us they are compatible with

    hich could drastically reduce

    ng to hydrogen as an energy

    drogen from these carriers is

    ree and can be used directly to

    l, most LOH systems with

    Dynamic Article LinksC

  • Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article Onlineintrinsically high hydrogen capacities4 are safe and cost effective

    for the future implementation of hydrogen economy.5

    The use of 9-ethylcarbazole as a reversible hydrogen storage

    material at moderate conditions was proposed by Air Products6,7

    with other researchers independently proposing the use of

    a general class of N-heterocycles.7,8 Initial theoretical and

    experimental studies indicated that the partial substitution of

    carbon atoms by nitrogen favours low temperature release of

    hydrogen from organic liquids.5,7,8 Thus, 9-ethylcarbazole is one

    promising material for hydrogen storage utilizing the LOH

    approach. The reversible catalytic hydrogenation of 9-ethyl-

    carbazole, studied in the present work, is shown in Fig. 1.

    The primary proposal was to execute the catalytic dehydro-

    genation reaction of 9-ethyl-perhydro-carbazole to supply

    hydrogen in a PEMFC vehicle.6,7 As a result, the dehydrogena-

    tion reaction and mechanism were extensively studied.913 On the

    other hand, there is only limited information on the hydroge-

    nation of 9-ethylcarbazole.1416 This is an important reaction

    from the point of view of material regeneration at hydrogen

    filling stations, although it is still likely that regeneration will be

    carried out at a central facility or in a factory. Therefore, an in-

    depth understanding of catalytic systems on reaction selectivity

    and yield is of importance for the future implementation of the

    LOH strategy in the application for mobile hydrogen storage.

    We have previously reported a comparative study of the liquid

    phase hydrogenation of 9-ethylcarbazole over a series of ruthe-

    nium based catalysts dispersed on different carriers14 and the

    studies for different unsupported metals for the same reaction

    were also iniated.15 Similarly, the catalytic performances over

    a commercial 5 wt% Ru/Al2O3 (ref. 10 and 13) and RANEYnickel catalyst16 were extensively studied by other researchers in

    this area. In this study, the particular interest is paid to the nature

    of metal and support on the overall catalytic performance for this

    reaction.

    Norskov and Hammer17 suggested that the adsorption prop-

    erties of many substrates depend on the surface electronic

    structure of the transition metal catalyst, which in turn is

    determined by its chemical composition and morphology. A

    good correlation between the position of the d-band centre of

    Fig. 1 Reversible catalytic hydrogenation of 9-ethylcarbazole.a particular metal, adsorption strength of atoms/molecules17,18

    and catalytic activity19,20 was established. In addition, the elec-

    tronic properties of metals can also be influenced by the presence

    and type of support used. The support can alter the electron

    density of the metal leading to a change in its catalytic proper-

    ties.21,22 Moreover, the support can sometimes increase the

    catalytic activity of the metal by providing additional active sites

    for substrate adsorption, as discussed by Chou and Vannice.23

    In this paper, a detailed comparison of the catalytic perfor-

    mances of ruthenium and rhodium over common support

    8622 | Energy Environ. Sci., 2012, 5, 86218630materials alongside commercial supported palladium was carried

    out in the liquid phase hydrogenation of 9-ethylcarbazole. The

    focus on the two metals was motivated by their high activities

    reported in the hydrogenation of pyrroles.24 Ruthenium was also

    previously found by our group to be the most active catalyst in

    the hydrogenation of 9-ethylcarbazole.15,25 Additionally,

    rhodium was also found to be active in hydrogenation of

    aromatic compounds.26 Thus, in the present work, the kinetics of

    the stepwise hydrogenation of 9-ethylcarbazole based on

    fundamental reaction pathways were compared over Ru and Rh

    supported catalysts. The influences of the type of active metal

    and support on the activity and selectivity of the catalysts in this

    reaction will also be discussed.

    2. Experimental

    2.1. Materials

    All the materials were used as received, without further purifi-

    cation. Sodium borohydride (99%), ruthenium(III) chloride

    hydrate (3842%), tris(acetylacetonato)ruthenium(III) (Ru-70,

    24.59%), rhodium(III) acetylacetonate (97%), oleic acid, dioctyl

    ether (99%), oleylamine (70%) and TiO2 (rutile) were purchased

    from Sigma-Aldrich. Commercial catalysts, namely: 5 wt% Ru

    on activated carbon (AC), 5 wt% Rh on AC, 5 wt% Ru on Al2O3and 5 wt% Rh on Al2O3 were also purchased from Sigma

    Aldrich. Activated carbon (activated, Ash 4%max), was received

    from Johnson Matthey. The HPLC grade solvents (acetone,

    cyclohexane, hexane and ethanol) were supplied by Fisher-

    Scientific. Hydrogen gas (99.9%) was provided by Air-products.

    2.2. Catalyst preparation

    Supported noble metal catalysts were either purchased from

    Sigma Aldrich (designated as COM) or were prepared using

    mainly two methods in the present work. The as-prepared

    catalysts were used without further pretreatments or activation.

    The single step method was a mild chemical reduction (desig-

    nated as CR) of the metal precursor impregnated on the

    support. Briefly, an appropriate amount of ruthenium(III)

    chloride hydrate salt (0.3 g) was dissolved in 10 mL of distilled

    water. This solution was then added dropwise to the aqueous

    slurry of the support (typically 2.9 g of support in 15 mL of

    distilled water). After 48 hours of stirring, the mixture was

    reduced by addition of a solution of sodium borohydride (0.4 g

    in 10 mL of distilled water) at 90 C. Subsequently, the catalystwas washed five times with distilled water and acetone and

    centrifuged at 8000 rpm for 10 minutes each time. Then it was

    left to dry overnight at room temperature under ambient

    atmosphere.

    The two-step method was based on a modified polyol process

    (designated as P), developed by Sun et al.27 followed by

    a subsequent impregnation of the metal colloid on the support.

    In this method, the sol of metal nanoparticles was produced

    following the polyol process. Mainly, an appropriate amount of

    tris(acetylacetonato)ruthenium(III)/rhodium(III) acetylacetonate

    (0.16 g/0.11 g) was dissolved in 10 mL of dioctyl ether (99%

    Aldrich). Then the appropriate amount of oleic acid and

    oleylamine (65 ml/68 ml) for ruthenium, and (42 ml/45 ml) for

    rhodium was added to the solution of the precursor, followedThis journal is The Royal Society of Chemistry 2012

  • Two main reaction intermediates, namely: 9-ethyl-tetrahy-

    drocarbazole (Pl 4[H] addition of 4 hydrogen atoms to the

    substrate) and 9-ethyl-octahydrocarbazole (Pl 8[H] addition of

    8 hydrogen atoms) were identified as the main key intermediates

    using GC-MS. The fully saturated product 9-ethyl-perhy-

    drocarbazole (Pl 12[H] addition of 12 hydrogen atoms) was

    separated into three different liquid fractions of isomers, namely

    Fig. 3 Reaction model of direct and stepwise hydrogenation of 9-eth-

    ylcarbazole to intermediate(s) and products over noble metal catalysts; Pl

    0[H]: 9-ethylcarbazole, Pl 4[H]: 9-ethyl-tetrahydrocarbazole, Pl 8[H]: 9-

    ethyl-octahydrocarbazole, Pl 12[H]: 9-ethyl-perhydrocarbazole. The

    stereoisomers of 9-ethyl-perhydrocarbazole (A, B and C) are shown in

    Fig. 2 and the numbering of the rate constants matches with those used

    for modelling of the kinetics.

    Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article Onlineby 0.14 g/0.9 g for ruthenium/rhodium of 1,2-hexadecanediol as

    a reducing agent. The mixture was stirred continuously under

    nitrogen atmosphere for 30 minutes in a refluxing set-up

    without heating. Then it was refluxed at 240 C for 40 minutesunder nitrogen to form a black sol. After cooling to room

    temperature, the catalysts were washed five times with a 50%/

    50% mixture of ethanol and hexane and centrifuged at 5000 rpm

    for 10 minutes each time. After washing was completed, the

    catalysts were left to dry at ambient conditions for 12 h.

    Subsequently, the produced metal catalyst was wet-impregnated

    onto the support. The solution containing 0.015 g of dried

    nanoparticles in 5 mL of ethanol was added dropwise to 0.285 g

    of support dispersed in 10 mL of ethanol. Finally, the supported

    nanocatalyst was dried at ambient conditions for 12 hours

    before testing. The dry powder was used for catalysis without

    any further pre-treatment.

    2.3. Catalyst characterisation

    The specific surface area of the catalyst was measured following

    a standard N2-BET method.

    The particle size and the dispersion were evaluated from the

    HRTEM JEOL 2010 with an accelerating voltage of 300 kV at

    a nominal magnification of 590k. Catalysts samples were

    dispersed ultrasonically in ethanol . A droplet of the suspension

    was then placed on a 200-mesh copper grid coated with formvar

    carbon and left to dry at 80 C overnight. Determination of themean particle size distribution was based on the measurements of

    the minimum of 100 particles from different areas of the sample

    using the Scandium software from Olympus Soft Imaging

    Solutions.

    2.4. Catalyst activity and selectivity

    The hydrogenation of 9-ethylcarbazole was carried out in a Parr

    stainless steel 300 mL batch reactor with continuous monitoring

    and control of stirring speed, temperature, and pressure. In

    a typical experiment, 1 gram of 9-ethylcarbazole was dissolved in

    100 mL of cyclohexane followed by addition of 0.2 gram of

    a selected catalyst. The reactor was then sealed, flushed with

    hydrogen and heated to 130 C. When the desired temperaturewas reached, 70 bar of pure hydrogen was charged into the

    reactor and the reaction time was taken as zero. After the

    required reaction time, the autoclave was cooled using a water

    bath. For study of the reaction kinetics, small aliquots of the

    reaction mixture (usually less than 0.5 mL) were removed from

    the reactor periodically and analysed. The composition of the

    samples was analysed using a GC-MS instrument (Agilent 6890-

    5975E GC-MS) equipped with a non-polar capillary column

    (Agilent 19091s-433), an auto-sampler and an MSD-Triple Axis

    Detector HED-EM. The experimentally obtained reproducibility

    of the results showed a deviation of less than 5% on the values

    obtained for both the conversion and selectivity in the repeated

    series of experiments.

    The catalytic activity and selectivity of the catalysts were

    estimated based on the eqn (1) and (2) listed below.

    Catalytic activity 9-ethylcarbazole convertedmolestotal metal contentgtimes (1)This journal is The Royal Society of Chemistry 2012Selectivity specific productmolessum of all the productsmoles 100 (2)

    2.5. Modelling of the reaction kinetics

    Fig. 2 Structures of the isomers of 9-ethyl-perhydrocarbazole obtained

    using NMR techniques.25 Fraction B actually contained two stereo-

    isomers indistinguishable by means of GC-MS. As a result, this mixture

    was treated as a single product in the quantitative analysis of the reaction

    kinetics. The numbering of the rate constants matches with the numbers

    used in the modelling of the kinetics.Energy Environ. Sci., 2012, 5, 86218630 | 8623

  • Pl 12[H] A, B and C (structures shown in Fig. 2). Furthermore,

    after the hydrogenation of the substrate was completed, the

    isomers of 9-ethyl-perhydrocarbazole were found to be inter-

    converted to each other at significant rates (see Fig. 2). More

    detailed analysis of the reaction products using NMR tech-

    niques was reported in our previous work.14,25 Based on the

    analysis of the experimentally obtained concentration versus

    time profiles, the reaction model was developed, which is shown

    in Fig. 3.

    In order to fit the experimental data into the proposed reaction

    model, the appropriate differential equations were readily solved

    to obtain the initial product concentrationtime profiles of each

    species by using the 4th and 5th order RungeKutta formulas.28

    For example, the rate equation describing the changes of

    the concentration of 9-ethyl-tetrahydrocarbazole (Pl 4[H]) is

    shown below.

    dP4

    dt k2P0 k5P4

    where P0 is the concentration of 9-ethylcarbazole (Pl 0[H]) and

    P4 is the concentration of 9-ethyl-tetrahydrocarbazole (Pl 4[H]).

    Similar rate equationswere solved for eachof the reactions shown

    in the reaction model in Fig. 3. The boundary conditions imposed

    were: at time zero, concentration of the substrate (P0[H] 100(mol%)) and the concentration of other species was equal to 0.

    The first rate constant k0 was determined by a regression

    analysis of the substrate concentrationtime profile (derived

    from the conversion) against a first order integrated rate law

    where [P0]t0 represents the concentration of the substrate P0 attime zero 100 (mol%). d represents the correction factor for anylag time due to experimental error, which was also determined

    from the regression analysis. The rate constant k0 represents the

    overall first order decomposition rate constant of the starting

    material, and it is equal to the summation of k1, k2, k3, k6 and k9.

    The remaining rate constants were derived using a NelderMead

    method (simplex method).29 The mismatch between experimental

    and modelled values was calculated using a root mean square of

    the difference (RMS).30 Thus, the unknown rate constants were

    solved iteratively to fulfil the conditions described by the reaction

    model and to optimise a set of k values that would result in

    a minimum mismatch. The rate constant k0 was treated as

    In Fig. 4, the obtained HRTEM images show distorted

    (0.2

    Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article Onlineaccording to the Equation shown below;

    [P0] [P0]t0exp(k0(t + d))

    Table 1 Physical properties of the studied catalysts. COM arecommercial catalysts, CR are chemically reduced by the NaBH4 methodand P are catalysts synthesized using the polyol process

    CatalystPreparationmethod TEM size (s.d.) [nm] BET [m2 g1]

    5 wt% Ru on AC COM 3.0 (1.0) 8095 wt% Ru on TiO2 CR 2.1 (0.6) 1915 wt% Ru on TiO2 P 3.1 (0.6) 1805 wt% Ru on Al2O3 COM 9.1 (4.7) 835 wt% Rh on AC COM 3.9 (1.8) 4425 wt% Rh on Al2O3 COM 3.3 (0.4) 2725 wt% Rh on AC P 5 wt% Pd/AC COM 2.7 (0.6) 801

    Table 2 Catalytic activity and selectivity quenched after 1 h of reaction

    CatalystCatalytic activity 102 [mM ETC per g of metal per s]

    5 wt% Ru/AC 27.85 wt% Ru on TiO2 (CR) 13.95 wt% Ru on Al2O3 (COM) 13.55 wt% Ru on TiO2 (P) 13.45 wt% Rh on AC (COM) 13.15 wt% Rh on Al2O3 (COM) 13.25 wt% Rh on AC (P) 8.35 wt% Pd/AC (COM) 4.78624 | Energy Environ. Sci., 2012, 5, 86218630spherical nanoparticles mostly uniformly distributed on the

    supports. This suggests that the samples contain highly

    g catalyst was used except in the case of 5 wt% Ru/AC (0.1 g))

    Selectivity (%)

    Pl 4[H] Pl 8[H] Pl 12[H]A Pl 12[H]B Pl 12[H]C

    1 32 44 19 30 9 4 77 100.5 58 6 33 20.5 54 7 39 36 6 50 33 6

  • Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article Onlineunsaturated sites such as edges and corners on the sharp crystal

    surfaces. From the image of 5 wt% Ru/Al2O3 (Fig. 4B), it can be

    seen that the particles are agglomerated/sintered probably due

    Fig. 4 Typical HRTEM images of the selected catalysts. A 5 wt% Ru/AC

    Ru/TiO2 (CR).

    This journal is The Royal Society of Chemistry 2012to the high temperature of reduction. In the HRTEM image of

    5 wt% Ru/TiO2 (Fig. 4D), the metal fringes are clearly visible in

    contrast to the support.

    (COM), B 5 wt% Ru/Al2O3 (COM), C 5 wt% Rh/AC and D 5 wt%

    Energy Environ. Sci., 2012, 5, 86218630 | 8625

  • 3.2. Hydrogenation activity and selectivity

    According to the Sabatier principle the catalytic activity is at

    its maximum if the substrate/intermediate chemisorbs onto the

    surface of the catalyst with an appropriate strength.31 If the

    adsorption is too weak, the catalytic activity will be low and

    the rate limiting step will be the adsorption of the substrate. On

    the other hand, if the adsorption is too strong, the rate will be

    limited by the regeneration of the surface sites of the catalyst. In

    this case of the hydrogenation of 9-ethylcarbazole, the ability of

    the surface d-band electrons of a particular metal to participate

    in bonding with the substrate is likely to be involved in the rate

    determining step. The shift of d band position across the periodic

    table suggests that the metal situated more towards the right

    hand side with a lower d band state as relative to the Fermi level

    consequently has a weaker surface bonding with the substrate

    and, as a result, has lower overall activity in the reaction.

    It is interesting to note that the selectivity of the key inter-

    mediate Pl 8[H] is proportional to the catalytic activity as

    Fig. 5 Catalytic activity versus selectivity towards stable intermediate Pl

    8[H] quenched after 1 h of the reaction.

    Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article OnlineThe obtained catalytic activity and selectivity of the supported

    metals in liquid phase hydrogenation of 9-ethylcarbazole in

    a stirred tank reactor are summarised in Table 2.

    The results were compared after the first hour of the reaction

    due to the high substrate conversions obtained with longer

    reaction time by some active catalysts. Based on the obtained

    data, the influence of different factors such as type of the metal,

    support and the synthesis method can be compared

    systematically.

    3.2.1. Influence of the different noble metals (Ru, Rh, Pd all

    supported on AC). Comparing Ru, Rh and Pd commercial

    catalysts dispersed on AC, Table 2 reveals that the activity per

    gram of metal decreases in this reaction despite the large differ-

    ences in the values of BET surface area (see Table 1) as the

    metal d-band centre decreases (Ru 1.41, Rh 1.73,Pd 1.83).32 This observation is also consistent with ourprevious report of the poor activity of Pt/AC,25 which has the

    lowest d band centre value (2.25 eV).32 The change of activity ofthe metals in this reaction can be described by the position of the

    active metal in the periodic table, and decreases as elements shift

    from the left hand to the right hand side in the second row of

    transition metals.Fig. 6 Time dependent product distribution obtained with 0.1 g of

    5 wt% Ru/AC (COM) catalyst during 4.5 hours of the reaction at 130 Cand 70 bar of hydrogen. The labels of different species are displayed.

    8626 | Energy Environ. Sci., 2012, 5, 86218630shown in Fig. 5. The most active 5 wt% Ru/AC gave the highest

    selectivity to Pl 8[H]. This suggests that the production of this

    intermediate is favourable at higher substrate conversions in the

    stepwise hydrogenation, but at the same time the readsorption

    for further conversion of this intermediate on the same catalyst

    surface is somewhat hindered on the surface of Ru metal. This

    could be due to the competitive adsorption between substrates

    (substrate, intermediates and products) on the same catalytic

    sites. On the other hand, the difference of a surface structure of

    Ru (hexagonal close packed structure) as compared to Rh, Pd

    or Pt (all face centre cubic structures) may account for the

    unusually high stability of the Pl 8[H] intermediate. In addition,

    the difference in surface coverage of hydrogen species over

    these different metals may also play a role in defining their

    selectivity.

    In order to compare the relative rates of competitive adsorp-

    tions (conversions) for different species on catalyst surfaces, their

    individual concentration versus time plots were obtained. The

    results for hydrogenations over Ru/AC, Rh/AC and Pd/AC

    commercial catalysts are shown in Fig. 68. A kinetic model for

    the stepwise hydrogenations shown in Fig. 3 was used to derive

    the fundamental rates for the elementary surface reactions.

    The rate constants for each elementary step with 5 wt%

    Ru/AC, 5 wt% Rh/AC and 5 wt% Pd/AC catalysts used for the

    Fig. 7 Time dependent product distribution obtained with 0.2 g of

    5 wt% Rh/AC (COM) catalyst during 5 hours of the reaction at 130 Cand 70 bar of hydrogen. The labels are the same as in Fig. 6.This journal is The Royal Society of Chemistry 2012

  • Table 3 The calculated rate constants in [h1] for 5 wt% Ru/AC (0.1 g catalyst), 5 wt% Rh/AC and 5 wt% Pd/AC

    Catalyst k0 k1 k2 k3 k4 k5 k6 k7 k8 k9 k10 k11

    5 wt% Ru/AC (COM) 3.4 0.6 1.6 1 0.001 3.7 0.22 0.001 1.058 0.002 0.13 0.0035 wt% Rh/AC (COM) 3.0 1.4 0.59 0.62 0.92 1.3 0.28 0.53 0.433 0.11 0.13 05 wt% Pd/AC (COM) 0.33 0.049 0.22 0 0.008 0.24 0.03 0.005 9.449 0.03 0.027 0.015

    Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article Onlinekinetic fitting of the obtained data to the model are summarised

    in Table 3.

    Careful inspection of the k values listed in Table 3 together

    with corresponding concentrationtime profiles, shows that the

    calculated k0 decreases in the order Ru > Rh > Pd, which is

    the same trend as the comparison of activity after the first hour of

    Fig. 8 Time dependent product distribution obtained with 0.2 g of

    5 wt% Pd/AC (COM) catalyst during 5 hours of the reaction at 130 Cand 70 bar of hydrogen. The labels are the same as in Fig. 6.

    Fig. 9 Time dependent product distribution obtained with 5 wt% Ru/

    Al2O3 (COM) catalyst during 5 hours of the reaction at 130C and 70 bar

    of hydrogen.

    Table 4 The calculated values of rate constants for the 9-ethylcarbazole hyd

    Catalyst k0 k1 k2 k3 k4

    5 wt% Ru/AC (COM)a 3.4 0.60 1.6 1 0.0015 wt% Ru/TiO2 (CR) 13 0.008 0.57 7.5 1.35 wt% Ru/Al2O3 (COM) 12 0.93 0.65 9.2 1.15 wt% Ru/TiO2 (P) 6.3 2.7 0.1 4.9 0.3

    a 0.1 g catalyst was used instead of 0.2 g.

    This journal is The Royal Society of Chemistry 2012the reaction (Table 2). It is noted that the derived k0 is not

    excessively large compared to other rates with the 5 wt% Ru/AC

    catalyst. Hence with the high substrate conversion during the

    first hour (Fig. 6), the rates (especially the k8) for Pl 8[H]

    conversions would suggest that this intermediate is able to

    Fig. 10 Time dependent product distribution obtained with 5 wt% Ru/

    TiO2 (CR) catalyst during 4 hours of the reaction at 130C and 70 bar of

    hydrogen. The labels are explained in Fig. 9.

    Fig. 11 Time dependent product distribution obtained with 5 wt% Ru/

    TiO2 (P) catalyst during 4 hours of the reaction at 130C and 70 bar of

    hydrogen. The labels are explained in Fig. 9.

    rogenation with supported ruthenium catalysts

    k5 k6 k7 k8 k9 k10 k11

    3.7 0.22 0.001 1.1 0.002 0.13 0.00314.4 2.0 0.47 0.035 3.5 0 0.1916 1.4 0.22 0.19 0 0.068 0.0334.0 0.3 0.003 0.16 0.36 0 0.021

    Energy Environ. Sci., 2012, 5, 86218630 | 8627

  • Similarly, the main kinetic stable intermediate Pl 4[H] was found

    when reaction was taking place over 5 wt% Pd/AC catalyst.

    Comparison of the rate constants in Table 3 with corresponding

    concentrationtime profiles (Fig. 68) revealed that over 5 wt%

    Ru/AC and 5 wt% Pd/AC, the reaction takes place mostly by

    consecutive addition of hydrogen atoms, whereas over 5 wt%Rh/

    AC, the reaction appears to progress primarily via the direct

    hydrogenation of the substrate, hence no significant accumula-

    tion of any intermediate product(s) was observed.

    Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article OnlineFig. 12 Time dependent product distribution obtained with 5 wt% Rh/

    Al2O3 (COM) catalyst during 5.5 hours of reaction at 130C and 70 bar

    of hydrogen.further convert to the fully hydrogenated products. It is thus

    interesting to find out why the selectivity for Pl 8[H] reaches

    >32% after the first hour of the reaction (Table 2). On the other

    hand, from Table 3 it can be also seen that in the case of 5 wt%

    Ru/AC, the rates of Pl 8[H] formation (k3, k5) are much higher

    than the rates of its further conversion (k4, k7, k8). But, the rates

    of Pl 8[H] formation over 5 wt% Pd/AC and 5 wt% Rh/AC are

    much lower than the rates of its further conversion over these two

    catalysts. Therefore, the unusual high stability of the Pl 8[H]

    intermediate using 5 wt% Ru/AC could be attributed to its weak

    interaction with the surface of this catalyst. This interaction is

    further decreased by the distorted structure of this intermediate

    resulting in the steric hindrance with the catalysts surface.

    Fig. 13 Time dependent product distribution obtained with 5 wt% Rh/

    AC (P) catalyst during 5.5 hours of reaction at 130 C and 70 bar ofhydrogen.

    Table 5 The calculated rate constants for 5 wt% Rh/Al2O3 (COM), 5 wt% R

    Catalyst k0 k1 k2 k3 k4

    5 wt% Rh/Al2O3 (COM) 1.6 1.1 0.50 0 0.0045 wt% Rh/AC (COM) 3.0 1.4 0.59 0.62 0.925 wt% Rh/AC (P) 1.1 0.73 0.52 0 0.87

    8628 | Energy Environ. Sci., 2012, 5, 86218630In addition, from Table 2, it can be seen that the composition

    of the fully saturated product (9-ethyl-perhydrocarbazole) is

    similar for the Ru/AC, Rh/AC and Pd/AC commercial catalysts

    with the most thermodynamically stable Pl 12[H]A as the main

    product. Thus, the formation of the isomers of Pl 12[H] does not

    seem to be affected by the type of the metal used but rather by the

    type of support.

    3.2.2. Influence of the type of support on catalytic performance

    of Ru based catalysts. Activated carbon is known to be an inert

    material giving insignificant metalsupport interaction. In order

    to evaluate the influence of support on catalytic performance,

    ruthenium was deposited on Al2O3 (alumina) and TiO2 (rutile) in

    order to compare with the results of ruthenium supported on

    activated carbon. The concentrationtime profiles obtained over

    these supported ruthenium catalysts are presented in Fig. 911.

    The experimental data analysed with the aid of our kinetic model

    (Fig. 3) generally showed a good fit. The calculated rate

    constants for these supported ruthenium catalysts are presented

    in Table 4.

    Table 4 shows that 5 wt% Ru/AC gives the significantly lowest

    k0 rate compared with other supports. We believe that this could

    be an artefact due to the use of a smaller amount of catalyst (0.1 g

    instead of 0.2 g) in this experiment. Although the calculated rates

    account for the quantity of metal, the coverage of the species on

    metal surfaces can be different. However, comparing the rate

    constants of direct production of Pl 12[H] isomers from the

    substrate (k1, k6 and k9) with the rate constants of the stepwise

    hydrogenations (k2 and k3), the reaction pathway clearly adopts

    the consecutive hydrogenation pathway over all these supported

    ruthenium catalysts, regardless of the type of the support used. In

    addition, the high selectivity to the Pl 8[H] intermediate is very

    pronounced over all supports as the rates of Pl 8[H] formation

    (k3, k5) are consistently higher than the rates of its further

    conversion (k4, k7, k8).

    One main difference when using different carriers is the

    composition of the final product Pl 12[H]. In the case of 5 wt%

    Ru/AC (COM), as mentioned before, the main product was the

    isomer Pl 12[H]A (see Fig. 6). This isomer accounted for over

    65% of the total sum of isomers of the 9-ethyl-perhydrocarbazole

    only after 1 hour of the reaction (Table 2). But, in the case of

    h/AC (COM) and 5 wt% Rh/AC (P)

    k5 k6 k7 k8 k9 k10 k11

    3.6 0 0.71 1.9 0 0.21 0.0041.3 0.28 0.53 0.43 0.11 0.126 01.0 0.09 0 0 0.43 0.046 3.7This journal is The Royal Society of Chemistry 2012

  • Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article Onlinebased catalysts. As discussed, the catalysts based on rhodium

    displayed slightly lower activity than those based on ruthenium,

    but they did not produce large quantities of intermediate(s),

    resulting in higher selectivity towards the fully hydrogenated

    products under comparable conditions. Therefore, it seems of

    importance to identify the influence of the type of support on Rh

    based catalysts.

    The concentrationtime profiles are shown in Fig. 12 and 13.

    The profile for 5 wt% Rh/AC was presented in Fig. 7 for

    comparison. The rate constants derived from our model are

    presented in Table 5.

    It is noted from Table 5 that regardless of the synthetic method

    or type of support used, the catalysts give a high rate for direct

    hydrogenation of 9-ethylcarbazole to Pl 12[H]A (k1). The main

    product for all the supported Rh catalysts is the isomer Pl 12[H]

    A, which is similar to the result with 5 wt% Ru/AC.

    It is interesting to note that the use of more hydrophilic

    alumina as the support (5 wt% Rh/Al2O3) resulted in the main

    initial product being Pl 12[H]B (see Fig. 12), which is similar to

    the case of Ru. However, during the progress of the reaction, the

    isomer Pl 12[H]B was further converted to Pl 12[H]A by an

    isomerisation step (k10). This result agrees very well with our

    previous observation that the composition of the final products

    (isomers of 9-ethyl-perhydrocarbazole) is influenced mainly by

    the type of support used. It is clear that the use of the activated

    carbon support tends to favour the thermodynamically most

    stable isomer Pl 12[H]A, whereas over alumina and rutile

    supports, the production of Pl 12[H]B is facilitated.

    4. Discussions and conclusion

    Fundamental understanding of the elementary steps of the

    reversible hydrogenation of 9-ethylcarbazole over different cata-

    lytic surfaces is important regarding the supply of hydrogen gas

    for mobile PEMFC devices. From the point of view of spent fuel

    recovery, the rate of hydrogenation for the carrier compound

    should be as high as possible. On the other hand, the existence of

    kinetically stable intermediates (for example, Pl 4[H] over Pd

    systems or Pl 8[H]withRu)may create technical difficulties. Thus,

    an ideal liquid organic hydride (LOH) storage material should be

    able to interchange between loaded and unloaded forms during

    hydrogenation (material regeneration) and dehydrogenation

    (delivery of hydrogen gas). Any stable intermediates would mean

    the reduction of total storage capacity and would definitely resultother supports, the main product was Pl 12[H]B instead. Thus, it

    seems that the final isomer distribution can be significantly

    influenced by the type of support used. According to the calcu-

    lated rate constants in Table 4, the stepwise hydrogenation of Pl

    8[H] to isomer Pl 12[H]A is via k4 and to isomer Pl 12[H]B is via

    k8 route (Fig. 3). While it is known that Pl 12[H]A is the most

    thermodynamically stable product, the surface of hydrophilic

    supports such as TiO2 and Al2O3 must apparently influence the

    position of the hydrogen addition resulting in the formation of

    the symmetrical Pl 12[H]B product (much higher k4). Thus, the

    role of the type of support can direct the stepwise hydrogenations

    to form the kinetically favoured but thermodynamically less

    stable Pl 12[H]B product.

    3.2.3. Influence of the support on catalytic performance of RhThis journal is The Royal Society of Chemistry 2012in longer times for hydrogen delivery and regeneration. In addi-

    tion, the produced undesirable isomers such as trans- isomers of

    9-ethyl-perhydrocarbazole in the hydrogenation of 9-ethyl-

    carbazole may not possess the stereochemistry required for the

    dehydrogenation steps. It should be noted that it is more favoured

    to remove two hydrogen atoms from amolecule in the cis form on

    terrace sites of the metal catalysts.7

    The catalytic hydrogenation of 9-ethylcarbazole on the noble

    metal surfaces appears to involve a series of consecutive reactions

    of addition of hydrogen to double bonds to produce the final

    fully hydrogenated 9-ethyl-perhydrocarbazole product. These

    elementary steps can take place purely on the metal surface as

    a direct hydrogenation process depending on the adsorption

    strengths of substrate/intermediate(s). Alternatively, the rela-

    tively weak adsorption of some intermediate(s) onto the metal

    surface with respect to their molecular solvation can result in the

    formation of kinetically stable intermediates in solution. As

    a result, only a combined concentrationtime profile analysis

    with rate constants elucidation from a kinetic model may allow

    systematic comparison of catalyst performances.

    In the present work it was found that Ru is the most active

    metal followed by Rh and Pd in liquid phase hydrogenation of

    9-ethylcarbazole, regardless of the type of support used. Here, we

    invoke the general d-band model which describes the interaction

    between renormalized valence states of a substrate molecule and

    the d-states of the transition metal catalyst. In atomistic sight; the

    bond strength of the substrate molecule on the catalyst surface

    would depend on the d-band position in relation to the Fermi

    level.17 Due to the fact that antibonding states are always above

    d-states, the position of the d-band centre with respect to the

    Fermi level is a good factor describing the reactivity of the

    transition metal surfaces.17 In principle, the higher the energy of

    the d-band centre in relation to the Fermi level, the higher the

    antibonding orbital and the stronger the bond.17,32 However, the

    changes in substrate, surface feature and geometry of adsorption

    can affect the adsorbate levels that couple to the metal and as

    a result can modify the bond strength. In the light of the relative

    positions of the d-band centres of the studied metals (Ru, Rh and

    Pd), the catalytic activities appear to correlate well with the

    position of the d-band. This clearly indicates that the rate

    determining step of hydrogenation of 9-ethylcarbazole lies in the

    ability of 9-ethylcarbazole adsorption onto the metal catalyst

    surface, which is not significantly influenced by the presence of

    support. It is thus clear that ruthenium is the most active metal in

    this reaction, however it suffers from relatively poor selectivity

    towards fully hydrogenated products, due to the production and

    accumulation of large quantities of the Pl 8[H] intermediate. Our

    results show that the surface interaction of Pl 8[H] with the

    surface of the ruthenium catalyst seems to be weak. This is due to

    the change in the compound structure from Pl 0[H] to Pl 8[H]

    (containing a stable pyrrole moiety which hinders the re-

    adsorption of this intermediate on terrace sites of the catalyst14)

    influencing its conversion into the fully saturated product.

    Similarly, the Pd based catalyst suffers from the accumulation of

    Pl 4[H] and poorer activity as compared to Ru in the 9-ethyl-

    carbazole hydrogenation reaction. Rh based catalysts with some

    characteristic surface features appeared to provide moderate

    binding strengths for both substrate and intermediates in this

    reaction, hence resulting in a good overall activity and selectivity.Energy Environ. Sci., 2012, 5, 86218630 | 8629

  • Based on the obtained results, it is clear that the interaction of the

    substrate and intermediates with a metal catalyst may be modi-

    fied through the use of a support or by alloying the active metal

    with another metal. The alloying of two metals shifts the position

    of the d-band centre of the metal and, as a result, modifies the

    strength of the interaction between catalyst and reaction

    substrates/intermediates/products.17 Regarding the support

    effect, it was found that the use of oxide support such as Al2O3 or

    TiO2 can facilitate the production of kinetically favoured cis-

    isomers of 9-ethyl-perhydrocarbazole (Pl 12[H]B). It is thought

    that hydrophilic surfaces at the metalmetal oxide interface can

    affect the adsorption of intermediate(s). However, the isomer-

    isation of Pl 12[H]B would further convert this product into the

    thermodynamically most stable isomer Pl 12[H]A.

    In conclusion, the type and stability of the intermediates in the

    hydrogenation of 9-ethylcarbazole is governed in majority by the

    electronic structure of the active metal whereas the initial

    Catal. Today, 2003, 82, 119.3 P. D. Tien, T. Satoh, M. Miura and M. Nomura, Fuel Process.

    7 G. Pez, A. Scott, A. Cooper, H. Cheng, F. Wilhelm andA. Abdourazak, US Pat., 7351395, Air Products and Chemicals,Inc., USA, April 11, 2008.

    8 A. Moores, M. Poyatos, Y. Luo and R. H. Crabtree, New J. Chem.,2006, 30, 1675.

    9 M. Sobota, I. Nikifordis, M. Amende, B. Sanmartin Zanon,T. Staudt, O. Hofert, Y. Lykhach, C. Papp, W. Hieringer,M. Laurin, D. Assenbaum, P. Wasserscheid, H. P. Steiruck,A. Gorling and J. Libuda, Chem.Eur. J., 2001, 17, 11542.

    10 F. Sotoodeh, L. Zhang and K. J. Smith, Appl. Catal., A, 2009, 362,155.

    11 Z. Wang, I. Tonks, J. Belli and C. M. Jensen, J. Organomet. Chem.,2009, 694, 2854.

    12 F. Sotoodeh and K. J. Smith, J. Catal., 2011, 279, 36.13 F. Sotoodeh and K. J. Smith, Ind. Eng. Chem. Res., 2010, 49, 1018.14 K. M. Eblagon, K. Tam, K. M. K. Yu, S. L. Zhao, X. Q. Gong,

    H. He, L. Ye, L. C. Wang, A. J. Ramirez-Cuesta and S. C. Tsang,J. Phys. Chem. C, 2010, 114, 9720.

    15 K. M. Eblagon, K. Tam, K. M. K. Yu and S. C. E. Tsang, J. Phys.Chem. C, 2012, 116, 7421.

    16 X. Ye, Y. An and G. Xu, J. Alloys Compd., 2011, 509, 152.

    Publ

    ished

    on

    17 Ju

    ly 2

    012.

    Dow

    nloa

    ded

    by U

    nive

    rsid

    ad A

    uton

    oma

    de N

    uevo

    Leo

    n on

    28/

    01/2

    016

    16:5

    6:48

    . View Article OnlineTechnol., 2008, 89, 415.4 R. B. Binwale, S. Rayalu, S. Devotta and M. Ichikawa, Int. J.Hydrogen Energy, 2008, 33, 360.

    5 R. H. Crabtree, Energy Environ. Sci., 2008, 1, 134.6 G. Pez, A. Scott, A. Cooper and H. Cheng, US Pat., 7101530, AirProducts and Chemicals, Inc., USA, May 9, 2006.and ISIS, UK and Dr Anibal J. Ramirez-Cuesta for various

    supports of this project.

    References

    1 D. Sebastian, E. G. Bordeje, L. Carvalio, M. J. Lazaro andR. Moliner, Int. J. Hydrogen Energy, 2008, 33, 1329.

    2 N. Itoh, E. Tamura, S. Hara, A. Shono, K. Satoh and T. Namba,composition of the fully saturated 9-ethyl-perhydrocarbazole is

    mostly influenced by the type of support present. It is believed

    that careful kinetic analysis of products and intermediates is

    crucial for future energy supply based on Liquid Organic

    Hydrogen Carriers.33,34

    Acknowledgements

    Mr Fernando Eblagon and Dr K. M. Kerry Yu are kindly

    acknowledged for fruitful discussions and help in data analysis.

    The authors are indebted to the University of Oxford, EPSRC8630 | Energy Environ. Sci., 2012, 5, 8621863017 B. Hammer and J. K. Norskov, Surf. Sci., 1995, 343, 211.18 W. Tang and G. Henkelman, J. Chem. Phys., 2009, 130, 194504.19 F. H. B. Lima, J. Zhang, M. H. Shao, K. Sasaki, M. B. Vukmirovic,

    E. A. Ticianelli and R. R. Adzic, J. Phys. Chem.C, 2007, 111, 404.20 V. R. Stamenkovic, B. S. Mun, M. Arenz, K. J. J. Mayrhofer,

    C. A. Lucas, G. Wang, P. N. Ross and N. M. Markovic, Nat.Mater., 2007, 6, 241.

    21 T. H. Fleisch, R. F. Hicks and A. T. Bell, J. Catal., 1984, 87, 398.22 G. Blyholder, J. Mol. Catal., 1996, 119, 11.23 P. Chou and M. A. Vannice, J. Catal., 1987, 107, 129.24 L. Hegedus, T. Matheb and A. Tungler, Appl. Catal., A, 1996, 147,

    407.25 K. M. Eblagon, D. Rentsch, O. Friedrichs, A. Remhof, A. Zuettel,

    A. J. Ramirez-Cuesta and S. C. Tsang, Int. J. Hydrogen Energy,2010, 35, 11609.

    26 N. Hiyoshi, M. Osada, C. V. Rode, O. Sato and M. Shirai, Appl.Catal., A, 2007, 331, 1.

    27 S. Sun, C. B. Murray, D. Weller, L. Folks and A. Moser, Science,2000, 287, 1989.

    28 K. J. Johnson,Numerical Methods in Chemistry, Marcel Dekker, NewYork, 1st edn, 1980.

    29 K. H. Borgwardt, The Simplex Method. A Probabilistic Analysis,Springer-Verlag, 1st edn, 1987.

    30 T. G. Kolda, R. M. Lewis and V. Torczon, SIAM Rev., 2003, 45, 385.31 G. A. Tritsaris, J. K. Norskov and J. Rossmeisl, Electrochim. Acta,

    2011, 56, 9783.32 A. Ruban, B. Hammer, P. Stolze, H. L. Skriver and J. K. Norskov, J.

    Mol. Catal., 1997, 115, 421.33 D. Teichmann, W. Arlt, P. Wasserscheid and R. Freymann, Energy

    Environ. Sci., 2011, 4, 2767.34 P. Makowski, A. Thomas, P. Kuhn and F. Goettmann, Energy

    Environ. Sci., 2009, 2, 480.This journal is The Royal Society of Chemistry 2012

    Comparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applications

    Comparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applications

    Comparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applicationsComparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applications