oxidative, reductive, infrared and catalytic studies of supported rhodium … · surface. so it is...

164
Oxidative, reductive, infrared and catalytic studies of supported rhodium catalysts Citation for published version (APA): Vis, J. C. (1984). Oxidative, reductive, infrared and catalytic studies of supported rhodium catalysts. Technische Hogeschool Eindhoven. https://doi.org/10.6100/IR114107 DOI: 10.6100/IR114107 Document status and date: Published: 01/01/1984 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 01. Apr. 2021

Upload: others

Post on 20-Oct-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Oxidative, reductive, infrared and catalytic studies ofsupported rhodium catalystsCitation for published version (APA):Vis, J. C. (1984). Oxidative, reductive, infrared and catalytic studies of supported rhodium catalysts. TechnischeHogeschool Eindhoven. https://doi.org/10.6100/IR114107

    DOI:10.6100/IR114107

    Document status and date:Published: 01/01/1984

    Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

    Please check the document version of this publication:

    • A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

    General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

    • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

    If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

    Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

    Download date: 01. Apr. 2021

    https://doi.org/10.6100/IR114107https://doi.org/10.6100/IR114107https://research.tue.nl/en/publications/oxidative-reductive-infrared-and-catalytic-studies-of-supported-rhodium-catalysts(8f684b51-ac04-4cb6-92cf-fe586f382f97).html

  • "It unfortunately often occurs that names are

    mistaken for explanations, and people deceive

    themselves with the believe that, for instance,in

    attributing chemical decompositions to affinity,

    attraction, contact-force, catalysis etc., they

    explain them."

    A.W. Williamson, 1851.

  • Dit proefschrift is goedgekeurd door de promotoren

    Prof.dr. R. Prins

    en

    Prof.dr. V. Ponec.

  • OXIDATIVE, REDUCTIVE, INFRARED AND

    CATALYTIC STUDIES OF

    SUPPORTED RHODIUM CATALYSTS

    PROEFSCHRIFT

    TER VERKRIJGING VAN DE GRAAD VAN DOCTOR IN DE TECHNISCHE WETENSCHAPPEN AAN DE TECHNISCHE HOGESCHOOL EINDHOVEN, OP GEZAG VAN DE RECTOR ~~GNIFICUS, PROF.DR. S.T.M. ACKERMANS, VOOR EEN COf'.1MISSIE AANGEWEZEN DOOR HET COLLEGE VAN DEKANEN IN HET OPENBAAR TE VERDEDIGEN OP

    VRIJDAG 24 FEBRUARI 1984, TE 14.00 UUR

    DOOR

    JAN CORNELIS VIS

    GEBOREN TE OUD VOSSEMEER

  • Contents.

    page

    1.

    1.1. Future needs of chemical industry 1

    1.2. The importance of metal catalysts 2

    1.3. Oxidation and reduction properties of supported

    metals

    1.4. CO adsorption and I.~.-spectroscopy

    1.5. Synthesis gas reactions and their mechanism

    1.6. References

    2. Reduction and oxidation of Rh/Y-Al 2 g 3 ~~~~·~-~~ catalysts as studied by Temperature Programmed

    Reduction and Oxidation

    2.1. Introduetion

    2.2. Experimental

    2.3. Results

    2.3.1. Hydragen chemisorption

    2.3.2. TPR and TPO of Rh/Y-Al 20 3 2.3.3. TPR and TPO of Rh/Ti0 2

    2.4. Discussion

    2.5. Conclusions

    2.6. References

    3. The and

    Y-Al 223 as studied with Temperature Programmed Reduction-Oxidation and Transmission Electron

    l"licroscopy

    3.1. Introduetion

    3.2. Experimental

    3.3. Results

    3. 3 .1. Hydragen

    3.3.2. TPR/TPO

    3.3.3. TPR/TPO

    chemisorption

    of the RA-series

    of the RT-series

    3.3.4. TEM measurements

    3.4. Discussion

    3.5. Conclusions

    3.6. References

    4

    4

    9

    12

    15

    17

    20

    23

    25

    26

    31

    32

    36

    39

    43

    45

    47

    53

    57

    60

    61

  • 4.

    and Ti0 2 catalysts at elevated pressures

    4.1. Introduetion

    4. 2.

    4.3. Results

    4. 3 .1. Rh/La 2o 3 4.3.2. Rhjy-Al 2o3 4.3.3. Rh/Ti0 2

    4.4. Discussion

    4.5. References

    5. an in-situ infrared cell and

    measurements

    5.1. Introduetion

    page

    66

    68

    73

    79

    82

    83

    85

    5.1.1. General introduetion 88

    5.1.2. of a new in-situ I.R.-cell 90

    5.1.3. CO on Rh, as studied with I.R.-spectroscopy 91

    5.2. Experimental 92

    5.3. Results and discussion 93

    5.4. Conclusions 101

    5.5. References 103

    6.

    6 .1. Introduetion 106

    6. 2. Experimental 109

    6. 3. Results

    6. 3.1. The structure of rhodium af ter reduction 114

    6.3.2. The structure of rhodium af ter co admission 121 6.4. Discussion

    6.4.1. The structure of rhodium af ter reduction 127

    6.4.2. The structure of rhodium af ter co admission 130 6.5. Conclusions 133

    6.6. References 134

  • General Discussion

    Surrunary

    Acknowledgements

    Samenvatting

    Dankwoord

    Curriculum Vitae

    List of publications

    page

    140

    144

    147

    148

    151

    153

    154

  • Chapter 1 page

    Chapter 1. lntrod u ct ion

    1.1. Future needs of chemical industry.

    The role of chemistry in nowaday society is more

    than most people think or wish. The chemical

    industry might be responsible for a major part of

    environrnental pollution, it also answers to numerous

    material needs of mankïnd. And last but not least, if

    solutions have to be found for pollution problems, it will have to be chemistry

    To a great extent this depends on crude oil as

    a raw material and energy carrier. Through various

    processes the crude oil is separated in various fractjons

    of hydrocarbons, which are thereafter mainly used as

    fuels. The very light fractions (containing 1, 2 or 3

    carbon atoms and therefore denoted as Cl, C2 and C3

    fractions) serve as raw material for many processes,

    producing alcohols, aldehydes, polymers and so on.

    The of chemical industry upon society, and the

    vulnerability of this industry through its dependenee on

    crude oil became very clear during the oil crisis of

    1973. Since that time the of crude kept

    Only a couple of years ago crude oil prices were

    predicted to reach $ 100 per barrel in 1990 (1973 $ 2.60). The chemical industry was forced to meet this

    challenge by returning to a cheaper and more abundant raw

    material: Coal. Coal can be to give a mixture

    of hydrocarbons resembling the heavy oil fractions. And

    with known refining and processes they can be

    converted to useful bulk raw materials and fuels. 'Coal

    can also be ~asified. It can react with steam under

    certain conditions and with certain catalysts to yiel.d a

    mixture of carbon monoxide and hydrogen, which is in a 1

    to 1 mixture known as synthesis gas {syngas). Syngas can

    be seen as the new raw material that would take the place

    1

  • Chapter 1 page

    of crude oil in a foreseeable future, and the chemistry

    involving the conversion of syngas to bulk chemieals

    forms, with some other reactions, the socalled c1 chemistry.

    Current projections predict however that the price of

    crude oil will reach only $ 50 in 1990 (current price

    about $ 30), half of the earlier level. This has narrowed the economie advantages presented by c1 chemistry for the

    and nineties. In potential it is still

    promising, but a more cornprehensive analysis of economie

    and technical factors is required to decide whether

    c1 chemistry projects will pay off in the near term or only later. This thesis deals with some of the

    chemical fundamentals of the conversion of syngas, and of

    some of the catalysts involved in it.

    1.2. The irnportance of roetal catalysts.

    Numerous processes in the manufacturing of crude oil

    involve roetal catalysts, because they are able to

    dissociate hydrogen and to transfer it to reacting

    molecules. Metal catalysts contain mostly noble or

    transition metals such as platinum (Pt), palladium (Pd),

    cobalt (Co), rhodium (Rh) or nickel (Ni). The conversion

    of syngas is an excellent example of a reaction proceding

    with metallic elements. In fact this conversion proceeds

    with all the metals mentioned above and some ethersas

    well, iron being the best known example from the Fischer-

    Tropsch process.

    An important item in syngas conversion is whether the

    oxygen atom from carbon monoxide is retained in the

    molecule, leading to products such as methanol, glycol,

    acetic anhydride etc., or removed in the form of water or

    carbon dioxide, leading to hydrocarbons. Rhodium then

    takes a rather special place arnong the catalytic active

    2

  • Chapter 1 page

    metals, since it seems to be able to catalyse processes

    in either direction. Recently, interest in rhodium has

    increased further because it is an irreplaceable compound

    in the three-way catalyst, applied in the USA for

    cleaning exhaust gases of automobiles. Plans exist to

    introduce this three-way catalyst on cars in Europe too,

    and this will undoubtedly promate the research on rhodium

    catalysts very much. However, this thesis is mainly

    directed towards studies of rhodium catalysts in relation

    with cl chemistry.

    As mentioned above,

    place on the roetal

    many catalytic conversions take

    surface. So it is of interest to

    create as large a surface- to volume ratio as possible

    for rhodium, especially in view of the availibility and

    price of Rh. Rh as RhCl3.xHi} casts about f 40/g; as a

    comparison, the price of gold is f 37/g. The ordinary way

    to create this Rh surface is to bring the roetal onto a

    high surface area support, such as silica-alumina's,

    alumina, silica, etc. These kinds of supports are very

    porous, so we can apply the roetal as a salt in an aqueous

    solution, which is suck up into the pores by capillary

    farces. Once we have established the pare-volume of a

    certain support, we can dissolve the wanted amount of

    roetal salt in that particular amount of water and add it

    to the support. It can be assumed that after careful

    drying the roetal salt is present within the pores of the

    support.

    Now we have to bring the catalyst into an active state.

    This is in certain cases the metallic state, which means

    we have to reduce the catalyst. In some cases it may be

    necessary to heat or to oxidize the catalysts first

    (calcination), in order to remave for instanee ligands

    from the roetal salt. This means that the oxidation-

    reduction behaviour of the catalytically active roetal is

    of vital interest.

    3

  • Chapter 1 page

    1.3. Oxidation and reduction properties of supported

    metals.

    An excellent way of Temperature Programmed

    TPR/TPO experiments a

    studying these processes is

    Reduction and Oxidation. In our

    certain amount of catalyst is

    placed in a horizontal, cylindrical quartz reactor,, held

    in an electrical furnace. A mixture of Hz/Ar or o2/He can flow through the catalyst bed, while the temperature of

    the oven can be varied continuously by a temperature

    controller. The ingoing gasstream (of known composition)

    is being compared with the outgoing gasstream. If there

    is a difference in composition (which means consumption

    of one of the components in the reactor), this leads to

    an electrical signal from the detector comparing the

    gasses, and this signal, integrated over time, is

    proportional to the amount of gas consumed. This signal

    is also recorded, and this recording is the so-called

    TPR- or TPü-profile the reader will encounter many times

    in the Chapters to come. A schematic representation of

    the TPR/TPO apparatus is given in Chapter 2. TPR and TPO

    experiments are the main subject of Chapters 2 and 3.

    1.4. CO adsorption and I.R.-spectroscopy.

    In hydracarbon and oxygenate synthesis from CO and H2

    the bonding between CO and the metal catalyst plays an

    important role. We will give a short description of this

    process, following the one given by Coulson (1).

    In the CO molecule there are ten valenee electrons,

    occupying the following molecular orbitals in order of

    increasing energy: 3v, 4v, lpiy, lpiz and Sa. The bonding

    in the molecule is provided for by the 3a and the two lpi

    orbitals. Two electrons occupy the non-bonding 4a

    orbital, made from the oxygen 2s and 2px orbitals and

    4

  • Chapter 1 page

    a

    -QjQ

    b

    therefore concentrated around the oxygen nucleus. The

    last two electrons occupy the 5o orbital formed from the carbon 2s and 2px orbital. They form the lone pair on the

    carbon atom, directed away from the oxygen atom. The

    charge density contours of these orbitals are shown in

    Fig. 1.

    0.1

    0.05

    0.15 Z(nm) -0.10 --0.05

    X(nm) t

    Figure 1. Charge density contours of the 30(a), 40(b), 1~ or 1~ (c) and So(d) orbitals of CO. From ref. (15). y

    z

    In the bonding with the metal this so lone is donated to an empty metal orbital, forming a dative a-bond. Then there is back donation from a filled metal d-

    orbital to the Lowest Unoccupied Molecular Orbital of the

    carbon monoxide molecule, which is the pi*-antibonding

    orbital. This back donation to an antibonding orbital

    causes the carbon-oxygen bond to weaken a little. The

    amount of weakening will be a measure for the electron

    donating capacities of the adsorbing metal atom. Since CO

    has a small dipole moment (0.1 , with the negative

    charge on the C-atom due to the directional properties of

    the 5 o orbi tal) the CO stretching vibration is infrared

    active and so we can study the weakening of the CO bond

    d

    5

  • Chapter 1 page

    upon adsorption by a metal by means of IR-spectroscopy.

    Of course under favourable circumstances.

    -0----0 \

    \... \ \ I I

    '',\~:'./'

    -- ::.-:s~0~- -/ .

    c::~--- --- --·~~ -

    Figure 2. The electrical "image" resulting from a positive charge above-, a dipale vibrating parallel with- and a äipole vibrating perpendicular to a metal surface.

    Pearce and Sheppard (2) discuss an important phenomenon

    first mentioned by Francis and Ellison (3). They argue

    that in the case of IR-studies of species, adsorbed upon

    a metal surface, it is necessary that the IR-radiation

    makes a large angle with the metal surface (near grazing

    incidence), because at the metal surface itself there is

    a knot in the standing-wave field (reflection) and

    absorption cannot be seen there. Only at grazing

    incidence the resulting field (in- plus out- going) will

    have sufficient intensity.

    Another point of interest is that a perfect metal

    surface should be perfectly polarizable and this leads to

    what is called the metal-surface selection rule (2).

    Their line of reasoning is that the electric field lines

    6

  • Chapter 1 page

    from a charge above a roetal surface

    roetal surface, will be directed

    going towards this

    as if there were an

    opposite charge of equal size under the roetal surface at

    equal distance.

    Likewise, a dipole

    virtual dipole of

    the roetal surface.

    above a roetal surface will induce a

    equal size but opposite signs, under

    As can be seen from Fig. 2 the

    vibrations of a dipole parallel to the surface will be

    annihilated by its mirror image under the surface, while

    for a dipole vibrating perpendicular to the surface its

    mirror image will reinforce it.

    For IR-radiation of for instanee 2000 cm-1, the

    wavelength is 5 ~m. Since this exceeds sufficiently the

    thickness of an adsorbed layer, and also the spacing

    between dipole and image-dipole, the physics described

    above predict that only dipole vibrations perpendicular

    to a roetal surface are infrared-active.

    In recent years the contribution of infrared studies to

    the understanding of adsorption processes at solid

    surfaces has increased substantially, a.o. by the

    introduetion of Fourier Transform Infrared Spectroscopy

    (usually indicated as FT-IR spectroscopy). This

    development had been initiated by the invention of the

    interferometer by Michelsen in the late nineteenth

    century (4, 5). Interference of light had been recognized

    long before that time, but with the Michelson

    interferometer it became possible to separate the two

    interfering beams in such a way that their relative path

    differences could be varied precisely. Michelsen already

    realized that the visibility curves of his interference

    patterns contained speetral information, but it was Lord

    Raleigh who recognized that the interferogram was related

    to the spectrum of the radiation passing through the

    interferometer through the mathematica! operatien known

    as the Fourier Transformation (4, 6). The interferometer

    7

  • Chapter 1 page

    played an important role in spectroscopy during the

    following years, but the technology required for the full

    Fourier Transformation was lacking at the time. The

    astronomist Fellgett recognized the possibilities of the

    interferometer in increasing the intensity of the signal

    on the detector compared with the grating spectro-

    photometer, where slits have to be used to create the

    energy dispersion characteristic for the latter

    instrument. By an interferometer, data from all speetral

    frequencies are measured simultanuously. The resulting

    reduction in measurement time is now known as Fellgett's

    advantage (4, 7). It can be expressed more mathematically if one realises that the signal-to-noise ratio increases

    for longer measurement times

    Fellgett

    in

    a lso

    proportion

    was the

    with the

    first to square root of time.

    actually perferm a

    calculate a spectrum.

    numerical Four~er Transformation to

    Another advantage of the interferometer is that the

    "throughput" of radiation,

    flow through the apparatus

    thanks to the absence

    Jaquinot advantage (4, 8).

    a measure for the total energy

    per second, is also greater

    of slits. This is known as the

    The real break-through

    spectroscopy came with the

    for mid-infrared FT-IR

    application of the Cooley-

    Tukey fast Fourier Transferm algorithm to interferometry,

    and the simultaneous development of fast, dedicated

    minicomputers. Spectra that took hours to measure and

    still considerable time to calculate, could now be

    plotted several seconds after starting the measurement of

    an interferogram. Because of these reasons, since 1968

    FT-IR spectroscopy gained an important place in the study

    of heterogeneaus catalysis. We started our infrared

    investigations with the development of an in situ

    infrared cell, and the first results obtained therewith

    are reported in Chapter 5.

    8

  • Chapter 1 page

    l.S. Synthesis gas reactions and their mechanism.

    We tried to point out in this introduetion so far the

    importance of the synthesis of certain chemieals from

    carbon monoxide and hydrogen apart from and together with

    the broad supply of chemieals from natural sources,

    whether or not after refining. We explained how these

    syntheses take place over supported metal catalysts, how

    these catalysts are made and how one can study their

    oxidation and reduction behaviour. Subsequently we dealt

    with the interaction of carbon monoxide with these metal

    catalysts and_ how we can study this interaction with FT-

    IR spectroscopy. What is left to introduce is a view on

    the theories of how the syntheses of chemieals over these

    catalysts are thought to take place.

    For this we refer only to the outstanding review of

    Biloen and Sachtler on the subject (9) and we restriet us

    here to their conclusions.

    The conversion of synthesis gas to hydrocarbons and

    alcohols can be described by three overall reactions:

    n CO + 2n H2 CnH2n + n H20 [1]

    m CO + ( 2m+l) H2 -- CmH2m+2 + m H20 [2] p co + 2p H2 - c p 1H2P_ 1CH20H + p-1 H20 [3] In the presence of the product water the watergas shift

    reaction can occur:

    The products are mainly unbranched parafins, olefins and

    alcohols, and the chain length distribution of the

    molecules usually obeys the so-called Schulz-Flory

    9

  • Chapter 1 page

    distribution, implying that the molecules are formed by

    step-wise chaingrowth propagation, followed by a singular

    termination step. This concept involves the presence on

    the surface of chains Yn and insertable monomers X, leading to the following reaction scheme:

    x - À. etc. [5] The subscript of the product P denotes the number of

    carbon atoms in the product molecule. The Schulz-Flory

    distribution means that the ratio Pn+l over Pn (denoted

    as a) is constant. This implies that the insertable

    monomer X is a Cl species. Biloen and Sachtler suggest it

    might be a -CH2- carbenic species, a conclusion arrived

    at after analysis of the literature concerned. However,

    some writers still prefer a cautious CHx notation. The

    initiating species YQ (and Yl) are also Cl species. The

    activation energy for the formation of methane is usually

    found to be the same as for the formation of higher

    hydrocarbons, so it is assumed that the rate-determining

    step in both processes is the same. Since hydragenation

    is known to be much faster than the Fischer-Tropsch

    reaction, the termination step cannot be rate-

    determining. And since the formation of methane does not

    involve the propagation step, this one can be ruled out

    too. Biloen and Sachtler show that in the overall

    reactions

    co - CH -x

    - CH -x

    [6]

    [7]

    in both cases the first step is a fast one. This implies

    10

  • Chapter 1 page

    that the conversion of -CHx- to -CH2- is rate

    determining. Biloen and Sachtler propose the Fischer-

    Tropsch propagation is a cis migration on the metal atom:

    [8]

    The terminatien steps, hydrogenation and S-H-elimination,

    lead to parafins and olefins, respectively.

    Rhodium catalysts are known to produce alcohols as well

    as hydrocarbons in the CO hydrogenation. Since insertion

    of CO in an alkyl group is thermodynamically favourable

    in the case of rhodium (10), one is tended to formulate a

    reaction mechanism for the formation of alcohols in which

    an insertion step of CO plays a key role. However, no

    definitive mechanism explaining the formation of

    oxygenated products has been given in literature yet. It

    should be remarked here that some related reactions are

    known which make some speculations plausible. For

    instanc;the Monsanto process, the conversion of methanol

    to acetic acid (11). This process is catalysed by a

    mononuclear rhodium complex. The oxidation state of the

    rhodium in the complex changes from I to III and back

    through oxidative addition of reactants and reductive

    eliminatien of products. Polyols can be formed also in a

    homogeneous process, and not only by rhodium catalysts

    (12,13). Watson and Somorjai roughly distinguished three

    temperature areas in which hydrogenation of CO would lead

    mainly to the products indicated (14):

    T

  • Chapter 1

    T >600 K CH _... I x s

    page

    [11]

    It is speculated that steps involving insertion of oxygen containing building blocks take place on oxidic sites, those not being able to dissociate CO because the electrens to donate to the pi*- antibonding orbital of the carbon monoxide. It is generally accepted that

    of hydrocarbons takes place over metallic formation rhodium ensembles. performance of some

    Catalytic catalysts

    studies concerning in the conversion

    the of

    synthesis gas are presented in Chapter 4. Some overall kinetic parameters of the processes described above have been determined.

    In Chapter 6 we come back to the special example of the interaction between carbon monoxide and a rhodium-alumina catalyst. There has long been a controversy in literature as to what processes took place during this interaction, and we used a number of complementary techniques to elucidate this problem. Since this Chapter will be published as such, we refer to its own introduetion for a review of the relevant literature.

    Chapter 7 presents a general discussion of the results presented in this thesis and some suggestions for further research.

    1.6. References.

    1. coulson, C.A. in "Valence", University Press, London 1961.

    2nd Ed., Oxford

    2. Pearce, H.A. and Sheppard, N., Surface Sci. 59, 205 {1976).

    12

  • Chapter 1 page

    3. Francis, S.A. and Ellison, A.H., J.Opt.Soc.Am., 49,

    131 (1959).

    4. Griffiths, P.R., in "Chemical Infrared Fourier

    Transform Spectroscopy", John Wiley and Sons, New

    York 1975.

    5. Miche1son, A.A., Phil.Mag. Ser. 5, 31, 256 (1891).

    6. Lord Ra1eigh, Phil.Mag. Ser. 5, 34, 407 (1892).

    7. Fe11gett, P., in "Proceedings of Aspen International

    Conference

    A.T. Stair

    1970.

    on Fourier Spectroscopy",

    and D.J. Baker, Eds.),

    (G.A. Vanasse,

    AFCRL-71-0019,

    8. Jaquinot, P., 17e Congres du GAMS, Paris, 1954.

    9. Bi1oen, P. and Sacht1er, W.M.H., in "Advances in

    Cata1ysis", Vol. 30, p 165 (D.D. Eley, H. Pines and

    P.B. Weisz, Eds.), Academie Press Inc., New York

    1981.

    10. Kuh1man, E.J. and Alexander, J.J.,

    Chemistry Reviews, 33, 195 (1980).

    Coordination

    11. Pau1ik, F.E. and Roth, J.F., J.C.S.Chem.Oomm., 1578

    (1968).

    12. De1uzarche, A., Fonseca, P., Jenner, G. and

    Kienneman, A., Erdoeh1 und Koh1e, 32, 313 (1979).

    13. Keim, w., Berger, M. and Sch1upp, J.J., J.Cata1., 61, 359 {1980).

    13

  • Chapter 1 page

    14. Watson, P.R. and Sornorjai, G.A., J.Catal., 74, 282

    (1982).

    14

  • Olapter 2 page

    Chapter 2

    The reduction-oxidation behaviour of supported Rh/y-

    Al203 and Rh/Ti02 catalysts as studied with Temperature

    Programmed Reduction and Oxidation.

    Summary.

    Careful preparatien of Rh/Al20 3 catalysts leads to

    ultradisperse systems (H/Rh>l.O). TPR shows that these

    catalysts are almast completely oxidized during

    passivation. Identical preparatien of Rh/Ti02 catalysts

    leads to less disperse systems (H/Rh=0.3), exhibiting two

    reduction peaks in TPR. These peaks are due to the

    reduction of smal!, wel! dispersed Rh20 3 particles, and

    of large, bulklike Rh2o 3 particles. In all cases

    reduction of Rh2o 3 is complete above 450 K. Tio 2 is

    partly reduced by a roetal catalysed proces above 500 K.

    2.1. Introduction.

    Over the past years it has become clear

    catalysts take a special position in

    that rhodium

    the field of

    supported transition roetal catalysts, because they are

    able to produce hydrocarbons as wel! as oxygenated

    products (alcohols, aldehydes, acids) from synthesis gas

    (1-10). Various workers have tried to influence the

    selectivity and activity of the rhodium catalysts via a

    special p .. :eparation (1,2), via additives (3-5) or mixed

    oxides (8-10), and where this worked out in the wanted

    direction of higher production of oxygenates at least two

    of them gave as the reasen the presence of stabilized Rh+

    15

  • Chapter 2 page

    in the surface (8,10). Same authors claimed the presence

    of isolated Rh+ sites in monometallic Rh catalysts on the

    basis of IR evidence (Worley et al. (11-13), Primet

    (14)), while others

    metallic rafts with

    (15)).

    found the Rh to be present as

    electron microscopy (Yates et al.

    In all cases it seems obvious that the support plays an

    important role in either bringing or keeping the roetal in

    a certain state of (un)reactivity. A special example of

    such a roetal support interaction has been discovered by

    Tauster et al. (16,17) and is now known as Strong Metal

    Support Interaction: Supported metals such as Pt and Rh

    loose their capability for chemisorption of H2 , co- and NO if they have been reduced at high temperatures (e.g. 773

    K) on supports such as Ti0 2 and V20 5. Normal

    chemisorption behaviour can be restored by oxidation at

    elevated temperatures, followed by low temperature

    reduction, at for instanee 473 K (16). SMSI has been

    related to the occurrence of lower oxides of the support

    (18, 19), but the exact nature of the interaction still

    remains unc1ear.

    Many of the above mentioned phenomena have to do with

    one camman property: The oxidation-reduction behaviour of

    supported Rh cata1ysts. We decided to study two systems,

    representative for many of the ones referred to above:

    2.3 wt% Rh/Al2o 3 and 3.2 wt% Rh/Ti02 . Via sintering (see Experimental) we have induced a variatien in partiele

    size (dispersion) in order to answer the fo11owing

    questions:

    - how is oxidation-reduction influenced by partiele size

    - how is oxidation-reduction influenced bu the support

    used

    -does partiele size show any effect upon SMSI.

    Befare we come to the experimental techniques, we have

    to introduce one last item: Passivation. Since it is

    16

  • Chapter 2 page

    obvious that reduced systems cannot simply be stared in

    air, we passivate and stabilize them by applying a layer

    of oxygen upon the metal particles in a controlled

    {see Experimental). Some authors have already

    way

    paid

    attention to the state catalysts are in after starage in

    air. Thus Burwell jr. et al. used Wide Angle X-ray

    Scattering, Extended X-ray Absorption Fine Structure

    (EXAFS), hydragen chemisorption and hydrogen-oxygen

    titration to characterize their supported Pt and Pd

    catalysts {20-23).

    We will show that a good insight in all these matters

    can be gained with the aid of Temperature Programmed

    Reduction and Oxidation (TPR and TPO), supported by

    chemisorption measurements. TPR as a characterization

    technique was presented by Jenkins et al. in 1975 {24,

    25) and has been used extensively in the past few years,

    as becomes obvious from a recent review by Hurst et al.

    (26). The technique allows one to obtain (semi)quan-

    titative information about the rate and ease of reduction

    of all kinds of systems, and once the apparatus has been

    built, the analyses are fast and relatively cheap. We

    used an aparatus as described by Boer et al. (27), which

    enabled us to extend the analyses to Ternperature

    Prograrnmed Oxidation, and to gather information about the

    rate and ease of oxidation as well.

    2.2. Experirnental.

    Tio2 (anatase, Tioxide Ltd., CLDD 1367, surface area 20

    m2/g, pore volume 0.5 cm3/g) and y-AlzOJ (Ketjen, 000-

    1.5E, surface area 200 m2/g, pore volume 0.6 cm3/g) were

    impregnated with ·an aqueous salution of RhC13 .xH20 via

    the incipient wetness technique to prepare a 2.3 wt% Rh/Al 2o 3 catalyst and a 3.2 wt% Rh/Tio 2 catalyst, as was established spectrophotornetrically afterwards.

    17

  • Chapter 2 page

    catalysts were dried in air at 355, 375 and 395 K for 2

    hrs successively, followed by direct pre-reduetion in

    flowing H2 at 473, 773 or 973 K for one hour. Prior to

    removing the catalysts from the reduction reactor they

    were passivated at room ternperature by replacing the

    hydrogen flow by nitrogen and subsequently slowly adding

    oxygen up to 20 %. Then the catalysts were taken out of the reactor and stored for further use.

    Programmer

    Vent.

    Figure 1. Schematic representation of the TPR/TPO apparatus.

    The TPR/TPO apparatus used is schematically represented in Fig. 1: A 5% H2 in Ar or a 5% 0 2 in He flow (300

    ml/hr) can be directed through a microreactor. The

    temperature of the reactor can be raised or lowered via

    linear programming. H~ or o2 consumption is being monitored continuously by means of a Thermal Conductivity

    Detector (TCD). A typical sequence of actions is as fellows:

    - the passivated or oxidized sample is flushed under Ar flow at 223 K

    - Ar is replaced by the Ar/H 2 mixture, causing at least

    an apparent H2 consumption (first switch peak)

    - the sample is heated under Ar/H 2 flow with 5 K/min to

    873 K

    18

  • Chapter 2 page

    - after 15 min at 873 K the sample is caoled down with 10 K/min to 223 K

    - the reduced sample is flushed with Ar Ar flow is replaced by the Ar/H~ mixture once again, now causing only an apparent H2 consumption (second

    switch peak).

    A similar sequence is followed during TPO. The switch peak procedure deserves sorne closer attention. The strong

    signa! we cal! the first switch peak is mainly due to the

    displacement of Ar by Ar/H2 in the reactor, but in sorne cases real hydragen consumption takes place, even at 223

    K. Therefore we repeat the whole procedure after the TPR

    has been perforrned: In that case the catalyst has been

    reduced and caoled down to 223 K, and as a consequence is

    covered with hydrogen. Then we replace the Ar/H 2 by pure Ar, resulting in a negative TCD signa!. Subsequently we switch back to Ar/H2 • Since we do nat expect any hydragen consumption frorn the reduced, hydragen covered sample at

    this time, the resulting switch peak will be due solely

    to the displacement of Ar by Ar/H 2. Sa the difference between the first and second switch peak reveals the real

    hydragen consumption at 223 K, if there is any.

    The reactions that might take place during TPO and TPR

    are:

    4 Rh 0.75) [1]

    [2]

    The quantities between brackets are the amounts of

    hydragen or oxygen consumption per rhodium expected for

    reduction of bulk Rh2o3 or formation of this very materal (apart from chemisorption of any kind).

    Chernisorption measurements were carried out in a

    conventional volumetrie glass apparatus after reduction

    19

  • Chapter 2 page

    of the passivated catalysts at 473 or 773 K in flowing H

    for 1 hr followed by evacuation at 473 K for one hr.

    After H2 admission at 473 K desarptien isotherms were

    measured at room temperature. As desarptien became only

    noticeable at pressures below 200 torr we believe that

    the chemisorption value above that pressure is

    representative of monolayer coverage (cf. Frennet et al.

    (28)). The H/Rh values thus obtained for the various

    systems are preseuted in Table 1.

    H/Rh T(pre-red)

    RT(LT) (l) (K) RA

    473 1. 70 0.37

    773 1. 53 0.29

    973 1.23 0.12

    (1): reduced in situ at 473 K. (2): reduced in situ at 773 K.

    RT(HT)( 2 )

    0.08

    0.05

    0.01

    Table 1. Hydrogen chemisorption of Rh/y-Al 2 o 3 (RA) and Rh/Ti0 2 (RT) catalysts.

    2.3. Results.

    2.3.1. Hydragen chemisorption.

    The hydragen chemisorption data as given in Table 1

    were obtained for Rh/Al2o3 (RA) after reduction of the passivated catalyst in situ at 473 K. Rh/Ti0 2 (RT) was

    also reduced in situ at 773 K, to induce SMSI behaviour.

    The catalysts will be denoted from now on as RA 773

    (Rh/Al2o3 , pre-reduced at 773 K), RT 973 (Rh/Ti0 2, pre-

    20

  • Chapter 2 page

    reduced at 973 K), etc., that is de code refers to the

    temperature of reduction of the dried, impregnated

    cata1yst (see Experimenta1). The results are represented

    graphically in Fig. 2, showing the value of H/Rh as a

    2.0

    1.5

    1.0

    0.5

    473 673 873 1073

    Temp. pre-red.(K)

    Figure 2. Hydragen chemisorption as a function of pre-reduetion temperature; a) Rh/Y-Alz03, reduced in situ at 473 K. b) Rh/Ti0 2 , reduced in situ at 473 K. c) Rh/Ti02, reduced in situ at 773 K.

    function of pre-reduetion temperature (i.e. reduction

    prior to passivation; reduction prior to chemisorption

    was at either 473 or 773 K, as indicated). The systems

    discussed here are represented by the separate dots.

    Va1ues for Rh/Al2o

    3 range from 1.70 to 1.00, for non-

    SMSI Rh/Ti0 2 from 0.37 to 0.12 and for SMSI Rh/TiC2 from

    0.10 to o.oL

    21

  • Chapter 2 page 22

    a

    b

    c

    d

    273 473 673 ~ 873 T(K)

    Figure 3. 2. 3 wt% Rh/Y-Alz03 catalyst; a) TPR of passivated catalyst, pre-reduced

    at 473 K. b) TPR of passivated catalyst, pre-reduced

    at 773 K. c) TPO following TPR of the catalysts. d) TPR following TPO of the catalysts.

  • Chapter 2 page

    The Temperature Programmed Reduction profiles of the

    passivated RA 473 and RA 773 catalysts are shown in Fig.

    3 a and b. The horizontal axis shows the temperature and

    the vertical axis the hydragen consumption (in arbitrary

    units).

    RA 473 shows a maximum H2 consumption at 330 K, and

    some further reduction above 473 K. RA 773 shows a single

    consumption peak around 273 K, foliowed by desarptien of

    H2 . The pre-reduetion had apparently been complete, and

    passivatien of this ultra disperse Rh catalyst had led to

    dissociative oxygen chemisorption, but nat full

    oxidation, since the net H2 consumption mounted up to 1.0

    H2/Rh. Apparently the pre-reduetion at 473 K (RA 473) had

    nat been complete, and the oxygen chemisorbed on this

    system was harder to remave than from the other ones (RA

    973 showed an identical TPR profile as RA 773). The

    subsequent TPO, Fig. 3c, which was identical for all

    three systems, confirms these observations: 02

    consumption starts at 223 K, in the switch peak,

    continues when the temperature ramp is started, reaches a

    maximum around 290 K, and then decreases very slowly

    towards higher temperatures. Integration of the 02

    consumption signa! proved to be difficult, due to the

    small sample sizes (typically 50-75 micromale of metal)

    and the small thermal conductivity of o2 but still mounted up toa rather satisfying 0.6-0.7 o2/Rh (0.75 was expected, since chemisorbed H2 had been removed by a heat

    treatment in between TPR and TPO). The equality of the

    TPO profiles of the RA 473, RA 773 and RA 973 catalysts

    is in accordance with the fact that all these catalysts

    have been brougt up to 873 K during the TPR run and it is

    also nat surprising to find that the TPR profiles of the

    23

  • Chapter 2 page

    completely oxidized systems (following TPO) are identical

    too (Fig. 3d). one single peak is observed around 360 K,

    corresponding with

    which agrees with the

    bulk Rh 2o3 showed

    a H2 consumption of about 1.5 H2/Rh, reduction of Rh 2o3 • Unsupported

    a reduction peak at 400 K in our

    apparatus, while bulk Rh roetal only started to become oxidized above 870 K.

    R3 473 673

    Figure 4. 3.2 wt% Rh/Ti02 catalyst;

    ~ T{K)

    a

    b

    c

    873

    a) TPR of ~assivated catalyst, pre-reduced at 473 K.

    b) TPO following TPR of the catalysts. c) TPR following TPO of the catalysts.

    24

  • Chapter 2 page

    2.3.3. TPR and TPO of Rh/Ti~.

    The TPR profile for the passivated RT 473 catalyst is

    shown in Fig.' 4a. The most striking feature is that H2 consumption starts already at 223 K in the switch peak,

    that is immediately when H2/Ar is being flushed through the reactor. Keeping in mind·the much lower H/Rh value of

    this catalyst compared to the Rh/Al 2o3 series, we think that passivatien here has caused only the formation of an

    outer layer of oxide on the relatively large roetal

    particles. If the remaining metallic care could be

    reached by the hydragen molecules, these molecules could

    be able to dissociate and provide atomie hydragen for an easy reduction of the oxide layer at low temperatures.

    There is also some H2 consumption just above 473 K, as

    for the corresponding Rh/Al203 catalyst, indicating that

    also for Rh/Ti02 the pre-reduetion at 473 K had not been

    complete. It can be seen that the Ti02 support is

    reducible also, leading to H consumption around 573 and

    700 K. The H2 consumption at 273 K amounts to about 0.4

    H2 /Rh.

    The TPR profiles for RT 773 and RT 973 are similar, but

    since the Rh surface area decreases with increasing pre-

    reduetion temperature (cf. Table L) the amount of

    passivatien oxygen decreases also, and so does,

    consequently, the H2 consumption at 223 K in the TPR

    profiles. Also the consumption just above 473 K has

    disappeared as a consequence of the higher pre-reduetion

    temperature.

    25

  • Chapter 2 page

    Fig. 4b shows the TPO profile of a reduced RT catalyst.

    The three TPO profiles of RT 473, 773 and 973 are alike:

    Three "areas" of oxygen consumption show up, Which we

    attribute respectively to chemisorption (which is the

    only phenomenon on Rh/Alz0 3 ), corrosive chemisorption and

    thorough oxidation. We will come back to this assignment

    in the Discussion part.

    The TPR profiles of the oxidized samples of the RT

    series are identical again (Fig. 4c) and show two clearly

    divided Hz consumption maxima, at 325 and 385 K.

    Oonsumption in the first peak is about 1.3 Hz/Rh, in the

    second one about 0.3 Hz/Rh. Taken together the Hz

    consumptions come close enough to the expected value of

    1.5 Hz/Rh to attribute them both to reduction of Rhz03·

    The peak at 385 K is assigned to bulklike Rh 2o3 particles (note that the peakmaximum for unsupported Rh zO 3 is at

    400 K) and the one at 325 K to a better dispersed ~0 3 phase. Furthermore the possibility of some support

    reduction taking place here as well, in advance of the

    support reduction around 573 and 700 K, cannot be

    excluded.

    2.4. Discussion.

    Hydrogen chemisorption has been used through the years

    by many workers to characterize metal surfaces (7, 9, 10,

    14, 16, 20-23, 29, 30) and when attempts were made to

    calculate metal surface areas from hydrogen chemisorption

    data, a hydrogen to metal stoechiometry of one was used.

    On the other hand, if CO was involved, some authors (31)

    chose a stoechiometry of one, while others (14) mentioned

    higher stoechiometries. We are of the opinion that if one

    accepts a metal atom, such as Rh, to adsorb two or more

    CO molecules, one should not reject the idea of the same

    atom adsorbing more than one hydrogen atom. So we think

    26

  • Chapter 2 page

    that although our experimental H/Rh results exceed unity

    (Table 1), they are real, and that the hydrogen

    chemisorbed does not exceed a monolayer (cf. Ftennet et

    al. (28)), and is all bound to the metal. We dit not try

    to make a distinction between socalled "reversible" and

    "irreversible" adsorption, because we believe that in the

    thermodynamic sense there is no such distinction. In

    following the procedures that lead to that supposed

    distinction one merely encounters the physical

    restrictions of ones apparatus, such as pumping speed and

    conductivity of tubing.

    As described in the Experimental section we admitted

    hydrogen at 473 K. This was simply done to accelerate

    adsorption but did not effect the ultimate amount of

    adsorption, as was checked via the measuring of

    adsorption isotherms at room temperature (32). This

    leaves us with chemisorption values above unity, and

    therefore, like some other authors (21, 30), we find it

    impossible to calculate a partiele size or a dispersion

    from these data, since there is no particular

    stoechiometry value to prefer. We imagine these Rh

    particles could be raftlike as suggested by Yates et al.

    (15) (although they were dealing with only 0.5 wt% Rh/Al 20 3), where the edgeatoms could have the possibility

    of adsorbi.ng more than one hydrogen atom. "Sintering" of

    these particles -H/Rh decreases from 1.7 to 1.0 upon

    reduction at 973 K- would then mean that the number of

    edgeatoms decreases, for example by growing of the rafts

    or even formation of (hemi)spherical particles. Still we

    are dealing with ultradispersed systems. We estimate that

    in all cases the Rh partiele size does not exceed 1 nm.

    Fbr the RT series the H/RH values as established after

    473 K reduction are much lower, which was to be expected

    taking into account the difference in surface area

    between Al 20 3 and TiO 2 and the similar me tal loadings.

    27

  • Chapter 2 page

    Here also we see a decrease in the H/Rh values as a

    result of an increase in the pre-reduetion temperature.

    But in this case this is simply due to growth of the

    metal particles, resulting in a decreased metal surface

    area. The H/Rh values after 773 K reduction show evidence for SMSI behaviour. In this case it means they are that

    small that they are of the order of magnitude of the

    experimental error. Therefore we cannot, at this point,

    draw any quantitative conclusion about a relationship

    between SMSI behaviour and partiele size. \'ie knew in

    advance this would be difficult, since both partiele

    growth and increasing SMSI behaviour lead to less

    hydrogen chemisorption, and so the limits of experimental

    accuracy might prohibit to make a distinction between the

    two effects. It is evident, though, that RT 773 and RT 973 must have been in the SMSI state after pre-reduction, but they showed normal chemisorption behaviour after

    passivation

    chemisorption

    following the

    state.

    and re-reduction at 473 K in the apparatus. This means that the passivation,

    pre-reduction, must have destroyed the SMSI

    Our TPR and TPO results also prove that passivation is

    rather drastic. Fbr the small metal particles on Alzo 3

    TPR of the passivated samples differs fram that of the

    oxidized ones only in the position of the peak, that is

    the ease of reduction. And from the shape of the TPO

    pattern we can understand that a long passivation time

    (like storage in air) comes close to a real oxidation.

    But it is very clear that no matter what the initia!

    state of the Rh was, passivated or oxidized, reduction is

    complete above 400 K.

    Worley and coworkers studied the oxidation state of Rh

    on various supports, and starting from various Rh

    precursors, applying IR specroscopy upon CO adsorption

    (11-13). Apart from IR absorptions attributed to CO

    28

  • Chapter 2

    adsorbed on metallic Rh,

    which they attributed

    they

    to CO

    page

    found some absorptions

    adsorbed on isolated Rh+

    sites. Fbr 2.2 wt% Rh catalysts they found these isolated

    Rh+ sites to be more abundant for Al2o 3 than for Ti02 as

    a support (13), and more abundant for RhCl3 than for

    Rh(N0 3) 3 as a precursor (12, 13). They concluded that one

    gets the worst reduction of Rh (they reduced at 673 K)

    when using Al 2o 3 as a support, and when starting from

    RhC1 3 as a precursor. From the TPR evidence presented

    here, we conclude that the systems they have studied must

    have been completely reduced prior to admission of 00,

    and therefore we prefer another explanation. One obtains

    the best dispersion of metallic Rh particles on Al203,

    and when starting from RhC1 3 . Upon co adsorption the smaller particles break up and create the isolated

    dicarbonyl species that were attributed to Rh+l by Worley

    et al. EXAFS proof for this explanation will be published

    elsewhere (33).

    Our findings for Rh/Ti02 go very well along with the

    results published by Burwell et al. for Pd and Pt on SiO and Al2o 3 (20-23). Upon passivatien the larger metal

    particles farm an oxide skin, the formation of which can

    beseen almast literallyin Fig.4b, the TPO of Rh/TiOz.

    The small metal particles on Alz03 show only one tailing

    chemisorption peak, but for Rh/Ti0 2 oxygen chemisorption

    is followed at higher temperatures by corrosive

    chemisortion and finally, around 700 K, by thorough

    oxidation. Apparently oxygen diffusion through the oxide

    layer is a strongly hindered process. Attribution of the

    intermediate temperature region of oxidation to corrosive

    chemisorption is supported by the results of a TPO we did

    on a 3 wt% Rh/Sio2 (Grace Silica, S.P. 2.-324,382, 290 m2/g, H/Rh=0.4). This TPO is shown in Fig. Sa, and

    exhibits two oxygen consumtion areas, one around 273 K

    which is due to oxygen chemisorption, and one around 500

    29

  • Chapter 2 page

    K due to corrosive chemisorption. All of the Rh on Si02 is in a well dispersed form, as was confirmed by the

    subsequent TPR of this Rh/Si02 catalyst, which showed

    only one hydrogen consumption peak at 335 K. This

    reduction peak corresponds to the first peak observed for

    Rh/Ti0 2 at 325 K (cf. Results).

    The oxidation peak around 700 K in the TPO of Rh/Ti02

    is attributed to the formation of rather large, bulklike

    Rh20 3 particles, the reduction of which is observed as a

    separate H 2 consumption maximum at 385 K in TPR (Fig.

    4c). That the large particles oxidize at higher

    temperature than the small particles and reduce at a

    higher temperature as well, has been proven by a rather

    simple experiment, shown in Figs. Sb and c. First a TPO

    is run with a reduced Rh/Ti0 2 system, up to 673 K.

    Subsequently a TPR is performed which demonstratee that

    the reduction peak at 385 K has completely disappeared.

    Thus the fraction of the Rh which reduces at 385 K

    oxidizes above 673 K, and vice versa.

    / ----,/ ~

    ~ \_

    273 473 673 ........ T(K}

    a

    b

    c

    873

    Figure 5. a) TPO of a reduced 3.0 wt% Rh/SiOz catalyst. b) TPO up to 673 Kof a reduced 3.2 wt%

    Rh/Ti02 catalyst. c) Subsequent TPR.

    30

  • Chapter 2 page

    That one can distinguish between a well dispersed phase

    and a bulklike phase of Rh2o3 on a· support has been noticed before by Yao et al. (34), although they used Al 20 3 as a support and had to oxidize for 12 hrs at 973 K

    to create the bulk phase.

    The fact that part of the Ti02 support is being reduced

    as well, by a metal-assisted process, is in agreement

    with findings for Pt/Ti02 (35, 36). Reoxidation of the support takes place during TPO,

    exceeds the expected o2/Rh value of but is apparently hidden in the

    since 02 consumption

    0.75 in all cases,

    TPO profile by the stronger 0 2 consumption caused by the Rh oxidation.

    2.5. Cbnclusions.

    The 2. 3 wt% Rh/A~ o3 catalysts proved to be "ultradisperse" with H/Rh values ranging fran 1. 7 to l.O.

    The catalysts behaved accordingly in TPR and TPO: Easy

    reduction and fast oxidation were observed to such an

    extent that even a mild passivation led to almast

    complete oxidation.

    The 3.2 wt% Rh/Ti02 catalysts were much less dispersed

    (H/Rh 0.37-0.12) and showed evidence of two distinct

    forms of Rh (and Rh2 o3 ), appearing as two reduction peaks in TPR (reduction of well dispersed and bulklike Rh2o 3) and three oxidation areas in TPO (oxygen chemisorption

    and corrosive chemisorption of well dispersed Rh, and

    thorough oxidation of bulklike Rh).

    A start was made in investigating the influence of

    partiele size upon oxidation-reduction behaviour, but we

    were not able yet to establish a relationship between

    SMSI behaviour and partiele size. All Rh/Ti0 2 samples

    could be brought into the SMSI state.

    We shall continue this search

    dispersion, via the me tal content

    by varying the

    and by varying the

    31

  • Chapter 2 page

    reduction procedure.

    This investigation has shawn unambiguously that the

    TPR-TPO technique is a very powertul tool in discriminating between the various ways in which oxygen

    can react with a metal, and so TPR-TPO allows a careful

    analysis of the state of dispersion of the metal on the

    catalyst to be made.

    2.6. References

    1. Ichikawa, M., Bull.Chem.Soc.Jap. 51, 2268 (1978).

    2. Ichikawa, M., Bull.Chem.Soc.Jap. 51, 2273 (1978).

    3. Leupold, E.I., Schmidt, H.-J., Wunder, F., Arpe, H.-

    J. and Hachenberg, H., E.P. 0 010 295 Al.

    4. Wunder, P.A., Arpe, H.-J., Leupold, E.I. and Schmidt,

    H.-J., Ger. Offen. 28 14 427.

    5. Bartley, W.J. and Wilson, T.P., Eur.Pat.Appl. 0 021

    443.

    6.

    7.

    s.

    9.

    Castner, D.G., Blackadar, R.L. and Somorjai, G.A. I

    J. Catal. 66, 2 (1980).

    Watson, P.R. and Somorjai, G.A. I J.Catal. 72, 347

    ( 1981).

    Watson, P.R. and Somorjai, G.A. I J .Catal. 74, 282

    (1982).

    Ichikawa, M. and Shikakura, K. in "Proceedings of the 7th International Congres on Catalysis", Tokyo 1980

    (T. seyana and K. Tanabe, Eds.), part A, p. 925,

    32

  • Chapter 2 page

    E1sevier, Amsterdam.

    10. Wi1son, T.P., Kasai, P.H. and E11gen, P.C., J.Cata1.

    69, 193 (1981).

    11. Rice, C.A., Wor1ey, s.o., Curtis, c.w., Guin, J.A. and Tarrer, A.R., J.Chem.Phys. 74, 6487 (1981).

    12. Wor1ey, s.o., Rice, C.A., Mattson, G.A., Curtis, c.w., Guin, J.A. and Tarrer, A.R., J.Chem.Phys. 76, 20 (1982).

    13. Wor1ey, s.o., Rice, C.A., Mattson, G.A., Curtis, c.w., Guin, J.A. and Tarrer, A.R., J.Phys.Chem. 86, 2714 (1982).

    14. Primet, M., J.C.S.Farad.Trans.II, 74, 2570 (1978).

    15. Yates, O.J.C., Murre11, L.L. and Prestridge, E.B.,

    J.Cata1. 57, 41 (1979).

    16. Tauster, S.J., Fung, s.c. and J.Am.Chem.Soc. 100, 170 (1978).

    Garten, R.L.,

    17. Tauster, S.J., Fung, s.c., Baker, R.T.K. and Hors1ey, J.A., Science 211, 1121 (1981).

    18. Baker, R.T.K., Prestridge, E.B. and Garten, R.L.,

    J.Catal. 50, 464 (1979) ..

    19. Huizinga, T. and Prins, R., J.Phys.Chem. 85, 2156

    {1981).

    20. Uchijima, T., Herrmann, J.M., Inoue, Y., Burwe11 jr.,

    R.L., Butt, J.B. and Oohen, J.B., J.Cata1. 50, 464

    33

  • Chapter 2 page

    (1977).

    21. Kobayashi, M., Inoue, Y., Takahashi, N., Burwell . ,

    R.L., Butt, J.B. and Cbhen, J.B., J.Catal. 64, 74

    (1980).

    22. Nandi, R.K., Georgopoulos, P., Cohen, J.B., Butt,

    J.B. and Burwell jr, R.L., J.Catal. 77, 421 (1982).

    23. Nandi, R.K., Molinaro, F., Tang, C., Oohen, J.B.,

    Butt, J.B. and Burwell jr., R.L., J.Catal. 78, 289

    (1983).

    24. Robertson, S.D., McNicol, B.D., de Baas, J.H., Kloet,

    s.c. and Jenkins, J.W., J.Catal. 37, 424 (1975).

    25. Jenkins, J.W., McNicol, B.D. and Robertson, S.D.,

    Chem.Tech 7, 316 (1977).

    26. Hurst, N.w., Gentry, S.J., Jones, A. and McNicol,

    B.D., Catal.Rev. Sci.Eng. 24, 233 (1982).

    27. Boer, H., Boersma, N.F. and

    Rev.Sci.Instr. 53, 349 (1982).

    Wagstaff, N.,

    28. Crucq, A., Degols, L., Lienard, G. and Frennet, A., ·

    Acta Chim. Acad.Sci.Hung. 1982, 111.

    29. Sinfelt, J.H. and Yates, D.J.C., J. Catal. 10, 362

    (1968).

    30. Wanke, S.E. and Dougharty, N.A., J.Catal. 24, 367

    (1972).

    33. Van 't Blik, H.F.J., van Zon, J.B.A.D., Huizinga, T.,

    34

  • Chapter 2 page

    Vis, J.C., Koningsberger, D.C. and Prins, R.,

    J.Phys.Chem. 87, 2264 (1983).

    34. Yao, H.C., Japar, s. and She1ef, M., J.cata1. 50, 407 (1977).

    35

  • Chapter 3 page

    Chapter 3

    The Morphology of Rhodium Supported on Ti0 2 and Alz 03 as

    Studied with Temperature Programmed Reduction-Oxidation

    and Transmission Electron Microscopy.

    Supported Rh/Al2o 3 and Rh/Ti0 2 catalysts with varying

    roetal loadings have been investigated with chemisorption

    and temperature programmed reduction and oxidation.

    Hydragen chemisorption shows that all the rhodium on

    Al2o 3 is well (H/Rh > 1 for loadings < 5 wt% and H/Rh > 0.5 up to 20 wt%), dispersion on Tio 2 is much

    lower. TPR/TPO shows this is due to the growth of two

    different kinds of rhodium /Rh2o 3 on Ti02 ; one kind

    easily reduced/oxidized, showing high dispersion, the

    ether kind harder to reduce/oxidize, showing lower

    dispersion. TEM has shown that the first kind of Rh 2o 3 consists of flat, raftlike particles, the secend kind of

    spherical particles.

    3.1. Introduction.

    Over the past years rhodium has been gaining importance

    in catalytic chemistry. Not only is rhodium widely

    recognized as the best catalyst to promate the reduction

    of NO in three way catalysts (1-3), it also takes a

    special place in the conversion of synthesis gas, since

    its product range can include oxygenated products

    (alcohols, aldehydes, acids) besides hydrocarbons (4-11).

    Various workers have tried to influence the selectivity

    36

  • Chapter 3 page

    and activity of the supported rhodium catalysts in syngas conversion via a special· preparatien (4,5), via additives

    (6-8) and via control over the oxidation state of the rhodium in the catalysts (9-11).

    Ichikawa deposited rhodium-carbonyl clusters on various supports (4, 5) and after pyrolysis of the clusters, he found a large range of selectivities towards oxygenates in the hydragenation of co. He explained this by the acid-bas& properties of the supports (12).

    Somorjai c.s. tried to hydrogenate CO over unsupported rhodium and rhodium foil, and found only hydrocarbons,

    unless they preoxidized the metal (9). H2 /co atmosphere led to the reduction to rhodium metal and the production

    of hydrocarbons, whereas Rh203.5H20 was better resistant towards reduction and produced oxygenates for a long time ( 10).

    Where the influence of additives or mixed oxides worked

    out into the right direction of enhanced oxygenate

    production, some workers traeed this to the presence of rhodium ions in the surface (11,13). The supposed presence of rhodium ions has been a point of discussion

    for quite some time. Several authors claimed it to be

    present in mono-metallic rhodium catalysts. W?rley et al.

    investigated a.o. a 0.5 wt% Rh/Al203 catalyst via infrared spectroscopy of adsorbed CO (14-16), and ascribed several infrared bands to CO molecules bound to

    isolated Rh(I) sites. This is in contrast with earlier findings by D.J.C. Yates et al. (18). They investigated

    some Rh/Al 2o 3 catalysts with electron microscopy and found rhodium to be present as metallic rafts. They did

    notmention any isolated Rh(I) sites, but, as Worley and

    coworkers stipulated themselves, the two groups used very

    different methods of preparatien of the samples.

    In all cases it seems obvious that the support plays an

    important role in either bringing or keeping the metal in

    37

  • Chapter 3 page

    a certain state of (un)reactivity. A special example of

    such an interaction between metal and support has been

    described by Tauster et al. (19, 20) and is now known as

    Strong Metal Support Interaction (SMSI). Supported noble

    and transition metals such as Ft, Rh, Ru etc. are

    normally capable of chemisorbing a.o. H2 and CO. But if

    they are supported on oxides as Ti0 2 , V205 and Nb205, and

    if they have been reduced at high temperatures (e.g. 773

    K), this chemisorption capability is greatly diminished.

    This SMSI phenomenon has been related to the occurrence

    of lower oxides of the supports (21), although these are

    known to be formed at lower temperatures than necessary

    to cause SMSI (22) and the exact nature of the

    interaction still remains unclear. The SMSI state can be

    destroyed according to Tauster c.s. by oxidation at

    elevated temperatures, followed by

    reduction (473 K): this procedure

    chemisorption behaviour (19).

    low ternperature

    restores normal

    All of the above rnentioned phenomena have to do with

    one common property: The oxidation-reduction behaviour of

    supported rhodium catalysts. We therefore decided to

    study a nurnber of Rh/Al20 3 and Rh/Ti02 catalysts with

    varying rnetal loadings (see Experimental). A1203 was

    chosen because it is known as a support giving good

    dispersions and stable catalysts, and Ti02 because it is

    known to exhibit SMSI. We varied the metal loading to

    create a variatien in partiele size, to see if and how

    oxidation-reduction and SMSI behaviour are influenced by

    partiele size.

    Befare we come to the experimental techniques we used,

    we want to introduce aother item: passivation. It is

    obvious that reduced catalyst systems cannot simply be

    removed from the reduction reactor and then be stared in

    air for later u se~ we stabil i ze the me tal surface by

    applying a layer of oxygen upon the metal particles in a

    38

  • Chapter 3 page

    controlled way (see Experimental): We passivate the

    catalysts. Although a simple low temperature reduction is

    sufficient to remove the passivatien oxygen again (as

    will be shown), some authors have given attention to the

    state the catalysts are in after storage in air. One can

    see this as a prolonged passivation, but without the

    precautions we take to prevent uncontrollable effects

    upon the first contact between air and the reduced roetal

    catalyst. So Burwell Jr. et al. used Wide Angle X-ray

    Scattering, Extended X-ray Absorption Fine Structure,

    hydragen chemisorption and hydrogen-oxygen titration to

    characterize their supported Pt and Pd catalysts (23-26),

    and they found their catalysts to be oxidized to a great

    extent after prolonged storage in air.

    We will show that a good insight in all these matters

    can be gained with the aid of Temperature Programmed

    Reduction and of Temperature Programmed Oxidation (TPR

    and TPO), supported by chemisorption measurements. TPR as

    a characterization technique was presented by Jenkins,

    Robertson and McNicoll in 1975 (27,28) and has been used

    extensively in the past few years. The development has

    been reviewed by Hurst et al. (29). The technique allows

    one to get (semi)quantitative information about the rate

    and ease of reduction of all kinds of systems, and once

    the apparatus has been built the analyses are fast and

    relatively cheap. We used an apparatus as described by Boer et al. (30), which enabled us to extend the analyses

    to Temperature Programmed Oxidation, and to gather

    information about the rate and ease of oxidation as well.

    3.2. Experimental.

    Tio 2 (anatase, Tioxide Ltd., CLDD 1367, surface area 20

    m2/g, pore-volume 0.5 cm3/g) and Y-Al203 (Ketjen, 000-

    l.SE, surface area 200 m2/g, pore-volume 0.6 cm3/g) were

    39

  • ctmpter 3 page

    wt% Rh

    2.3

    4.6

    8.5

    H/Rh

    1.53

    0.96

    0.81

    0.67

    0.54

    Table 1. Hydragen chemisorption of

    the Rhjy-Al203 (RA) catalysts.

    11.6

    20.0

    H/Rh Table 2. Hydrogen chemi-wt% Rh

    LT( 1 J HT(ZJ sorption of the Rh/TiOZ

    (RT) catalysts. 0.3 1.10 0.00

    0.7 0.61 0.01

    1.0 0.41 0.01

    2.0 0.35 0.01

    3.2 0.22 0.02

    8.1 0.12 0.04

    ( 1 ) : reduced in situ at SZ3 K. ( 2) : reduced in situ at 773 K.

    impregnated with aqueous so1utions of RhCl3.xH20 via

    the incipient wetness technique to prepare the catalysts.

    Their characteristics are presented in Table 1 and Table

    2. The catalysts will be denoted from now on as RT

    (Rh/Ti02 ) and RA (Rh/Alz03 ) catalysts, foliowed by the

    metal loading. After impregnation the catalysts were

    dried in air at 355, 375 and 395 K for 2 hours

    successively, foliowed by direct pre-reduetion in flowing

    Hz at 773 K for one hour. Prior to removing the

    from the reduction reactor they were

    temperature by replacing the Hz flow by Nz and

    40

  • Chapter 3 page

    subsequently slowly adding 0 up to 20 %. Then the catalysts were taken out of the reactor and stared for

    further use.

    In the TPR-TPO apparatus used a 5 % H2 in Ar or a 5 %

    o 2 in He flow can be directed through a microreactor, which is connected to a temperature programmer. H2 or 02

    consumption is being monitored continuously by means of a

    Thermal Oonductivity Detector (TCD). A typical sequence

    of experiments is as fellows:

    - the passivated or oxidized sample is flushed under Ar

    at 223 K

    - Ar is replaced by the Ar/H2 mixture, causing at least

    an apparent H consumption (first switch peak)

    the sample is heated under Ar/H 2 flow with 5 K/min to

    873 K

    after 15 min at 873 K, the sample is caoled down with

    10 K/min to 223 K

    - the reduced sample is flushed with Ar

    - Ar flow is replaced by the Ar/H2 mixture once more, now

    causing only an apparent H consumption (second switch

    peak).

    An identical sequence is followed during TPO, so the

    final oxidation temperature in TPO is also 873 K, unless

    stated otherwise.

    The switch peak procedure deserves some closer

    attention. The streng signal we call the first switch

    peak is due mainly to the displacement of Ar by Ar/H2 in

    the reactor, but in some cases real hydragen consumption

    might take place, even at 223 K. Therefore we repeat the

    whole procedure after the TPR has been performed: In that

    case the catalyst has been reduced and caoled down to 223

    K, and as a consequence it is covered by hydrogen. Then

    we replace the Ar/H 2 by pure Ar. Subsequently we switch

    back to Ar/H2 . Since we cannot expect any hydragen

    consumption from the reduced, hydragen covered sample at

    41

  • Chapter 3 page

    this time, the resulting second switch peak will be due

    solely to the displacement of Ar by Ar/H2· So the difference between the first and second switch peak

    reveals the real hydragen consumption at 223 K, if there

    is any.

    The reactions that might take place during TPO and TPR

    are:

    4 Rh 0.75) [1]

    The quantities between brackets are the hydragen or

    oxygen consumptions in TPR or TPO expected for reduction

    of bulk Rh 2o 3 or formation of this very material (apart from chemisorption of any kind). In a standard experiment

    a TPR is done on a passivated catalyst, followed by TPO

    (on the now reduced catalyst), followed by TPR (on the

    now oxiclized catalyst).

    1.2

    • 0.8

    0.4

    4 8 12 16 20

    wt'l. Rh~

    Figure 1. Hydragen chemisorption of Rh/y-Al 2 0 3 catalysts as a function of roetal loading. T(red) is 773 K.

    42

  • Chapter 3 page

    Chemisorption measurements were carried out in a

    conventional glass apparatus after reduction of the

    passivated catalysts at 773 K in flowing Hz for 1 hour,

    followed by evacuation at 473 K for one hour. After

    hydragen admission at 473 K desorption isotherms were

    measured at room temperature. Fbr the RT series 773 K

    (High Temperature} reduction will induce SMSI, and so the

    RT catalysts were also reduced at 523 K (Low Temperature)

    to study their normal chemisorption behaviour. In

    measuring the desorption isotherms desorption became only

    noticeable at pressures below 200 torr (l torr 133.3

    N/m2} so we believe that the chemisorption value above

    that pressure is representative of monolayer coverage

    (cf. Frennet. c.s. (31}).

    Transmission Electron Microscopy was carried out on a

    Jeol 200 ex top entry stage microscope. Photographs were taken at a magnification of 430,000 times, and then

    enlarged further photographically to a final magnifi-

    cation of 1,290,000 times. TEM measurements were done on

    samples pre-oxidized at 900 K. This temperature was

    chosen because, as TPO measurements will show, for some

    samples this temperature is necessary to cause total

    oxidation. Samples were prepared by applying a slurry of

    the catalyst in alcohol onto a coal coated capper

    and evaporating of the alcohol.

    Metal loadings were established for the passivated

    samples spectrophotometrically.

    3.3. Results.

    3.3.1. Hydragen Chemisorption.

    The hydragen chemisorption data as given in Table 1 for

    the RA series (Rh/Al2o 3 } are represented graphically in Fig. 1. In a similar way the data from Table 2, for the

    43

  • Chapter 3

    o.a

    0.6

    0.4

    0.2

    2 4 8 wt% Rh..._

    s:: a:: ...... :I:

    0.004

    0.002

    page

    Figure 2. Hydragen chemisorption of Rh/Ti0 2 catalysts as a function of metal loading. T(red) is 473 k (LT) or 773 K (HT).

    RT series (Rh/Ti0 2 ), are presented in 2. Fbr

    Rh/Al 2o 3 , the H/Rh value drops below 1.0 somewhere around

    5 wt % loading but H/Rh is still above 0.5 at 20 wt % (which catalyst had to be prepared via two successive

    impregnation and drying steps). Fbr Rh/Tio 2 reduced at

    523 K H/Rh decreases much faster with increasing metal

    loading, and drops below 1.0 before the metal lóading

    reaches 0.5 wt %. Of course we must keep in mind ~he 10 times larger surface area of the alumina, but it is

    obvious that anatase is less capable of stabilizing small

    rhodium particles than alumina. Fbr Rh/Ti02 reduced at

    773 K, some hydrogen chemisorption is still measurable at

    high metal loadings, but the measured values are of the

    order of magnitude of experimental error.

    The attentive reader will have noticed by now that the

    catalysts must have been already in the SMSI state after

    44

  • Chapter 3 page

    the pre-reduetion at 773 K. The fact that they do show normal chemisorption behaviour after 523 K reduction implies that the passivation treatment they received has nullified SMSI.

    The reduction-oxidation behaviour of a selected number of these catalysts, RA 2.3, 4.6 and 20.0 and RT 0.3, 1.0, 3.2 and 8.1, will be presented below.

    3.3.2. TPR/TPO of the RA series.

    The Temperature Programmed passivated RA catalysts

    Reduction is shown

    profile in Fig.

    of 3.

    the The

    horizontal axis shows the temperature, the vertical axis hydro~en consumption in arbitrary units.

    ) ~ ~

    473 673 873 T(K)~

    Figure 3. Temperature Programmed Reduction ot passi-vated 2.3 wt% Rh/y- Al 03.

    All three catalysts show a hydrogen consumption maximum below 300 K, followed by a slight desorption. Hydrogen consumption decreases from 1.33 H2/Rh for RA 2.3 (which is almost enough to account for the reduction of

    stoechiometric Rh 20 3), to 0.49 H2/Rh for RA 20.0 (average

    45

  • O'l.apter 3 page

    oxidation state of the rhodium in this case was +1).

    V r---r---r----.

    lÎ r----. ---~ 1'"--..

    --v- ! ['--._ 273 473 673 !173

    T(K)....,_ Figure 4. Temperature Programmed Oxidation of reduced

    Rh/y-Al 20 3 ; 2.3 wt% (a), 4.6 wt% (b) and 20.0 wt% (c).

    The subsequent TPO profiles show more difference. For

    all three catalysts Fig. 4a, b, c), oxygen consurnp-

    tion starts at 223 K, that is in the switch peak. For RA

    2.3 the oxygen consurnption rises at the beginning of the

    temperature ramp, to reach a maximum at 300 K, and to

    fall down slowly towards higher temperatures. Total

    oxygen consumption mounts up to 0.65 o2/Rh. The behaviour of RA 20.0 (Fig. 4c) is quite different. After some

    oxygen consumption i.n the switch peak, oxygen consumption

    46

  • Chapter 3 page

    keeps a low level for several hundreds degrees of

    628 K. o2/ 4.6 was an

    temperature raise, to reach a maximum at

    rhodium is 0.70. The behaviour of RA

    intermedia te one (Fig. 4b). Integration of the oxygen

    consumption signal proved difficult, because of the small

    therrnal conductivity of 0 2 , and due to the small sample

    sizes (typically 50-75 micromale of metal).

    lJ \ i'.. -273 473

    i

    i

    673

    i

    673 T(K)__"._

    Figure 5. TPR of oxidized 20.0 wt% Rh/y-Al 2 o 3 •

    The TPR profile of the oxidized catalysts is shown in

    Fig. 5, and is alike for all three of them. One sharp and

    well-defined peak at about 340 K, with an H2/Rh value

    ranging frorn 1.3 (RA 20.0) to 1.6 (RA 2.3). In all three

    cases we assurne we have to do with the reduction of

    supported Rh2o 3• Unsupported Rh203 in showed a reduction peak at 400 K, while

    our apparatus

    bulk rhodium

    roetal only started to becorne bulk-oxidized above 870 K.

    3.3.3. TPR/TPO of the RT series.

    Fig. 6 shows the TPR profiles of the passivated

    47

  • Chapter 3

    ) ( ~

    \

    \

    273

    I

    473 673 873 T(K)~

    page

    Figure 6. TPR of passivated Rh/Ti0 2 ; 0.3 wt% (a), 1.0 wt% (b), 3.2 wt% (c) and 8.1 wt% (d).

    48

  • Chapter 3 page

    catalysts RT 0.3, RT 1.0, RT 3.2 and RT 8.1. With

    increasing rnetal loading, in other words decreasing H/Rh

    values, we see how the reduction peak disappears into the

    switch peak at 223 K. Fbr RT 0.3 the reduction peak is

    still cornpletely separated from the switch peak, with a

    maximum at 273 K and an H2 / rhodium of 0.56, for RT 8.1

    the reduction peak has merged with the switch peak and

    the hydragen consumption has d~opped to H2/Rh =0.25. We

    believe the explanation for this lies in the fact that

    --~

    / I-

    --- - -----""' 1'---

    _.-/ ~ L

    273 473 673 873 T(K)~

    Figure 7. TPO of reduced Rh/Ti0 2 ; 0.3 wt% (a), 1.0 wt% (b), 3.2 wt% (c) and 8.1 wt% (d).

    49

  • Chapter 3 page

    with decreasing H/Rh (and for RT 1.0 the H/Rh ratio is

    nat more than 0.41, still lower than for RA 20.0) the

    metal particles keep a metallic care

    Hydragen molecules can diffuse

    upon passivation.

    through the outer

    passivated layer of the particle, reach the metallic care

    and dissociate there to provide atomie hydragen for an

    easy reduction of the oxide layer at low temperature.

    We find a support for this idea in Fig. 7 showing the

    TPO profiles of the respective catalysts. All catalysts

    show some oxygen consumption as soon as the temperature

    ramp has been started. The two low-loaded catalysts, RT

    0. 3 and RT l. 0 ( 7a and 7b) have an ear 1 y consumption

    maximum around 300 K, likeRA 2.3, but also show distinct consumption around 600 K, while RA 20.0 had a maximum

    around 620 K. In Fig. 7c, RT 3.2, we can distinguish

    three areas of oxygen consumption apart from the switch

    peak: Around 350 K, 600 K and a new maximum at 770 K.

    This consumption at 770 K is dominant in Fig. 7d, the TPO

    of the reduced RT 8.1 catalyst. At first glance we can

    attribute the low temperature oxygen consumption to

    chemisorption and the high temperature oxygen consumption

    to thorough oxidation, but we will come back to this in

    the Discussion. Anyway, it is this last observation, the

    high temperature needed for thorough oxidation, that

    confirms the presence of a metallic care in larger

    particles after passivation. Oz/Rh is abaut 0.7 for all

    RT catalysts. In Fig. 8 we see

    catalysts, and a

    here. RT 0.3 and RT

    maximum, 330-340

    the TPR profiles of the oxidized

    very interesting phenomenon shows up

    1.0 show only one clear consumption

    K, Hz/Rh=l.20, like the RA catalysts

    (Fig. 5). But RT 3.2 and RT 8.1 have a secend consumption

    maximum at 385-400 K, abaut the temperature where

    unsupported bulk Rh2 0 3 reduces. Taken bath peaks

    together, H2/Rh is about 1.70 for RT 3.2 and 1.41 for RT

    8.1, so bath peaks can without any doubt be ascribed to

    50

  • Chapter 3

    (\

    _)

    ""' _.... 1----v---

    lJ" v--... -

    \..

    v \ ----...._

    _./ 1--..__

    11

    '--IJ ) \

    273 473 673

    Figure 8. TPR of oxidized Rh/Ti0 2 ; 0.3 wt wt% ( b) , 3. 2 wt% ( c) and 8. 1 wt%

    873 T(K)....,_

    page

    (a), 1. 0 (d).

    51

  • Chapter 3 page

    the reduction of Rh2o 3. To be more specific, the low temperature TPR peak must belang to well-dispersed, ill-

    defined Rh 2o 3, which is also easily formed, while the high temperature TPR peak belongs to the reduction of

    bulklike, crystalline Rh2o 3, which can only be formed by the high temperature oxidation. This was proven earlier

    (32) by halting a TPO experiment at a temperature

    intermediate between the two maxima at 600 and 770 K. A

    subsequent TPR proved to miss the second reduction peak,

    so the 770 K TPO peak is connected to the 400 K TPR peak,

    and the low temperature TPR and TPO peaks must, as a

    consequence, be connected too.

    I ./ \ ----- !'---. 273 473 673 873

    I

    I v--~ ~ --373 S73 773 973

    T (K)......_

    Figure 9. 1.0 wt Rh/Ti0 2 , wet reduced; TPR of oxidized system (a) and TPO of reduced system (b). Note the temperature axis in TPO.

    52

  • Chapter 3 page

    In all cases, af ter a slight desorption of H2, H2 conswnption rises again above baseline, shawing two maxima around 600 and 740 K. We think this must be attributed to reduction of the support in the neighbourhood of the metal; tli.e bare support does not reduce below 800 K.

    3.3.4. TEM measurements.

    To make the difference between the R~o3 reducible at 330 K and the Rh 2o 3 reducible at 400 K more clear, we

    reduced part of the impregnated RT 1.0 batch at 773 K

    without drying, a method already used by Burwell c.s.

    (24) to prepare catalysts with law dispersions. After

    oxidation at 900 K for 1 hr we performed a TPR, shown in

    Fig. 9a. One single reduction peak for Rh203, at 410 K

    was observed. The subsequent TPO, shown in Fig. 9b, run

    this time up to 973 K, shows that oxidation proceeds

    readily only at about 900 K. The H/Rh value of the catalyst, after the 900 K oxidation and a 523 K reduction in situ, was 0.20.

    We examined this

    comparable RT 1.0

    catalyst,

    (Fig.

    RT

    8b,

    Transmission Electron Microscopy

    prepared by applying a slurry of

    l.O(wet), and the

    7b) catalyst with

    (TEM). Samples we re

    the at 900 K oxidized

    catalysts in absolute alcohol on a coal coated grid, and

    evaporating the alcohol. TEM micrographs are shown in

    Fig. 10. Fig. lOa shows the bare support, anatase, as

    received from the manufacturer. Clearly visible is that

    the Ti0 2 particles (in the order of magnitude of 50 nm in

    diameter) are covered with very tiny seed-crystals of

    TiC2, which apparently, due to the poor sintering

    capacities of these kinds of oxides, did not get the

    chance to grow into larger particles. In Fig. lOb (RT

    1.0) we find them back as conglomerates (clustered

    together in the impregnation and drying steps of the

    53

  • Chapter 3

    1oo.&

    F i gure 10 a. T r a n sm i ss io n e lectro n mi cr o rap h of anatase T i 0 2

    page 54

  • Chapter 3 page

    Figure 10 b. Transmission electron micrographo f 1. 0 wt% Rh z03/ Ti0 2 , H/ Rh = 0.4

    55

  • Olapter 3 page

    Figure 10 c. Transmi s sion electron mLcrograph of 1. 0 wt% Rh 2 0JITi0 2 , H/Rh = 0 . 2

    56

  • 3 page

    on the Ti02 surface show the uniform density

    and srnooth outline characteristic for heavy roetal oxide

    , diameter ranging frorn 1-6 nrn. The uniform

    colour of the particles is an indication that they are

    actually flat (raftlike). In Fig. lOc (RT l.Owet) the

    Ti02 conglornerates are no langer visible. Apparently they

    got the chance during the high ternperature reduction in

    the presence of water to spread over the surface of the

    large particles. The particles we do see, are Rh2o3 . The lighter contours of the particles indicate

    that they are spherical. Their diameter is about 7 nrn.

    3.4. Discussion.

    Hydragen chernisorption has been used through the years

    by rnany workers to characterizernetal surfaces (10, 12,

    13, 17, 19, 23-26, 33, 34) and when atternpts were made to

    ca1culate roetal surface areas frorn hydragen chernisorption

    data, a hydragen-metal stoechiornetry of one was a1ways

    used. On the other hand, if CO was involved, sorne authors

    (35) chose a stoechiornetry of one, while others (18)

    preferred stoechiornetries. Frorn Figs. 1 and 2 it

    becornes clear though, that a1so H/Rh values can exceed

    unity. Fbr Rh/Ti02 this occurs for roetal loadings below

    0.5 %, and for Rh/Al2o 3 even for roetal loadings up to 5 %. Although in this study only two cata1ysts occur with

    an H/Rh va1ue above unity, we have reported about other

    ultra-dispersed systerns elsewhere (32, 36). And in our

    opinion, if one accepts a roetal atorn, such as rhodium, to

    adsorb two or more CO molecules, one should shou1d not

    reject the idea of that same atorn adsorbing more than one

    hydragen atorn. So we think that a1so in those cases where

    H/Rh exceeds unity, the hydragen chernisorbed does not

    exceed a rnono1ayer (cf. Frennet (31)) and is all bound to

    the rnetal. A consequence of this is the irnpossibility to

    57

  • Chapter 3

    calculate a partiele size

    highly dispersed systems,

    page

    fram chemisorption data for

    since there is no way of

    calculating the number of rhodium surface atoms. Fbr

    larger, well defined particles as w found in RT l.O(wet),

    with an H/Rh of 0.2, the calculated partiele size is 6 nrn

    which is in good accordance with our TEM measurements

    (Fig. lOc).

    The H/Rh values measured for Rh/Ti02 after reduction at

    773 K (Fig. 2), that is with rhodium in the SMSI state,

    seem to show a tendency to increase, but the values

    measured are still that small that we do not want to draw

    any conclusions frorn that without further study.

    Fram all TPR/TPO measurements presented here the

    following picture comes up. Rhodium on a support can

    occur in either a dispersed farm, or in a bulklike farm.

    The dispersed farm is easy to reduce to the metal, giving

    a reduction peak in TPR around 340 K. It is also easily

    oxidized, which manifests itself in TPO's such as Figs.

    4a, 7a and 7b, and from the fact that upon passivatien

    the oxidation state of the Rh can almast reach +3 (cf.

    Fig. 3, H2/Rh=l.33). That reaction with oxygen is

    vehement, even at room temperature, shows also in the

    fact that prolonged passivatien breaks SMSI (if not, one

    would not have seen any hydragen consumption at all after

    the low temperature reduction, since the RT catalysts

    would still be in the SMSI state induced by the 773 ~

    pre-reduction). The exact assignrnent of the three TPO

    peaks is difficult since the literature does not provide

    much material about the mechanism of oxidation of

    support

  • Chapter 3 page

    follawing logarithmic rate equations, and leading to an

    oxide film with a thickness from 2 (38) to 100 (39, 40)

    rum. Finally, Wagners oxidation theory (41) describes how

    the oxidation process goes on, rate-controlled by volume

    diffusion of the reacting ions and/or of electrens

    through the growing oxide scale, leading to parabalie

    rate equations. We can imagine that analogous models can

    describe the oxidation processes in supported metal

    particles as well, dependent on their sizes.

    It is very clear that no matter how the rhodium is

    present on the support, dispersed or not, oxidized or

    passivated, reduction to metallic rhodium is complete at

    400 K, that is