Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H....

download Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garofalini -- Molecular dynamics simulations of α-alumina and γ-alumina surfaces.pdf

of 12

Transcript of Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H....

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    1/12

    Surface Science 295 (1993) 263-274

    North-Holland

    . .:.:.:i........:.:;.:/.......,:,:.> .:,:.>:.:_,.,

    p::~:l:~::::::~:l:.:................

    l......... . . . . . . .......................~.~....r).::; ::::.

    i.

    li::::z::.

    . >;:,~,:z:i

    surface science

    Molecular dynamics simulations of a-alumina and y-alumina surfaces

    S. Blonski 1 and S.H. Garofalini

    Depart ment of Cerami cs and Int erfacial M olecular Science Laboratory , Insti tut e for Engineered M aterials, Rutgers U nioersit y,

    P.O. Box 909, Piscataw ay, NJ 08855-0909, USA

    Received 18 January 1993; accepted for publication 15 June 1993

    Molecular dynamics simulations of crystalline aluminum oxide were performed for cu-Al,O, and y-Al,O, phases. Both bulk

    crystals and surfaces of each phase were studied. For each of the surfaces, several possible atomic terminations were examined and

    surface energies, density profiles, and atom configurations have been calculated. It was found that due to processes of surface

    relaxation and reconstruction some terminations of the a-alumina surfaces become more likely to appear. For y-alumina, the

    occurrence of cation vacancies in the crystal structure has a significant influence on surface morphology. On the surfaces,

    additional active sites were observed which are not predicted by idealized models which omit vacancies.

    1. Introduction

    Aluminum oxides have a great number of tech-

    nological applications. While a-alumina is mainly

    seen as a structural, optical and electronic mate-

    rial, y-alumina is usually used as a catalytic sup-

    port. Because surface properties of the crystals

    affect the successful application of these alumi-

    nas, the surfaces have been an object of a number

    of theoretical and experimental studies. For (Y-

    alumina, only one possibility of surface termina-

    tion has been considered for each surface studied

    by theoretical modeling. However, experimental

    studies show that there exist different terminating

    layers on a-alumina surfaces [1,2]. Thus, to fill

    this lack of knowledge, molecular dynamics simu-

    lations of a-alumina surfaces with different ter-

    minating layers of atoms were performed. Con-

    versely, for y-alumina, the concept of different

    terminations has been known and examined pre-

    viously. However, theoretical models of y-alumina

    surfaces are usually based on an idealized struc-

    ture of the crystal, with an excess of aluminum

    atoms [3]. In this study, an extended approach for

    On leave from the Department of Applied Physics, Techni-

    cal University of Gdansk, Poland.

    y-alumina crystals was used to find more realistic

    structures of the surfaces. Due to the nature of

    the y-alumina crystals, experimental observations

    of their surfaces have not yet been done. There-

    fore, simulations can show for the first time how

    the surfaces look and should allow for a better

    understanding of the properties of y-alumina.

    Both (Y- and y-aluminas were studied using the

    same model to allow for comparison between

    properties and behavior of both materials.

    2. Computational procedure

    Constant-volume simulations were performed

    with a fifth-order Nordsieck-Gear algorithm used

    to integrate Newtons equations of motion with a

    time step of the integration of 0.2 fs. The total

    potential energy of the system is composed of

    contributions from two- and three-body interac-

    tions. The two-body part is a modified Born-

    Mayer-Huggins form, given as

    Qij =Aij exp( -rij/pij)

    + ( 4i4je2/rij) erfc ij/Pij) >

    0039-6028/93/ 06.00 0 1993 - Elsevier Science Publishers B.V. All rights reserved

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    2/12

    264

    S. Blonski, S.H. Garof ali ni / M olecular dynamics simul ati ons of a- and y-al umina surfaces

    where qi and q, are the ionic charges, rij is the

    separation distance between ions i and j, e is the

    elementary charge, and erfc is the complemen-

    tary error function 0[4]. The softness parameter,

    pCj, is equal to 0.29 A for all pairs of the ions. The

    values for the adjustable parameters, Aij and pIj,

    are as follows:

    A o_o = 0.0725 fJ,

    po_o = 2.34 A,

    A ,_,,=0.2490J, /3O_A, = 2.34 A,

    A *,_*, 0.0500J,

    P/,_A, = 2.35 A.

    Similar values of the parameters were used in the

    studies of sodium aluminosilicate glasses [5,61,

    but for the reason of obtaining a rational value of

    the pressure in the simulated bulk crystals, the

    parameters were modified for the current studies.

    The main difference is for the A,,_,,arameter,

    but even in this case the force between two ions

    is changed by no more than 6% at the distances

    occurring in the simulations. Nevertheless, the

    adjustment of the parameters was necessary, be-

    cause the previous values were developed for an

    amorphous system in which Al formed only a

    fraction of the total number of cations and all of

    the Al ions were biased toward tetrahedral coor-

    dinations so as to test ideas regarding glass prop-

    erties [51.

    function for the triplet with an oxygen ion at the

    vertex has a minimum at about 109 which is

    appropriate for the tetrahedral coordination, To

    take into account two possible coordinations of

    the Al atoms in the alumina crystals, the different

    angular function is used for the triplets with an

    aluminum ion at the vertex. That function has a

    broad minimum for 0 in the range from 90 to

    110 as well as another minimum for f3 = 180.

    This allows Al atoms to be both tetrahedrally and

    octahedrally coordinated. The adjustable paramc-

    ters, h j r k , Yi , ,

    and R,j,ave the following values:

    A ,_o_/,, = 0.001 fJ,

    A,~,,_, = 0.024 fJ,

    yo_*, = 2.0 A,

    y&-o = 2.8 A,

    R

    omA, = 2.6 A,

    R,,_o = 3.0 A.

    The same values of the three-body parameters

    were used in the previous studies of glasses.

    The three-body interactions imposed on all of

    the Al-O-Al and O-Al-O triplets have the fol-

    lowing functional form:

    q j i l k = j i k exP[ Yi j / i , - Rr j )

    +Y i k / t T i k -Rzk ) ]a j i k ,

    if r i j < Rij and

    r ik < Rik;

    Pjik 0,

    if

    r r l 2

    Rtj or

    rik 2 R,k ;

    with the angular part, fljik, given by

    fijik = (cos O,ik + l/3)*,

    for AI-O-Al;

    finirk = [(co, ejik + l/3) sin 19,~~ os tijik]*,

    for O-Al-O;

    The simulations were run for 50000 time steps

    each with the initial 5000 steps being used for

    temperature equilibration. All the reported simu-

    lations have been performed at 300 K. After the

    bulk crystals were simulated, the desired surfaces

    were exposed and surface simulations were per-

    formed in the way described by Garofalini [4l. A

    surface was created by removing periodic bound-

    aries in one dimension, while keeping them in the

    other two dimensions. Simultaneously, atoms in

    the layers most apart from the surface were im-

    mobilized, so that they retained their original

    bulk-like configuration. Samples consisting of

    2560 to 3600 atoms were simulated, respiting in

    surface dimensions of appr?ximately 25 A X 25 A

    and a height of about 50 A. To allow for move-

    ment of as many atoms as possible, the thickness

    of the layer of immobile atoms was chosen to only

    slightly exceed the interaction cut-off distance

    (5.5 A,.

    The structures of the simulated crystals were

    characterized in terms of radial distribution func-

    tions (RDFs) of the atomic positions. The partial

    radial distribution functions were calculated for

    each pair of ion types from the formula [7 :

    g i , t r ) =W j r ) /N o r ) 7

    where ejik is an angle formed by the ions j, i, and

    where N ,, r) denotes the number of ions of type j

    k with the ion

    i

    placed at the vertex. The angular

    in a shell between

    r - Ar/2

    and

    r + Ar /2

    around

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    3/12

    S. Blonski, S.H. Garofal ini / M olecular dynamics simul ati ons of a- and y-alumi na surfaces

    265

    an ion of type i. The average number of atoms,

    N&r), in the same shell in an ideal gas at the

    same number density p is

    N,(r) = (%/3)[(

    r + Ar/2)3 - (r - Ar/2)3].

    The shells with the thickness Ar = 0.01 A were

    used. The number density p was calculated by

    dividing the total number of ions in the simulated

    crystal by the volume of the simulation cell. The

    total radial distribution functions were obtained

    by summation of the partial RDFs over all pairs

    of ion types.

    Number density profiles in the direction per-

    pendicular to a surface were calculated by count-

    ing the

    surface.

    w each

    cell.

    number of ions in slabs parallel to that

    The slabs have the thickness of about 0.1

    depending on the size of the simulation

    3. Structures of bulk crystals

    The initial structure of a-alumina was defined

    according to Wyckoff [8], with the lattice parame-

    ters given by Newnham and de Haan [9]. Fig. 1

    shows the radial distribution function obtained

    from the simulation of bulk cy-Al,O, composed

    of 3600 atoms, i.e. 120 crystallographic cells with

    hexagonal symmetry. For comparison, the static

    pair distribution for the structure given by Wyck-

    5

    4

    3

    C

    xl

    2

    1

    0

    simulation

    diffraction

    2 4

    r [Al

    6 8

    10

    Fig. 1. Total radial distribution function for a-alumina: solid

    line is from the simulations, dotted lines are from the X-ray

    diffraction studies [8].

    off is also shown. The excellent agreement be-

    tween crystallographic data and simulation re-

    sults indicates that the potentials used in the

    simulations adequately describe the structure of

    a-alumina, preserving the initial configuration of

    atoms in the molecular dynamics simulations.

    The structure of y-Al,O, is still a matter of

    discussion; the common assumption being that

    -y-alumina is a defective spinel. The defects have

    to occur because the stoichiometry of Al,O, does

    not fit the spine1 structure. If all of the cation

    positions of the spine1 structure were filled by

    aluminum atoms, there would be an excess of

    aluminum atoms. Thus, some cation positions of

    the spine1 structure have to be vacant in y-

    alumina. Cation sites of two kinds appear in the

    spine1 structure: octahedral and tetrahedral. The

    question which remains is where the vacancies

    are located. Most of the experimental data sug-

    gest that vacancies occur mainly on tetrahedral

    sites [lo], but just the opposite statements can

    also be found in the literature [ll]. Nevertheless,

    it was assumed in the simulations that all vacan-

    cies are located on tetrahedral sites. Recently

    published results of other molecular dynamics

    studies of bulk y-alumina support such an as-

    sumption [12]. Those simulations were started

    from configurations which had the vacancies

    placed randomly among all cation sites of the

    spine1 structure. During the time-evolution of the

    system, nearly all of the octahedral vacancies

    were filled by aluminum atoms and the vacancies

    survived almost exclusively on tetrahedral posi-

    tions.

    In the present simulations the initial structure

    of the bulk crystal of y-Al,O, was defined as the

    cubic spine1 described by Wyckoff [8]. Locations

    of eight cation vacancies per every 3 crystal cells

    were chosen randomly. Initially, simulations of

    200 different configurations of vacancies were

    performed. The configuration with the lowest en-

    ergy was chosen for further studies. It was noted

    that among all of the configurations, differences

    in energy were no greater than 1.3 kJ/mol.

    Moreover, ali the samples of y-alumina have

    higher energies than the crystal of a-alumina.

    The lower-energy y-alumina crystals are usually

    characterized by a more uniform spatial distribu-

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    4/12

    266

    S. Blonski, S.H. Garof ali ni / M olecular dynamics simul ati ons of a- and y-alumi na surfaces

    2,

    I

    o-o

    1

    ~~~

    I,:

    : ,,

    , ,:

    ~ I /

    ~\

    \ \

    0

    Fig. 2. Partial radial distribution functions for y-alumina:

    solid and dashed lines are from the simulations with lattice

    constants 8.03 and 7.91 A, respectively; dotted lines are from

    the X-ray diffraction studies [8].

    tion of the vacancies, but the correlation between

    the energy and the distribution is rather weak.

    To obtain a reasonable value of the pressure in

    the sample, a lattice constant of 8.03 A was used,

    which is slightly greater than values usually re-

    ported in the literature (7.91 A> [ll]. Fig. 2 shows

    partial radial distribution functions (PDFs) for

    the relaxed structures obtained from the simula-

    tions of the crystals with both lattice constants.

    Although the expansion is clearly visible, the

    shapes of the PDFs are unchanged. The simu-

    lated PDFs agree well with the PDFs calculated

    for the static structure found by Verwey from

    X-ray crystallography [ 131. The most significant

    difference is an additional peak near 2.4 A in the

    simulated oxygen-oxygen PDF. However, such a

    peak occurs in the radial distribution function of

    a-alumina (both experimentally and in the simu-

    lation; see fig. 1). It originates from distorted

    oxygen octahedra which are detectable in the

    sharp diffraction spectra of a-alumina. Since the

    spectra of y-alumina are more diffuse, the distor-

    tion could have been omitted in the generation of

    the crystallographic structure of y-alumina.

    4. Surfaces of cu-alumina

    Three surfaces of a-alumina were simulated:

    (000 l),

    11 001

    and (1 120). These are the sur-

    faces which frequently occur in natural and artifi-

    cial corundum crystals and have been the subject

    of several experimental studies [14-161. From

    density profiles of the crystals with different ori-

    entation, it was observed that each of the surfaces

    can be formed by terminating at a different layer

    of atoms. Such different atomic terminations of

    surfaces could be formed during fracture of the

    bulk crystal. In previous static calculations of

    a-alumina surfaces, such a possibility was not

    taken into account [17,18]. Therefore, all possible

    terminating layers which could preserve a two-di-

    mensional periodicity on the surfaces were simu-

    lated. Surface energies obtained from our simula-

    tions are presented in table 1, along with the

    energies obtained from other theoretical studies

    of a-alumina surfaces. Experimental values of

    surface energy for a-alumina are known only for

    Table 1

    Calculated surface energies of cu-alumina

    Surface Surface energy (J m-*)

    This Ref.

    Ref.

    Ref.

    work

    [I71

    [I81 [211

    Unr ela xed surfaces

    (0001) A

    12.77

    _ _ _

    B 12.85

    C

    5.04 6.53 5.95

    6.72

    (1120) A 14.32

    _ _ _

    B 3.49 5.17 4.37

    _

    C

    14.41

    _ _ _

    (iioo) A 5.56 6.87 6.46

    5.65

    Relaxed surf aces

    (0001) A 8.04

    _ _ _

    B 2.19

    _ _ _

    C 2.04

    2.97 2.03 5.32

    (1120) A 8.39

    _ _

    B

    2.21 2.65 2.50

    _

    C

    4.17 -

    _ _

    CliOO) A 2.35 2.89

    2.23

    5.59

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    5/12

    S. Blonski, S.H. Garofal ini / M olecular dynamics simul ati ons of a- and y- alumi na surfaces

    267

    d

    E, = 8.04 J m-z

    3

    I

    I

    E, = 2.19 J m-*

    m

    I

    E

    2

    I

    I

    E, = 2.04 J mz

    1

    Height [A]

    CB

    Fig. 3. Density profiles for various terminations of the (0001)

    surface of a-alumina and for the bulk crystal in this direction.

    Dashed lines show the periodic boundaries for the bulk crys-

    tal and the limit of the immobile layer for the surfaces.

    elevated temperatures: Kingery reported 0.9 J/m2

    at 2123 K [19]. This is less than the values from

    our room temperature simulations, but it is known

    that surface energy of ceramics decreases with

    increasing temperature [20]. Although the experi-

    mental value of surface entropy, which describes

    the temperature dependence of the surface en-

    ergy, is also unknown, estimations made by Tasker

    [17] and Mackrodt [18] suggest that the calculated

    surface energies presented in table 1 are correct.

    Simulation results for the particular surfaces are

    discussed in detail in the following subsections.

    4.1,

    0001)urface

    Three possible terminations of the (0 0 0 1) sur-

    face are shown in fig. 3. Two of them (B and C)

    are terminated by the layers of aluminum atoms.

    These terminations have a surface energy signifi-

    cantly lower than the third one (A), which is

    terminated by a layer of oxygen atoms. The sur-

    face with the terminating layer C has the lowest

    surface energy among all surfaces studied for

    a-alumina. Therefore, this should be the surface

    which most frequently occurs in a-alumina crys-

    tals. Some observations have show that it is really

    the case [14]. This surface has also been chosen

    as the subject of earlier theoretical studies (see

    table 1). The surface is only slightly relaxed and

    has the symmetry of the bulk crystal. As fig. 4

    shows, for all of the terminations there exists a

    two-dimensional periodicity on the surfaces. Sur-

    faces B and C are surprisingly similar, consider-

    ing the difference between the initial configura-

    tions from which they were created. There is an

    excess of aluminum atoms on the unrelaxed B

    surface, which during relaxation shift toward the

    inside of the crystal. This causes other Al atoms

    to move deeper into the crystal and to create

    interstitial defects in the bulk. Thus. the surface

    layer

    layer

    Fig. 4. Atom configurations for the terminations of the (000 1)

    surface of a-alumina. For the layers B and C, only the

    aluminum atoms located above the surface are shown.

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    6/12

    26X S. Blonski, S.H. Garofal ini / M olecular dynamics simulat ions of a- and y-al umina surfaces

    Al atoms are the source of the crystal defects and

    their occurrence increases the surface energy, but

    they do not influence the appearance of the outer

    surface layer itself.

    Both B and C surfaces are stoichiometric and

    on both of them exist under-coordinated alu-

    minum atoms which can significantly influence

    the chemical properties of the surface. For the

    third terminating layer, the A surface, there are

    mainly oxygen atoms on the surface. This can also

    be seen as a deficiency of the aluminum atoms on

    the unrelaxed surface. Due to the field of unbal-

    anced electric dipoles, aluminum atoms from the

    Al layer adjacent to the surface move toward the

    surface during the relaxation. Some Al atoms

    from the deeper aluminum layer also move in the

    same direction and cause the A surface to be

    non-stoichiometric. In the simulated samples,

    there are 90 oxygens in each oxygen layer, there-

    fore, there should be 60 Al atoms in each alu-

    minum layer. However, fig. 4 shows that there are

    69 Al atoms in the uppermost aluminum layer of

    the termination A. The excess Al atoms originate

    from the deeper aluminum layer, hence there is

    also a deficiency of Al atoms in the deeper layer.

    Because of the similarity of the B and C sur-

    faces, the simulations show that only two differ-

    ent regions should be observed on the (000 1)

    surface. Recent experiments indicate that this

    may, in fact, be the case [16]. Two regions were

    observed in reflection electron microscopy experi-

    ments performed under different resonance con-

    ditions. Contrast reversals suggest that the atomic

    configurations in the regions are different, very

    possibly due to the surfaces terminated at differ-

    ent layers within one crystal cell. However, the

    observations are not fully understood yet. Results

    of further studies might be very important to

    clarify the problem. However, simulation results

    presented here provide a useful interpretation of

    the experimental features.

    4.2. (I I 20) surface

    Three terminations of the (1 120) surface have

    been studied. Density profiles obtained from the

    simulations are shown in fig. 5, in comparison

    with a density profile of the bulk crystal. The

    E, = 4.77 1 m-

    Height [A]

    CBA

    3

    Fig. 5. Density profiles for various terminations of the (1 1 ZO)

    surface of a-alumina and for the bulk crystal in this direction.

    termination B, which is formed when the crystal

    is split at the plane lying between two oxygen

    layers, has the lowest surface energy. Surpris-

    ingly, it is achieved with only modest relaxation of

    the surface. The surface profile remains very sim-

    ilar to the profile of the bulk crystal. It is contrary

    to the behavior of the (000 1) surface, for which

    the density profile peaks are significantly broader

    than for the bulk crystal.

    For the terminating layer A, a significant re-

    construction of the surface occurs. The depth

    involved in the reconstruction has a thickness of

    about 7 A. Despite the reconstruction, the sur-

    face energy of this termination is high. As a result

    of the reconstruction, there is a layer of non-

    bridging oxygen (NBO) atoms at the top of the

    surface. The NBOs and aluminum atoms bonded

    to them also form a regular pattern on the sur-

    face (see fig. 6). This layer of NBOs might result

    in strong reactivity of the surface. For the termi-

    nating layer C the reconstruction is even deeper

    than for the layer A, because there is initially a

    layer of Al atoms on the top of the termination C.

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    7/12

    S. Blonski, S.H. Garofal ini / M olecular dynamics simulat ions of LX- nd y-al umina surfaces 269

    ;

    layer

    a-A O,

    11

    TO)

    layer c

    Fig. 6. Atom configurations for the terminations of the (1120)

    surface of a-alumina.

    The resulting surface has energy higher than the

    layer B, but lower than the layer A. As can be

    seen in fig. 6, the surface C is partially disor-

    dered. From the results, it is believed that longer

    simulations are needed to form an entirely peri-

    odic surface. Additionally, it should be noted

    that, although there are only octahedrally coordi-

    nated aluminum atoms in bulk a-alumina crys-

    tals, the surface Al atoms are often four- or

    five-coordinated, as indicated in fig. 6.

    4.3. 1700)

    surface

    There is only one possible termination of the

    (1iOO) surface. After splitting the crystal at that

    level, the surface undergoes significant recon-

    struction. In the most external layer, two-coordi-

    nated oxygens move slightly above the surface

    formed by three-coordinated oxygens and four-

    coordinated aluminum atoms. Other changes also

    occur in deeper layers, even 5 A from the surface

    (see fig. 7b). The formed surface, shown in fig. 7a,

    displays two-dimensional periodicity. The energy

    of the (liO0) surface is higher than that of the

    lowest-energy terminations of the (0 00 1) and

    (112 0) surfaces. Therefore, contrary to results of

    Tasker [17] and Mackrodt et al. [18], the order of

    surface energies predicted in the present work is:

    (0001) < (1120) < (iioo),

    which coincides with the statistical observations

    presented by Hartman [141.

    5. Surfaces of y-alumina

    The crystal structure of -y-alumina is much

    more complicated than the structure of cy-

    alumina. In a-alumina, Al atoms have only octa-

    hedral coordination, but in y-alumina, aluminum

    atoms are coordinated octahedrally as well as

    tetrahedrally. Occurrence of vacancies at some

    (a>

    20

    k

    10

    a-AI,O,

    (iioo)

    layer A

    0

    (b) lox

    I

    ,

    I

    4

    E, = 2.35 J me2

    Height [A]

    A

    Fig. 7. (a) Atom configurations on the (IiOO) surface of

    a-alumina; (b) density profile for this surface and for the bulk

    crystal in this direction.

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    8/12

    270

    S. Blonski, S.H. Garof ali ni / M olecular dynamics simul ati ons of a- and y-al umina surfaces

    Table 2

    Calculated surface energies for y-alumina (letters in paren-

    theses show notation of surfaces used by other authors [3])

    Surface

    (001) A (EJ

    B (F)

    (110) A (D)

    B (Cl

    (111) A (A)

    B (B)

    Surface energy (J rn-)

    Unrelaxed Relaxed

    surfaces

    surfaces

    3.24

    1.94

    3.37 0.79

    4.62

    2.54

    4.62

    1.21

    9.45 0.87

    14.08

    0.88

    cation sites creates another complication to the

    structure. Hence, surfaces of y-alumina have a

    more complex structure than a-alumina. In par-

    ticular, surface relaxation in y-alumina can vary

    in different regions depending on the distribution

    of Al vacancies in the atomic layer below the

    surface. Such behavior was observed in the simu-

    lations.

    Therefore, in order to find the average proper-

    ties of the y-alumina crystal surfaces, 12 to 16

    different samples for each of the surfaces were

    studied. Three y-alumina surfaces were exam-

    ined: (00 l>, (1 lo), and (111). The average sur-

    face energies obtained are shown in table 2. The

    surface energies for y-alumina are for the most

    part lower than the energies for a-alumina. The

    surfaces which have the lowest energies became

    amorphous during reconstruction. However, sim-

    ulations of bulk amorphous alumina resulted in

    higher energy than the crystalline y-alumina, in-

    dicating that the crystal is the more stable bulk

    structure and the amorphization of the surface is

    not an artifact.

    Surface disordering during simulations is ac-

    companied by a change in coordination of alu-

    minum atoms: the number of tetrahedral Al atoms

    increases. For each of the surfaces studied, there

    is an apparent dispersion of the values of surface

    energy, but the structural characteristics are simi-

    lar. Neither the number of defects nor the extent

    of disordering was directly related to the surface

    energy. However, the energy is lower for the

    surfaces which contain more vacancies located in

    the layer adjacent to the surface. Results for

    particular surfaces of y-alumina are discussed in

    more detail in the following subsections.

    5.1. 001)

    urface

    There are two possible terminating layers of

    the (0 0 1) surface. Fig. 8 shows density profiles of

    both surfaces compared with the one of the bulk

    crystal. To enhance details, only top regions of

    the crystals are shown. Density profiles of the

    entire crystals show that surface relaxation in

    y-alumina is extended to a depth similar to that

    in the a-alumina crystals, but interesting features

    occur close to the surface level. The crystal struc-

    ture of y-alumina is preserved quite well on the

    (00 1) surface. For the A-layer only some oxygen

    atoms move slightly above the surface. For the

    B-layer, which has a surface energy significantly

    lower than the A-layer, some aluminum atoms

    were located above the surface at the beginning

    of the simulation. However, during surface relax-

    ation these atoms move toward a layer of oxygen

    atoms and hide among them. This must be the

    main source of the drop in the surface energy of

    the B-layer due to relaxation. From the density

    profiles, it can be concluded that on the (00 1)

    surface of y-alumina, oxygen and aluminum atoms

    are located mainly on the same level, with some

    I I

    cl I

    Fig. 8. Density profiles for various terminations of the (00 1)

    surface of y-alumina and for the bulk crystal in this direction.

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    9/12

    S. Blonski, S.H. Garofal ini / M olecular dynamics simul ati ons of a- and y-alumi na surfaces

    oxygen atoms placed a little above the surface of

    the A-layer.

    Fig. 9 shows atom configurations of the simu-

    lated surfaces and compares them with that of

    the idealized ones. In the idealized case, no va-

    cancies are included in the crystal structure,

    whereas they are included in the simulated sys-

    tem. The configurations of the idealized and sim-

    ulated A-layers are very similar, however, vacan-

    cies are present at some positions of the tetrahe-

    drally coordinated aluminum atoms located

    slightly below the surface. Due to these vacancies

    and the lack of the attracting cations, some pairs

    of two-coordinated oxygen atoms rotate above

    the surface. This is a major difference from the

    idealized A surface, which only consists of three-

    coordinated oxygens. Because atoms do not

    change their coordinations during the simula-

    tions, only surface relaxation takes place for the

    A-layer. For the B-layer, there is clearly a surface

    reconstruction instead of relaxation. All the two-

    coordinated aluminum atoms, which at the begin-

    ning of the simulation were located above the

    surface, change their coordination as they move

    toward oxygen atoms and form new bonds with

    them. Mainly four- and five-coordinated Al atoms

    appear on the surface, with some three-fold Al

    Idealized

    0 10

    20

    Fig.

    9.

    Atom configurations for the terminations of the (00 1)

    surface for the idealized and simulated models of y-alumina.

    Al

    -0

    I :

    271

    0

    Fig. 10. Density profiles for various terminations of the (1 10)

    surface of y-alumina and for the bulk crystal in this direction.

    atoms also present due to defects in the structure

    of the surface. Additionally, there are more alu-

    minum atoms on the simulated surface then on

    the idealized one because some Al atoms from

    the lower level move toward the surface during

    the reconstruction. There is also a difference in

    coordinations of the oxygen atoms present on the

    surface. There are fewer tetrahedrally coordi-

    nated oxygens on the simulated surface than on

    the idealized one. The oxygen atoms are mainly

    three-fold, but two-coordinated oxygens also oc-

    cur. The difference may be very important be-

    cause the lower surface energy of the B-layer

    makes it the most probable termination on the

    (00 1) surface of y-alumina.

    5.2.

    110) surface

    There are also two possible terminations for

    the (110) surface. Fig. 10 shows density profiles

    of the sample surfaces compared with that of the

    bulk crystal. Fig. 11 shows atom configurations of

    the respective surfaces. Configurations of the

    simulated surfaces are also compared in this fig-

    ure with that of the idealized ones.

    For the A-termination, surface relaxation en-

    compasses three layers of atoms. Oxygen atoms,

    which are located above the cation vacancies,

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    10/12

    272

    S. Blonski, S.H. Garof ali ni / M olecular dynamics simul ati ons of a- and y-al umina surfaces

    rotate slightly above the surface. They are bonded

    to only one aluminum atom each. Thus, the sur-

    face of the A-layer consists not only of the two-

    and three-coordinated oxygens, but also of NBOs.

    Coordination of the aluminum atoms does not

    change during the relaxation, but the density

    profile shows that some of them slightly follow

    the NBOs in the move upward. In spite of differ-

    ences in surface energy, the configurations of all

    of the (110) A-layers studied here are very simi-

    lar to each other. They differ in the number of

    NBOs on the surface, due to the different va-

    cancy distributions in each simulation, but the

    surface energy does not seem to be correlated to

    that number.

    There is also a difference in the surface energy

    among the B-terminations. Decrease in the sur-

    face energy is accompanied by the disorder ap-

    pearing on the surface. The B-surface shown in

    fig. 11 is the one with the moderate surface

    energy and with the moderate number of defects

    on the surface. The two top atom layers are

    involved in the surface reconstruction with even

    some oxygen atoms from the second layer moving

    to the surface. As a result, five coordinated alu-

    minum atoms appear on the surface, in addition

    to the three- and four-fold ones. The oxygen

    dealized

    Simulated

    I.. . . . . . . /

    I

    AsA A.A A.A A:

    r..

    ,. AA AA AA AA;

    . . . . . . . .

    A+A A.A A&A A;

    AA AA AA AA;

    k

    90 A GQ I

    O.O*A,A A.A

    layer B

    o1...

    . . . .

    .L ,

    &A&,

    -

    10

    2

    0 18 20

    Fig.

    11.

    Atom configurations for the terminations of the (1 10)

    surface for the idealized and simulated models of y-alumina.

    1

    Fig. 12. Density profiles for various terminations of the

    (I

    1 1)

    surface of y-alumina and for the bulk crystal in this direction.

    atoms are no longer three-fold only, but some of

    them are rather two-coordinated. Large defects

    occurring on the surface additionally expose the

    second layer of atoms, therefore, it is difficult to

    estimate the number of atoms on the surface.

    5.3. Cl I 1) surf ce

    There are several possible terminations of this

    surface, however only two of the ones which

    expose aluminum atoms above the level of oxy-

    gens were simulated (see fig. 121. Those termina-

    tions have been usually considered as the A- and

    B-layers of the (1 1 I> surface [3]. Surface energies

    of both terminations are high, but different, be-

    fore reconstruction; after the reconstruction the

    average energies are equal. The energy is usually

    lower for layers with the larger number of the

    cation vacancies present near the surface. For all

    of the (1 11) surfaces studied, the reconstruction

    is extensive and the surfaces become amorphous

    or at least significantly distorted. Fig. 13 shows

    that the regular patterns of the idealized termina-

    tions vanish during the simulations. A driving

    force of the reconstruction is a tendency to in-

    crease coordination of the aluminum atoms. As a

    result, the number of three-coordinated Al atoms

    is much lower on the simulated surfaces than on

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    11/12

    S. Blonski, S.H. Garofal ini / M olecular dynamics simul ati ons of a- and y-alumi na surfaces

    273

    Idealized

    Fig. 13. Atom configurations for the terminations of the (111)

    surface for the idealized and simulated models of y-alumina.

    the idealized ones. One-coordinated Al atoms,

    initially present in the A-layer, also bind to addi-

    tional oxygens during reconstruction. Because of

    this, disorder is much broader on the A-layers

    than on the B-layers. It is clearly shown by figs.

    12 and 13 which present examples of both A and

    B terminations of the (111) surface with the

    similar surface energies which are also close to

    the average values.

    6.

    Conclusions

    The simulations show that surfaces of the alu-

    mina crystals can be terminated by different lay-

    ers of atoms. Although some terminations seem

    to be unfavorable when the unrelaxed crystal

    surfaces are considered, they become more prob-

    able after surface relaxation or reconstruction

    during the simulations. For basal surface of (Y-

    alumina, the termination C which was usually

    considered in earlier theoretical studies has the

    lowest surface energy in the present simulations.

    In addition, the simulations show that another

    termination (B) has the energy only slightly higher

    then C and looks exactly the same as C. This is in

    agreement with experimental observations of two

    terminating layers on the (0 0 0 1) surfaces. Simu-

    lation results cannot be compared with experi-

    mental observations of the (1 i 00) and (1 120)

    surfaces, because such data are not yet available.

    However, regularity of the surfaces suggests that

    they might be an object of studies on the atomic

    level, e.g. by atomic force microscopy.

    Surfaces of y-alumina obtained from the simu-

    lations differ greatly from idealized surfaces which

    have been considered so far in previous studies.

    One reason for the difference is the presence of

    cation vacancies in the current simulations. Al-

    though the presence of vacancies in the crystal

    structure of -y-alumina has been recognized for a

    long time, the vacancies were usually ignored in

    previous models of the surfaces. Differences are

    also caused by the occurrence of surface relax-

    ation or reconstruction. Atoms with various coor-

    dinations, not expected from the idealized mod-

    els, appear on the surfaces as a result. Some

    surfaces are often disordered, but their energy is

    significantly lower than the lowest one for (Y-

    alumina. This may explain the experimental dif-

    ference in surface area between these two forms

    of alumina.

    Acknowledgement

    The authors acknowledge support from the

    Center for Ceramic Research at Rutgers Univer-

    sity.

    References

    [l] N. Yao, Z.L. Wang and J.M. Cowley, Surf. Sci. 208

    (1989) 533.

    [2] Z.L. Wang, Surf. Sci. 271 (1992) 477.

    [3] H. Knazinger and P. Ratnasamy, Catal. Rev.-Sci. Eng. 17

    (1978) 31.

    [4] S.H. Garofalini, J. Non-Cryst. Solids 120 (1990) 1.

    [5] D.M. Zirl and S.H. Garofalini, J. Am. Ceram. Sot. 73

    (1990) 2848.

    [6] D.M. Zirl and S.H. Garofalini, J. Am. Ceram. Sot. 75

    (1992) 2353.

    [7] M.P. Allen and D.J. Tildesley, Computer Simulation of

    Liquids (Clarendon, Oxford, 1990).

    [8] R.W.G. Wyckoff, Crystal Structures, 2nd ed. (Intersci-

    ence, New York, 1963).

  • 7/23/2019 Surface Science Volume 295 issue 1-2 1993 [doi 10.1016%2F0039-6028%2893%2990202-u] S. Blonski; S.H. Garof

    12/12

    274

    S. Blonski, S.H. Garofal ini / M olecular dy namics simulat ions of Y- nd y-alum ina surfaces

    [9] R.E. Newnham and Y.M. de Haan, Z. Kristallogr. 117

    (1962) 235.

    [lo] V. Jayaram and C.G. Levi, Acta Metall. 37 (1989) 569.

    [1 l] W.H. Gitzen, Alumina as a Ceramic Material (American

    Ceramic Society, Columbus, OH, 1970).

    [12] L.J. Alvarez, J.F. Sanz, M.J. Capitan and J.A. Odriozola,

    Chem. Phys. Lett. 192 (1992) 463.

    [13] E.J.W. Verwey, Z. Kristallogr. 91 (1935) 65.

    [14] P. Hartman, J. Cryst. Growth 96 (1989) 667.

    [15] T. Hsu and Y. Kim, Surf. Sci. 258 (1991) 119.

    [16] Y. Kim and T. Hsu, Surf. Sci. 258 (1991) 131.

    [17] P.W. Tasker, in: Surfaces of Magnesia and Alumina, Ed.

    W.D. Kingery (American Ceramic Society, Columbus,

    OH, 1984).

    [18] W.C. Mackrodt, R.J. Davey and S.N. Black, J. Cryst.

    Growth 80 (1987) 441.

    [19] W.D. Kingery, H.K. Bowen and D.R. Uhlmann, Intro-

    duction to Ceramics, 2nd ed. (Wiley, New York, 1976).

    [20] S.H. Overbury, P.A. Bertrand and G.A. Somorjai,

    Chem.

    Rev. 75 (1975) 547.

    [21] M. Causa, R. Dovesi, C. Pisani and C. Roetti, Surf. Sci.

    215 (1989) 259.