Sie sind auf Seite 1von 7

Top Catal (2011) 54:1319

DOI 10.1007/s11244-011-9634-9

ORIGINAL PAPER

Assembly of Chiral Amino-Acids at Surfaces from a Single


Molecule Perspective: Proline on Cu(110)
M. Forster M. S. Dyer M. Persson
R. Raval

Published online: 21 January 2011


Springer Science+Business Media, LLC 2011

Abstract We present a molecule-by-molecule mapping


of an amino-acid monolayer on a Cu(110) surface in which
individual molecular configuration, bonding and adsorption
footprints are identified. This detailed mapping provides
direct evidence that the (4 9 2) structures of enantiopure
and racemic proline are governed by a strict heterochiral
adsorption footprint template. For the racemic system this
has an interesting consequence, namely that it leads to a
2-D random solid solution, a situation that is rarely
encountered in 3-D.
Keywords Scanning tunnelling microscopy  Random
solid solutions  Chiral surfaces  Amino acids  Footprints

1 Introduction
Introducing enantioselectivity to heterogeneous catalysis is
a significant scientific and technological challenge and
currently little understanding exists of key reaction mechanisms that underpin such processes at solid surfaces.
Enantioselectivity is critical to the pharmaceutical industry
where different enantiomers of a molecule may induce very
different physiological responses, an example being the
thalidomide tragedy. In many cases it is imperative that
products are enantiomerically pure thus fuelling the need

M. Forster  M. S. Dyer  M. Persson  R. Raval (&)


Surface Science Research Centre and Department of Chemistry,
University of Liverpool, Liverpool L69 3BX, UK
e-mail: R.Raval@liv.ac.uk
M. Persson
Department of Applied Physics, Chalmers University
of Technology, SE-412 96 Goteborg, Sweden

for enantioselective catalysis [1]. This stringent requirement is also present in other technological sectors such as
fragrances, flavours and agrochemicals. There are a number of different approaches which have been used to create
heterogeneous chiral catalysts including: immobilising
chiral transition metal complexes at surfaces [2]; attaching
chiral auxiliaries to the reactant to induce an asymmetry in
the surface reaction [3]; creating intrinsically chiral surfaces in which asymmetric configurations of the reactive
surface are displayed [4]; and, finally, adsorbing chiral
molecules at achiral solid surfaces to induce asymmetry
[5]. The work discussed here directly impinges on the last
case, where there are several successful examples of
introducing enantioselectivity to heterogeneous reactions
by modifying surfaces via the adsorption of chiral organic
molecules. Examples of surface modifiers include tartaric
acid and the a-amino-acid, alanine in the hydrogenation of
b-ketoesters over supported Ni catalysts [610]. The
hydrogenation of a-ketoesters to a-hydroxyesters may also
be made enantiospecific by the modification of Pt and Pd
surfaces with chiral cinchona modifiers [1113]. In a different application, the amino-acid proline and its derivatives have proved to be highly effective as chiral auxiliaries
in diastereoselective heterogeneous catalysis [14, 15]. We
note that (S)-proline was also implicated to act as a chiral
surface modifier in the Pd/C catalysed asymmetric hydrogenation of isophorone [11, 16]. Recent work by Lambert
and co-workers [17] has questioned this interpretation and
suggested that enantiodifferentiation in the reaction occurs
in solution and not at the surface, thus highlighting the
extreme difficulty of elucidating mechanisms in enantioselective heterogeneous catalysis.
A vital platform for progress in this field is establishing
a fundamental understanding of how chiral modifiers
interact with surfaces. The complexity of real catalytic

123

14

conditions prohibits detailed insights of the enantioselective site from being extracted and, therefore, model systems have been utilised extensively. In such studies the
behaviour of organic modifiers on well defined single
crystal metal surfaces is investigated under ultra high
vacuum (UHV) conditions, allowing a broad range of
surface science techniques to be utilised. There are a
number of reviews that provide a comprehensive overview
of research in this area [1820]. In addition to this, a parallel body of work has endeavored to produce a nanoscale
understanding, down to the single molecule level, of how
chirality is nucleated, expressed and propagated [2127].
Of particular interest is the behaviour of amino-acids at
surfaces [2840] as they represent cheap, abundant materials and, in addition to implanting enantioselectivity in
heterogeneous reactions [7, 8], they have potential use in
nano-technological applications based upon bio-inorganic
interfaces, sensors and molecular electronics.

2 Model Systems: The Adsorption of Amino-Acids


on a Cu(110) Surface
2.1 The Challenge of Structure Determination
within Highly Organised Amino Acid Monolayers
The most challenging aspect of providing an in-depth
description of amino-acid organisations is attaining
knowledge on how each molecule bonds to the surface. A
widely used convention for describing the interaction of a
molecule with the surface is the adsorption footprint
which describes the 2-D space enclosed by the bonding
points of the molecule upon adsorption [22]. Monte Carlo
simulations by Zgrablich and co-workers [41, 42] demonstrate that the evolution of molecular surface organisation
and local patterns is strongly influenced by the adsorption
footprints the system adopts. It has previously been
established experimentally that molecules in the (3 9 2)
structures of alanine and glycine on Cu(110) [3140] bond
to the surface by three distinct points, via the two oxygen
atoms of the carboxylate group and the nitrogen atom of
the amino group. The carboxylate oxygens were found to
be located at on-top positions along the close-packed Cu
rows, with the nitrogen positioned at an on-top position on
the adjacent Cu row. The adsorption footprint is, thus,
described by a right-angled triangle which may be either
left or right handed depending upon the position of the
amino group relative to the carboxylate group, Fig. 1. The
left and right handed footprints are mirror images in twodimensions and are, therefore, chiral [22]. A distinction
between molecular and footprint chirality is provided in
Fig. 1. The key question in the field is whether an entire
amino-acid assembly utilises the same chiral footprint

123

Top Catal (2011) 54:1319

Fig. 1 Schematic representation of the bonding geometries and


associated triangular adsorption footprints of amino acids on Cu(110).
Colour code: Red sphere = oxygen, Blue sphere = nitrogen (hydrogen atoms not included). Chiral systems can be divided into homochiral systems where only one enantiomer of an object exists or as
heterochiral where both enantiomers co-exist. For adsorbed surface
species, it is often useful to de-convolute chirality in terms of the
inherent chirality of the adsorbed molecule and the chirality of
the adsorption footprint that describes the bonding contacts of the
molecule with the surface [23, 25]. This figure shows two distinct
bonding geometries and the corresponding footprints which may
result from the interaction of a single enantiomer of an amino acid
with a surface. The adsorption footprints of the two species are mirror
images in 2-D. Therefore, this surface system possesses molecular
homochirality but may display footprint heterochirality

(footprint homochirality) or whether both mirror forms of


the footprints are utilised (footprint heterochirality). This
question is particularly difficult to address experimentally
since techniques such as reflection absorption infrared
spectroscopy (RAIRS) and near edge X-ray absorption fine
structure (NEXAFS) are averaged across the whole surface
and do not provide information at the single-molecule
level. Although, nanoscale probing by scanning tunneling
microscopy (STM) should be able to reveal this information, in practice there is generally insufficient image contrast to allow discrimination between the various
conformers and their adsorption footprints [34].
2.2 Enantiopure (S)-Proline on Cu(110): Probing
Adsorption Footprints at the Single Molecule Level
In order to provide an unambiguous characterisation of
amino-acid adlayers, we turned to the amino-acid proline
which uniquely has the amino group incorporated within a
pyrrolidine ring (Fig. 2), leading to greatly increased
structural rigidity. As a direct consequence of this rigidity
it was anticipated that any changes to the bonding

Top Catal (2011) 54:1319

Fig. 2 Chemical structure of the molecular and anionic forms of the


amino acid proline

configuration of the amino group relative to the fixed carboxylate group (i.e. different adsorption footprints) would
trigger a significant reorientation of the attached pyrrolidine ring. This would, in turn, lead to image contrasts in
STM and thus act as a marker of each adsorption footprint
and molecular geometry. This level of information should
then enable a molecule-by-molecule analysis of the
organised assembly.
Adsorption of enantiopure (S)-proline on Cu(110) at
room temperature leads to a highly organised (4 9 2)
overlayer (Fig. 3a) containing the anionic prolate species
(Fig. 2) [2830]. The larger unit cell of the proline organisation compared to the alanine and glycine (3 9 2)
structures [3140] is attributed to the larger molecular
dimension of proline due to the pyrrolidine ring. STM
images of the organisation show the (4 9 2) structure
contains two distinct prolate conformers within the two
molecule unit cell, which image as bright and faint protrusions (Fig. 3a, inset) [28]. DFT calculations demonstrate
that the two observed conformers arise as a direct consequence of the molecule possessing different adsorption
footprints. We find that the bonding motif outlined for the
alaninate and glycinate systems is also adopted here and
the two carboxylate oxygen atoms bind to adjacent copper
atoms in the 110 row. In addition, a third bonding contact

Fig. 3 a STM image of the (S)-proline (4 9 2) structure on Cu(110)


with a coverage of 0.25 monolayers (the (4 9 2) unit cell is
2, I(t) = -0.4 nA, V(t) = -806 mV). Inset
highlighted) (94 9 102 A
Enlarged STM image of a (4 9 2) unit showing conformers A (bright)
and B (faint), arrows point to relaxed adsorption geometries for

15

occurs via the nitrogen to a copper atom across from one of


the oxygen binding sites in the neighbouring row. If the
nitrogen atom bonds to the copper atom to the left, a lefthanded triangular adsorption footprint is described and the
pyrrolidine ring tilts significantly away from the surface
(conformer A, Fig. 3b), imaging as a bright protrusion in
STM. However, if the nitrogen bonds to the right, a righthanded triangular adsorption footprint is described, the ring
lies flat (conformer B, Fig. 3b) and images as a faint protrusion. The orientation of the ring, therefore, does indeed
act as a marker of the molecular conformation and
adsorption footprint [28].
It is now important to consider how the prolate conformers organise within the (4 9 2) structure. There are in
fact eight possible (4 9 2) arrangements that can be constructed using combinations of the two conformers identified [28]. Adsorption energies were calculated for each
possible arrangement and an STM image was simulated for
direct comparison with experiment. It was found that the
most stable structure utilises both conformers A and B in a
heterochiral footprint arrangement and provides a simulated STM image in excellent agreement with experiment
(Fig. 4a and b). Mapping the footprints from experimental
STM images reveals that both left and right handed
adsorption footprints are incorporated in the overlayer with
alternate rows of each footprint propagating along the
[001] surface direction (Fig. 4c). Therefore, in terms of
footprint chirality the assembly possesses a strict heterochiral organisation. Periodic density functional theory
(DFT) calculations are also able to provide insights into the
driving forces behind such an assembly. Specifically, it is
shown that the formation of one-dimensional intermolecular hydrogen bonded chains along the 110 direction,
arising through the interaction of the hydrogen of the

conformers A and B shown in (b) from left to right: view from above,
the side, and in terms of their triangular adsorption footprint (Colour
code for atoms: red = oxygen, blue = nitrogen, black = carbon,
white = hydrogen)

123

16

Fig. 4 a Simulated STM image for the most energetically favourable


(S)-proline (4 9 2) arrangement, the positions of the atoms in the
prolate anions are included. The calculated image is shown in a
(12 9 6) unit cell at a bias voltage of -1.25 V with an average tip . b Experimental STM image of the (4 9 2)
surface height of 7.5 A
2, It = -0.43 nA, Vt = -1250 mV)
structure (31.1 9 21.7 A
c Schematic representation of the heterochiral adsorption footprint
arrangement (Filled triangle = conformer A, Open triangle = conformer B)

amino group with the oxygen atom of a neighbouring


carboxylate group, stabilises the overlayer [28].
2.3 Racemic (RS)-Proline on Cu(110): Formation
of a 2-D Random Solid Solution
The complexities of providing a detailed characterisation
of a racemic amino-acid overlayer are even greater than for
the enantiopure phase and little is known of the enantiomer
arrangements within such layers. Thus, the insights gained
into the enantiopure (S)-proline overlayer provide a unique
opportunity to unravel the behaviour of the racemic proline
assembly [29]. Fig. 5a displays an STM image of the
racemic (RS)-proline (4 9 2) overlayer on Cu(110) and
reveals that the racemic overlayer also contains both bright
and faint protrusions, showing that conformers A and B
identified in the enantiopure system are also favoured

123

Top Catal (2011) 54:1319

here [29]. However, in contrast to the regular (4 9 2)


pattern of the enantiopure layer, the distribution of conformers in the racemic overlayer is random (Fig. 5a). Upon
close inspection of high resolution STM images (Fig. 5c
and d) more detailed information may be extracted. First,
the protrusions of the flat-lying conformer B image with a
distinct triangular shape which lie in one of two mirror
directions with respect to the 110 axis, indicated with
triangles in Fig. 5c. These opposing alignments are similar
to those exhibited by the (R)- and (S)-enantiopure systems
(Fig. 5b) and we, therefore, attribute the mirroring protrusions of conformer B to molecules of opposing chirality.
Second, at high image contrast, an oval shaped protrusion
can be observed for the upright conformer A, also possessing mirroring directions with respect to the 110 axis,
indicated with ovals in Fig. 5d. Again, by comparison to
the enantiopure systems (Fig. 5b) these are associated with
conformers of opposite chirality. Thus, the STM images
associated with each conformer of a specific chirality can
be identified and are shown in Fig. 5b. Furthermore, if the
associated footprints of the conformers are examined, it can
be seen that conformer A of the (R)- enantiomer possesses
the same adsorption footprint as conformer B of the
(S)- enantiomer and vice versa (Fig. 5b).
By combining the observations above, we are now in a
position to identify both the molecular and footprint chirality at the single-molecule level within the ordered
racemic amino acid assembly. If one considers the molecules within a row along the [001] axis, it can be seen that a
random mixture of conformer A (bright) and conformer B
(faint) are present (Fig. 5c). However, along a row both
conformers image as protrusions that are oriented in the
same direction, regardless of the conformation. Examining
the adjacent rows directly above and below, it can be seen
that the protrusions are oriented in the opposite direction
(Fig. 5c and d). Given this information, a detailed picture
of the adlayer can be constructed in which the molecular
chirality and the adsorption footprint is identified molecule-by-molecule. We have done this explicitly for a region
of the (4 9 2) overlayer in Fig. 6a and obtained a map of
individual chirality, conformation and footprints within the
adlayer, Fig. 6b. Although, this arrangement looks rather
complex, it can be deconvoluted into a separate map of
molecular chirality and a map of the adsorption footprints
adopted, Fig. 6c. From this, a very interesting picture
emerges. First, it can be seen that the distribution of
molecular chirality within the adlayer is random, i.e. a
pseudo-racemate is created and, therefore, the assembly
is a two-dimensional random solid solution. However,
when one scrutinises the adsorption footprints, a highly
significant observation emerges, namely that the adsorption
footprints within the (4 9 2) organisation adhere strictly to

Top Catal (2011) 54:1319

17

Fig. 5 a STM image of the


(RS)-proline overlayer on
2,
Cu(110) (158 9 125 A
I(t) = -0.62 nA, V(t) =
-1238 mV). b STM images
showing the individual
conformers (A and B) for both
the (S) and (R) enantiomers, the
associated adsorption footprint
is indicated
(Red = (S) enantiomer,
Blue = (R) enantiomer, Filled
triangle = conformer A, Open
triangle = conformer B). c and
d High resolution STM images
showing mixed conformer
mixed chirality overlayer
2, I(t) = -0.49 nA,
(44 9 50 A
V(t) = -856 mV) at differing
image contrast

a specific heterochiral pattern (Fig. 6c), i.e. the same pattern as observed for the enantiopure (S)-proline system
[29]. Essentially, the evidence clearly points to the fact that
the chirality of the adsorption footprint, and not that of the
molecule, dictates the arrangements of both enantiopure
and racemic proline at the surface. Consequently, for the
racemic system, molecular chirality within a particular row
is not conserved; rather, the randomly distributed enantiomers adapt their conformers so that the adsorption footprint chirality is maintained along the row.
DFT calculations show that the heterochiral footprint
arrangement is preferred as it minimises repulsive interactions and maximises intermolecular hydrogen bonding
[29]. Specifically, the relative position of the bonding
carboxylate groups is a crucial driving force behind the
heterochiral footprint arrangement. If carboxylate groups
bind to every atom in the same copper row a highly
repulsive compressive strain is incurred thus an unfavourable adsorption energy is computed. The heterochiral
footprint arrangement staggers carboxylate groups between
copper rows and, therefore, this energy penalty is avoided.
Consideration of hydrogen bonding interactions show that
within the specific heterochiral footprint arrangement,
conformer A of (R)-proline is equivalent to the B conformer of (S)-proline, with respect to the formation of
hydrogen-bonded chains. Thus, each adsorption position in

the footprint template can be occupied by either enantiomer


resulting in a 2-D random solid solution at the surface [29].
2.4 Comparison with Previous Amino Acid Studies
on a Cu(110) Surface
Previous calculations of the (3 9 2) organisations of glycine and alanine on Cu(110) by Sholl [31, 32] and Jenkins
[33] predicted that a heterochiral adsorption footprint
arrangement is adopted, however no experimental proof of
such an amino-acid assembly existed. Using the (4 9 2)
assembly of proline on Cu(110) we are able to provide the
first direct experimental verification of the predicted heterochiral adsorption footprint arrangement. In the enantiopure (S)-proline system the individual adsorption
footprints are distinguishable in STM due to the image
contrast of the pyrrolidine ring for different adsorption
footprints [28]. More striking, however, is the confirmation
of Sholls calculation of racemic alanine on Cu(110) which
predicted that the system would behave as a random solid
solution [32]. The STM images collected for the racemic
proline (4 9 2) organisation allow both the molecular and
footprint chirality to be accessed at the single molecule
level, from which it is shown that the distribution of chirality is indeed random, however the assembly is dictated
by the heterochiral adsorption footprint arrangement [29].

123

18

Top Catal (2011) 54:1319

There remains a significant gap between the conditions


of these experiments and those of a working catalytic
system. However, as pointed out previously by Raval [24],
a number of concepts developed here are amenable to
translation to the working heterogeneous enantioselective
system. For example, although the systems under investigation in this work are idealised, catalytic studies show that
the size of the metal particles is a critical factor in the
catalyst performance, with higher enantiomeric excesses
observed for larger particles, reaching a maximum at
before reaching a plateau [7, 8]. Particles of this
100300 A
size would be capable of sustaining two-dimensional
molecular assemblies of the type described here. An exact
knowledge of the bonding motifs of chiral modifiers on
surfaces is critical in designing enantioselective catalysts
since this will govern the orientation of the functional
groups which form the enantioselective pocket. The
nanoscale insights gained from molecular-level probing by
surface science techniques and high-level DFT are beginning to reveal the delicate balances which exist between
the molecule-surface interactions and intermolecular forces
that determine chiral surface structures. These factors must
be present in the real catalytic system, albeit in a modified
form when solvents and promoters, etc. are present. The
acquisition of a fundamental knowledge, down to the single-molecule, of the surface behaviour of chiral modifiers
will thus enable rational designs of a new generation of
heterogeneous catalysts.
2). b the
Fig. 6 a Enlarged STM image from Fig. 5c (37 9 43 A
chirality and the adsorption footprint of each molecule is indicated
(key for footprints and chirality as in Fig. 5. c de-convolution of the
image into its molecular chirality and adsorption footprint arrangement

3 Conclusions
Gaining a molecular level understanding of the behaviour
of chiral modifiers at surfaces is pivotal to advances in
enantioselective heterogeneous catalysis. Amino-acids
represent an important class of surface modifiers and much
effort has been made to provide accurate descriptions of
their organisation, conformation and bonding at surfaces.
In this article we have shown how an amino acid overlayer
may be characterised in unprecedented detail. A combined
STM and DFT study demonstrated that the (4 9 2) surface
organisation of (S)-proline on Cu(110) adopts a strict heterochiral adsorption footprint arrangement stabilised by
intermolecular hydrogen bonded chains [28]. The corresponding racemic (RS)-proline organisation is also governed by the heterochiral adsorption footprint arrangement
and results in a random distribution of molecular chirality
at the surface thus creating a two-dimensional random solid
solution [29].

123

Acknowledgments R.R and M.F acknowledge the Engineering and


Physical Sciences Research Council (EPSRC), EU seventh framework small scale collaborative project RESOLVE (NMP4-SL-2008214340) and the University of Liverpool for equipment grants,
funding and a studentship for M.F. M.P acknowledges the Swedish
Research Council (VR) and the Marie Curie Research Training
Network PRAIRIES, contract MRTN-CT-2006-035810. M.S.D
acknowledges the University of Liverpool for funding of a postdoctoral fellowship and computational resources.

References
1. Sheldon RA (1993) Industrial synthesis of optically active compounds. Chirotechnology Dekker, New York
2. Dubois V, Jannes G (eds) (1995) Chiral reactions in heterogeneous catalysis. Plenum, New York, p. 33
3. Baiker A, Blaser HU (1997) In: Ertl GH, Knoezinger H,
Weinheim J (eds) Handbook of heterogeneous catalysis, vol 5.
VHC, New York, p 2442
4. McFadden CF, Cremer PS, Gellman AJ (1996) Langmuir
12:2483
5. Besson M, Debleqc F, Gallezot P, Neto S, Pinel C (2000) Chem
Eur J 6:949
6. Izumi Y (1983) Adv Catal 32:215
7. Tai A, Harada T (1986) Asymmetrically modified nickel catalysts. In: Iwasawa Y (ed) Tailored metal catalysts, Reidel, Tokyo,
p 265
8. Orito Y, Imai S, Niwa S (1980) J Chem Soc Jpn 670

Top Catal (2011) 54:1319


9. Webb G, Wells PB (1992) Catal Today 12:319
10. Blaser HU, Jalett HP, Muller M, Studer M (1997) Catal Today
37:441
11. Tungler A, Kajtar M, Mathe T, Toth G, Fogassy E, Petro J (1989)
Catal Today 5:159
12. Baiker A (1997) J Mol Catal A 115:473
13. Wells PB, Wilkinson AG (1998) Top Catal 5:39
14. Besson M, Blanc B, Champelet M, Gallezot P, Nasar K, Pinel C
(1997) J Catal 170:254
15. Ranade VS, Consilglio G, Prins R (1999) Catal Lett 58:71
16. Tungler A, Mathe T, Petro J, Tarnai T (1990) J Mol Catal A
61:259
17. McIntosh AI, Watson DJ, Burton JW, Lambert RM (2006) J Am
Chem Soc 128:7329
18. Raval R (2001) Cattech 5:1
19. Ma Z, Zaera F (2007) In: Ozkan US (ed) Design of heterogeneous
catalysis: new approaches based on synthesis, characterization,
and modelling, Wiley-VCH, Weinheim
20. Tysoe WT (2007) Nanotoday 2:53
21. Raval R (2009) Chem Soc Rev 38:707
22. Perez-Garcia L, Amabilino DB (2007) Chem Soc Rev 36:941
23. Barlow SM, Raval R (2003) Surf Sci Rep 50:201
24. Ernst K-H (2006) Supramolecular surface chirality. Top Curr
Chem 265:209
25. Barlow SM, Raval R (2008) Curr Opin Colloid Inter Sci 13:65
26. Humblot V, Barlow SM, Raval R (2004) Prog Surf Sci 76:1
27. Forster M, Raval R (2009) Chiral expression by organic architectures at metal surfaces: the role of both adsorbate and surface
in inducing asymmetry. In Rioux R (ed) Model systems in

19

28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.

catalysis: single crystals to supported enzyme mimics, Springer,


New York, p 97
Forster M, Dyer MS, Persson M, Raval R (2009) J Am Chem Soc
131:10173
Forster M, Dyer MS, Persson M, Raval R (2010) Angew Chem
Int Ed 49:2344
Mateo Marti E, Barlow SM, Haq S, Raval R (2002) Surf Sci
501:191
Rankin RB, Sholl DS (2005) J Phys Chem B 109:16764
Rankin RB, Sholl DS (2004) Surf Sci 548:301
Jones G, Jones LB, Thibault-Starzyk F, Seddon EA, Raval R,
Jenkins SJ, Held G (2006) Surf Sci 600:1924
Barlow SM, Louafi S, Le Roux D, Williams J, Muryn C, Haq S,
Raval R (2005) Surf Sci 590:243
Barlow SM, Louafi S, LeRoux D, Williams J, Muryn C, Haq S,
Raval R (2004) Langmuir 20:7171
Haq S, Massey A, Moslemzadeh N, Robin A, Barlow SM, Raval
R (2007) Langmuir 23:10694
Williams J, Haq S, Raval R (1996) Surf Sci 368:303
Barlow SM, Kitching KJ, Haq S, Richardson NV (1998) Surf Sci
401:322
Booth NA, Woodruff DP, Schaff O, Giessel T, Lindsay R,
Baumgartel P, Bradshaw AM (1998) Surf Sci 397:258
Kang J-H, Toomes RL, Polcik M, Kittel M, Hoeft J-T, Efstathiou
V, Woodruff DP, Bradshaw AM (2003) J Chem Phys 118:6059
Unac RC, GilRabaza AV, Vidales AM, Zgrablich G (2007) App
Surf Sci 254:125
Lopez RH, Roma F, Gargiulo V, Sales JL, Zgrablich G (2008) J
Phys Chem B 112:8619

123

Das könnte Ihnen auch gefallen