03.09.2014 Views

Dynamic screening and electron–electron scattering - Donostia ...

Dynamic screening and electron–electron scattering - Donostia ...

Dynamic screening and electron–electron scattering - Donostia ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Surface Science 601 (2007) 4546–4552<br />

www.elsevier.com/locate/susc<br />

<strong>Dynamic</strong> <strong>screening</strong> <strong>and</strong> electron–electron <strong>scattering</strong><br />

in low-dimensional metallic systems<br />

V.M. Silkin a , M. Quijada a,b ,R.Díez Muiño a,c , E.V. Chulkov a,b,c , P.M. Echenique a,b,c, *<br />

a <strong>Donostia</strong> International Physics Center (DIPC), P. de Manuel Lardizabal 4, 20018 San Sebastián, Spain<br />

b Departamento de Física de Materiales, Facultad de Ciencias Químicas UPV/EHU, Apartado 1072, 20080 San Sebastián, Spain<br />

c Unidad de Física de Materiales, Centro Mixto CSIC-UPV/EHU, P. de Manuel Lardizabal 3, 20018 San Sebastián, Spain<br />

Available online 1 May 2007<br />

Abstract<br />

The modification of dynamic <strong>screening</strong> in the electron–electron interaction in systems with reduced dimensionality <strong>and</strong> tunable oneparticle<br />

electronic structure is studied. Two examples of such systems are considered, namely, the adsorbate-induced quantum well states<br />

at the Na adlayer covered Cu(111) surface, <strong>and</strong> metal clusters of sizes up to few nanometers. The dependence of the electron–electron<br />

decay rates on the Na coverage in the former case <strong>and</strong> on the cluster size in the latter is investigated. The role played by the dynamical<br />

screened interaction in such processes is addressed as well.<br />

Ó 2007 Elsevier B.V. All rights reserved.<br />

Keywords: Many-body <strong>and</strong> quasiparticle theories; Surface electronic phenomena; Noble metals; Adsorption; Quantum well states; Clusters<br />

1. Introduction<br />

* Corresponding author. Address: Departamento de Física de Materiales,<br />

Facultad de Ciencias Químicas UPV/EHU, Apartado 1072, 20080<br />

San Sebastián, Spain. Tel.: +34 943 015963; fax: +34 943 015600.<br />

E-mail address: wapetlap@sc.ehu.es (P.M. Echenique).<br />

Electronic excitations in metallic media can decay<br />

through various mechanisms, the most important of them<br />

being electron–electron (e–e) <strong>and</strong> electron–phonon (e–ph)<br />

<strong>scattering</strong>. Over the last years, a large amount of theoretical<br />

<strong>and</strong> experimental work has been devoted to underst<strong>and</strong><br />

<strong>and</strong> predict the lifetime of these excitations in solids [1–5]<br />

<strong>and</strong> at surfaces [6–10]. Focusing on the e–e <strong>scattering</strong> process,<br />

one of the key quantities that determine the rate at<br />

which the electronic excitations can decay is the dynamic<br />

<strong>screening</strong> among the medium electrons. However, the<br />

<strong>screening</strong> properties of metallic systems are profoundly affected<br />

by their dimensionality. Confinement <strong>and</strong> surface effects<br />

are, among others, two features that can drastically<br />

change the electronic <strong>screening</strong> <strong>and</strong> thus the lifetime of<br />

electronic excitations as compared with the bulk case.<br />

The modification at wish of the electronic properties of metal<br />

systems <strong>and</strong> the subsequent change in the lifetime of<br />

electronic excitations is not a question of purely academic<br />

interest, but has important implications for many technological<br />

applications.<br />

In present work, we analyze two cases in which the<br />

reduction of dimensionality leads to significant modifications<br />

in the electron–electron interaction. First, we study<br />

the case of quantum well states formed under the deposition<br />

of alkali atoms on metal surfaces [11,12]. Whereas<br />

the number of adsorbed alkali atoms is small, <strong>and</strong> they<br />

can be considered as independent, the picture of almost total<br />

transition of s valence electron from the alkali atom to<br />

the substrate is appropriate. However, when the coverage<br />

reaches rates of around 1 monolayer (ML) [13] an adsorbate-induced<br />

electron b<strong>and</strong> which forms a quasi twodimensional<br />

electron gas confined between the vacuum<br />

barrier <strong>and</strong> the metal substrate [14,15] (called a quantum<br />

well state (QWS)), unoccupied for low coverage, starts to<br />

fill [16]. A prominent example of this is the case of Na<br />

adsorption on Cu(111), for that the properties of this<br />

QWS have been intensively studied. Thus, the dynamics<br />

of holes created at wave vector k k = 0 has been studied<br />

for this system, both experimentally <strong>and</strong> theoretically<br />

0039-6028/$ - see front matter Ó 2007 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.susc.2007.04.232


V.M. Silkin et al. / Surface Science 601 (2007) 4546–4552 4547<br />

[17]. In this study it was proven that e–e <strong>and</strong> the e–ph<br />

mechanisms are equally important for the hole decay for<br />

1 ML coverage. Very recently a joint experimental <strong>and</strong> theoretical<br />

investigation of electron dynamics for Cs <strong>and</strong> Na<br />

adlayers on Cu(111) was performed as well [18]. Here,<br />

we investigate the energy dependence of the e–e contribution<br />

for several positions of the QWS b<strong>and</strong> relative to the<br />

Fermi level E F , which, in turn, depends on the Na coverage.<br />

For the sake of comparison with the experiment, a reasonable<br />

assumption could be that the e–ph contribution<br />

should not change significantly for such small variation<br />

of the QWS energy.<br />

As a second case study, we analyze dynamic <strong>screening</strong><br />

properties in metal clusters. In clusters <strong>and</strong> nanoparticles,<br />

chemical, optical, <strong>and</strong> electronic properties can be tuned<br />

by varying the size <strong>and</strong>/or the shape of the system. Besides<br />

its fundamental interest, tailoring the electronic properties<br />

can be of vast importance for many processes of technological<br />

interest, such as photochemical reactivity. Over the last<br />

years, the advent of femtosecond lasers has fueled the<br />

experimental study of electron relaxation processes. Particular<br />

attention has been paid to the dependence on size of<br />

the electron–electron interaction processes [19–21]. We will<br />

show that the interplay between <strong>screening</strong> effects <strong>and</strong> space<br />

localization of the initial excitation makes the lifetime of<br />

the excitation vary with respect to the bulk equivalent.<br />

2. Theoretical methods<br />

In very recent years, improvements in experimental<br />

methods, in particular in the field of ultrafast laser spectroscopies,<br />

are making it possible to study electronic excitations<br />

at time scales below the femtosecond [22,23]. In this<br />

time scale, the <strong>screening</strong> of excited electrons in condensed<br />

matter can be incomplete <strong>and</strong> the description of the electronic<br />

excitations as quasiparticles is questionable [24].<br />

However, most st<strong>and</strong>ard experimental techniques provide<br />

information on time scales for which the <strong>screening</strong> of excited<br />

electrons can be approximated as instantaneous in<br />

practice. In this case, the description of the many-body<br />

electronic excitation can be simplified by means of the quasiparticle<br />

picture. In the following, we focus into this<br />

situation.<br />

The <strong>scattering</strong> rate of an excited electron or hole due to<br />

inelastic e–e <strong>scattering</strong> can be evaluated by means of<br />

many-body calculations based on the electronic self-energy<br />

[25,26]. In this approach the lifetime broadening of a quasiparticle<br />

in a quantum state characterized by an energy E 0<br />

<strong>and</strong> wave function w 0 is obtained as the projection of the<br />

imaginary part of the self-energy R(r,r 0 ,E 0 ) onto the state<br />

itself (atomic units are used throughout, i.e., h = e 2 =<br />

m e = 1, unless otherwise is stated)<br />

Z Z<br />

C e–e ¼ s 1<br />

e–e ¼ 2 drdr 0 w 0 ðrÞImRðr; r0 ; E 0 Þw 0 ðr 0 Þ: ð1Þ<br />

Frequently the self-energy R is represented by the so-called<br />

GW approximation [25], which is the first term in a series<br />

expansion of R in terms of the screened Coulomb interaction<br />

W. Usually, the non-interacting Green function G 0 is<br />

used to replace the full one-electron Green function G. In<br />

the following we provide a brief description of the numerical<br />

methods used in the calculation of <strong>scattering</strong> rates in<br />

the two examples considered in this work, namely, quantum<br />

well states at metal surfaces <strong>and</strong> metal clusters. A more<br />

detailed account can be found elsewhere [26,27].<br />

2.1. Quantum well states<br />

In the case of Na adlayers on the Cu(111) surface, we<br />

assume that the charge density <strong>and</strong> the one-electron potential<br />

only vary along the z-direction perpendicular to the<br />

surface <strong>and</strong> are constant in the (x,y) plane parallel to the<br />

surface. This assumption is valid for Na induced QWS’s<br />

at coverage rates close to 1 ML, since the wave functions<br />

lie mainly in the Na layer <strong>and</strong> in the vacuum side, i.e., in<br />

a region with little potential variation in the direction parallel<br />

to the surface [28]. This leads to significant simplification<br />

of Eq. (1), which now takes a form<br />

C e–e ¼ s 1<br />

e–e ¼ 2 X 0<br />

f<br />

Z<br />

Z Z dqk<br />

ð2pÞ 2<br />

dzdz 0 / 0 ðzÞ/ f ðz0 Þ<br />

Im½ W ðz; z 0 ; q k ; E 0 E f ÞŠ/ 0 ðz 0 Þ/ f ðzÞ; ð2Þ<br />

where the summation is performed over all available final<br />

states f, W(z,z 0 ,q k ,x) is the two-dimensional Fourier transform<br />

of the screened Coulomb interaction, <strong>and</strong> / n (z) is the<br />

z-dependent component of the wave function w n (r):<br />

w n ðrÞ ¼p 1 ffiffiffi e iq kr k<br />

/ n ðzÞ<br />

ð3Þ<br />

S<br />

with S being a normalization area. To describe the Cu(111)<br />

surface covered by Na adlayers, we employ a slab containing<br />

31 atomic layers of Cu together with a region corresponding<br />

to Na <strong>and</strong> a vacuum region corresponding to 20<br />

Cu interlayer spacings. For the description of the surface<br />

electronic structure, the model potential of Ref. [17], based<br />

on the model potential of the bare Cu(111) surface [29], has<br />

been employed. This potential reproduces the QWS energy<br />

position of E QWS ¼ 0:127 eV at the surface Brillouin zone<br />

C<br />

center for 1 ML Na coverage [30]. To investigate the linewidth<br />

dependence on the QWS energy position relative to<br />

the Fermi level <strong>and</strong> demonstrate the dramatic change of<br />

the <strong>screening</strong> upon the QWS b<strong>and</strong> shrink below E F , we evaluate<br />

C e–e for the cases with QWS energies E QWS ¼ 0eV 1 <strong>and</strong><br />

E QWS<br />

C<br />

¼ 0:042 eV. We assume that the small variation of<br />

the QWS binding energy at the scale of 0.1 eV does not<br />

significantly affect the QWS wave function, <strong>and</strong> therefore<br />

use the same set of wave functions obtained for the<br />

E QWS<br />

C<br />

¼ 0:127 eV case, just shifting the QWS b<strong>and</strong> to the<br />

corresponding energy positions.We also assume here that<br />

1 As the QWS energy position at the C point is gradually moved down<br />

with increasing Na coverage [16] this energy position should roughly<br />

correspond to h 0.9.<br />

C


4548 V.M. Silkin et al. / Surface Science 601 (2007) 4546–4552<br />

the QWS wave functions do not change along the QWS<br />

energy b<strong>and</strong>, although this can have a significant effect on<br />

the excited electron linewidth in some cases [31,32]. For<br />

the QWS energy dispersion curve vs the two-dimensional<br />

wave vector k k , we use a parabolic form<br />

E QWS ðk k Þ¼E QWS þ k 2 C k =2mQWS , taking as effective mass the<br />

experimental value m QWS = 0.7 [30]. Further calculation<br />

details on the evaluation of the screened Coulomb interaction<br />

W(z,z 0 ,q k ,x) can be found in Refs. [33,34].<br />

2.2. Metal clusters<br />

The spherical jellium model [35,36] is used to represent<br />

metal clusters of variable size. The average one-electron<br />

radius r s in the cluster can be defined from r s = R/N 1/3 ,<br />

where R is the radius of the cluster <strong>and</strong> N the number of<br />

electrons in it. If the cluster is neutral, the equality<br />

1=n þ 0 ¼ 4pr3 s =3 holds as well, with nþ 0<br />

being the constant<br />

background positive charge of the cluster.<br />

We restrict our discussion to neutral clusters in closedshell<br />

configurations. The ground state properties of the<br />

cluster are described using density functional theory<br />

(DFT) <strong>and</strong> Kohn–Sham (KS) equations [37]. The exchange-correlation<br />

potential is calculated in the local density<br />

approximation (LDA) with the parametrization of<br />

Ref. [38]. The KS wave functions w(r) are calculated<br />

numerically after expansion in the spherical harmonic basis<br />

set Y lm (X r ). The set of KS equations are solved self-consistently<br />

using an iterative procedure.<br />

In real space, when the exact Green function G is replaced<br />

by the independent-electrons Green function G 0 in<br />

Eq. (1), one can show that<br />

C e–e ¼ 2 X Z<br />

0<br />

drdr 0 w 0 ðrÞw f ðr0 Þ<br />

f<br />

ImW ðr; r 0 ; xÞw 0 ðr 0 Þw f ðrÞ;<br />

where w 0 (r) is a KS wave function with eigenvalue E 0 ,<br />

x = E 0 E f , <strong>and</strong> the sum over f runs over all unoccupied<br />

KS states of eigenvalues E f < E 0 . Hence we are approximating<br />

the energy levels <strong>and</strong> wave functions in the decay<br />

by the eigenvalues <strong>and</strong> wave functions of unoccupied KS<br />

levels. This is a reasonable first-order approximation [39].<br />

In practice, the spherical symmetry of the problem simplifies<br />

the calculation [27]. All quantities in Eq. (4) are exp<strong>and</strong>ed<br />

in the spherical harmonics basis set. The screened<br />

interaction W(r,r 0 ,x) is calculated from the response function<br />

of the system v(r,r 0 ,x) obtained in the r<strong>and</strong>om phase<br />

approximation (RPA) [40,41]. The latter is built in a similar<br />

way to that of Ref. [42]: one-electron Green functions are<br />

built from the KS wave functions for every x. These Green<br />

functions are used to calculate the multipole components<br />

of the independent-electrons response function v 0 (r,r 0 ,x),<br />

<strong>and</strong> the RPA integral equation is solved in real space after<br />

matrix inversion for every polar component. Finally, the<br />

decay rate C e–e is calculated as a sum over the independent<br />

contributions of each of the available final states [27].<br />

ð4Þ<br />

3. Calculation results <strong>and</strong> discussion<br />

3.1. Quantum well states<br />

In Fig. 1, the calculated dependence of the linewidth C e–e<br />

on the excited energy is presented for three Na QWS energy<br />

positions. A large variation of the linewidth as a function of<br />

coverage is observed. A particularly strong enhancement of<br />

C e–e for small excitation energies <strong>and</strong> for the two systems<br />

with partially occupied QWS b<strong>and</strong> is observed, with respect<br />

to the case of unoccupied QWS. Additionally, one can note<br />

the different behavior between the two occupied QWS cases.<br />

For small energies, C e–e for E QWS ¼ 0:042 eV is larger than<br />

C<br />

C e–e for E QWS ¼ 0:127 eV. Indeed, the two curves cross<br />

C<br />

around E = 0.25 eV, <strong>and</strong> the linewidth for E QWS ¼ 0:127<br />

C<br />

starts to be the largest for higher energies. This finding is<br />

rather difficult to underst<strong>and</strong> in simple terms such as phase<br />

space available for the quasiparticle decay, since this phase<br />

space is the same for all three systems as is seen in Fig. 2.<br />

Note that these intrab<strong>and</strong> transitions within the QWS b<strong>and</strong><br />

itself are the main source for excited electron decay in all<br />

systems under study. Thus, for small energies this contribution<br />

accounts for almost 100% of the total decay rate. With<br />

the increase of the excitation energy the contribution from<br />

interb<strong>and</strong> transitions to the empty bulk electronic states increases<br />

faster than that from intrab<strong>and</strong> transitions. Nevertheless,<br />

even for energies around 2 eV the intrab<strong>and</strong><br />

contribution accounts for more than 90% of the total<br />

C e–e . Hence, the two-dimensional effects in Na/Cu(111) systems<br />

are much stronger than in the case of surface states at<br />

clean metal surfaces [8,43]. This fact can be explained by the<br />

larger decoupling of the QWS state wave functions from the<br />

bulk electron system [17].<br />

To underst<strong>and</strong> the behavior of C e–e as a function of the<br />

QWS energy position let us see Eq. (2), where the double<br />

integration in z <strong>and</strong> z 0 is performed. As mentioned above,<br />

Linewidth (meV)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0 0.5 1 1.5 2<br />

Energy (eV)<br />

Fig. 1. Calculated e–e contribution, C e–e , to the linewidth of a quantum<br />

well state in Na/Cu(111) for QWS energy position E QWS ¼ 0:127 eV<br />

C<br />

(solid line), E QWS ¼ 0:042 eV (dashed line), <strong>and</strong> E QWS ¼ 0 (dotted line).<br />

C<br />

C


V.M. Silkin et al. / Surface Science 601 (2007) 4546–4552 4549<br />

Fig. 2. Quantum well states energy dispersion for E QWS ¼ 0:127 eV<br />

C<br />

(solid line), E QWS ¼ 0:042 eV (dashed line), <strong>and</strong> E QWS ¼ 0 (dotted line).<br />

C<br />

C<br />

Grey area shows the regions of projected bulk electronic structure of bare<br />

Cu(111) surface being most efficient decay channel in Na/Cu(111)<br />

systems as well [18]. The two main decay channels, namely intrab<strong>and</strong> <strong>and</strong><br />

interb<strong>and</strong> transitions, are shown schematically by arrows.<br />

the QWS wave functions / 0 (z) <strong>and</strong> the bulk electronic structure<br />

for all the three different QWS energy positions have<br />

been hold the same in the present calculation. In addition<br />

to the above-mentioned similarity of the available phase<br />

space for decay, the remaining difference between these calculations<br />

is the screened Coulomb interaction W(z,z 0 ,q k ,x)<br />

originated from different QWS energy positions. We perform<br />

calculation of W(z,z 0 ,q k ,x) for all the three systems<br />

<strong>and</strong> analyze them below. First of all, we note a strong<br />

variation of ImW as a function of all its variables z, q k ,<br />

<strong>and</strong> x. To give a general idea of it in Fig. 3, we present<br />

ImW(z,z,q k ,x) as a function of z <strong>and</strong> x for the E QWS ¼<br />

C<br />

0:127 eV <strong>and</strong> E QWS ¼ 0:042 eV cases for two-dimensional<br />

momentum value q k = 0.05 a.u.<br />

1 . The use of<br />

C<br />

ImW(z,z,q k ,x) as an example is reasonable as inside the<br />

solid <strong>and</strong> near the surface ImW(z,z 0 ,q k ,x) has the maximum<br />

magnitude at z = z 0 <strong>and</strong> only far from the surface into<br />

the vacuum the maximum of this quantity occurs at z 5 z 0<br />

[44]. InFig. 3 one can observe a strong enhancement of<br />

ImW(z,z,q k ,x) for z P 0 <strong>and</strong> energies x 6 0.25 eV <strong>and</strong><br />

x 6 0.18 eV, respectively. Comparing both figures one<br />

can see that the amplitude of ImW in low left corner for<br />

the E QWS ¼ 0:042 eV case is much larger than for the<br />

C<br />

E QWS ¼ 0:127 eV case. Therefore, although the phase<br />

C<br />

space in the former case is smaller as one can appreciate<br />

in Fig. 5, this ImW enhancement leads to a larger linewidth<br />

for the E QWS ¼ 0:042 eV case for small energies. When the<br />

C<br />

energy increases, the relative role of this enhanced region<br />

of ImW decreases leading to a smaller linewidth for<br />

the E QWS ¼ 0:042 eV case, as compared to the E QWS ¼<br />

C<br />

C<br />

0:127 eV one. In Fig. 4, ImW(z,z,q k ,x) for E QWS ¼<br />

C<br />

0:127 eV <strong>and</strong> E QWS ¼ 0 are presented for two values of<br />

C<br />

q k , demonstrating that, at the surface, the amplitude of<br />

Fig. 3. Imaginary part of the screened interaction, ImW(z,z,q k ,x), for the<br />

E QWS ¼ 0:127 eV (upper panel) <strong>and</strong> E QWS ¼ 0:042 eV (lower panel)<br />

C<br />

C<br />

cases as a function of distance z <strong>and</strong> energy x for the two-dimensional<br />

momentum q k = 0.05 a.u.<br />

1 . Na adlayer is located at z = 0 <strong>and</strong> negative<br />

(positive) z values correspond to the solid (vacuum) side.<br />

ImW is two orders of magnitude larger than in the former<br />

case, whereas the latter is very similar to the clean Cu(111)<br />

case [44].<br />

Furthermore, in Fig. 3, one can see additional peaks at<br />

the maximum energy for QWS intrab<strong>and</strong> e–h excitations<br />

which correspond to the collective electronic excitations<br />

similar to the acoustic surface plasmon (ASP) at the clean<br />

metal surfaces [33,34]. Their dispersions closely follow the<br />

upper edge of regions for e–h excitations within the QWS<br />

b<strong>and</strong>s <strong>and</strong> are shown in Fig. 5. The origin of this collective<br />

mode is the coexistence at the surface of both the QWS <strong>and</strong><br />

bulk electrons. More detailed analysis of its properties will<br />

be given elsewhere. From Fig. 5, it is clear that the ASP<br />

contributes to the total linewidth C e–e . Nevertheless, it is<br />

difficult to separate this contribution from the very efficient<br />

decay channel due to intrab<strong>and</strong> QWS e–h excitations.<br />

3.2. Metal clusters<br />

One of the most appealing features of low-dimensional<br />

systems such as clusters is that their properties can often


4550 V.M. Silkin et al. / Surface Science 601 (2007) 4546–4552<br />

Fig. 5. Phase space for intrab<strong>and</strong> electron–hole excitations in the QWS<br />

b<strong>and</strong>. Grey colored area shows the region for the E QWS ¼ 0:127 eV case<br />

C<br />

<strong>and</strong> hatched area st<strong>and</strong>s for E QWS ¼ 0:042 eV. Intrab<strong>and</strong> bulk <strong>and</strong><br />

C<br />

interb<strong>and</strong> QWS-bulk electron–hole excitations are allowed everywhere.<br />

Thick solid <strong>and</strong> dashed lines on the left sides of each area show the<br />

dispersion of corresponding acoustic surface plasmons. As an example,<br />

thin dashed (dotted) line delimits the phase space region available for<br />

intrab<strong>and</strong> transitions for an excited electron of energy of 0.3 eV above E F<br />

for the E QWS ¼ 0:127 eV ðE QWS ¼ 0:042 eVÞ case.<br />

C<br />

C<br />

Fig. 4. Imaginary part of the screened interaction, ImW(z,z,q k ,x), as a<br />

function of z for q k = 0.05 a.u.<br />

1 (upper panel) <strong>and</strong> q k = 0.10 a.u.<br />

1 (lower<br />

panel). Thick (thin) lines st<strong>and</strong> for the E QWS ¼ 0:127ðE QWS ¼ 0Þ case.<br />

C<br />

C<br />

Solid, long dashed, dashed, dashed-dotted, <strong>and</strong> dotted lines show ImW<br />

for x = 0.06 eV, 0.12 eV, 0.18 eV, 0.24 eV, <strong>and</strong> 0.30 eV, respectively.<br />

be tuned by varying the size <strong>and</strong> thus they can be different<br />

from the bulk case. For this reason <strong>and</strong> in the following, we<br />

focus into clusters of variable size <strong>and</strong> fixed average electronic<br />

density (r s = 4, that represents for instance sodium<br />

clusters). In this respect, the difference between the dynamics<br />

of hot electrons in clusters <strong>and</strong> in bulk has been discussed<br />

bringing up two different effects. First, the electron<br />

lifetime can be enhanced in the cluster because of the discretization<br />

of levels that reduces the number of final states<br />

to which the electronic excitation can decay. Then again,<br />

the electron lifetime can decrease in the cluster because<br />

of the reduction of dynamic <strong>screening</strong> in the proximity of<br />

the cluster surface. We show in the following that the<br />

<strong>screening</strong> in clusters is not only modified due to the presence<br />

of the surface but also due to the different mobility<br />

of the cluster electrons with respect to those in bulk. The<br />

dynamically screened interaction between the cluster electrons<br />

is affected by this factor, in particular for low excitation<br />

energies.<br />

In Fig. 6, we show the linewidth C e–e of electron excitations<br />

with energy E 0 1 eV measured from the Fermi level<br />

of the cluster. Due to the discretization of levels in the cluster,<br />

the initial energy of the excitation cannot be fixed exactly<br />

to 1 eV, but the eigenvalue closest in energy is<br />

chosen. In any case, the variation with respect to the reference<br />

value is never larger than 10%. We come back to this<br />

Fig. 6. Linewidth (in meV) of electron excitations with energy E 0 1eV<br />

above the Fermi level of the cluster, as a function of the cluster radius (in<br />

a.u.). All systems have r s = 4. The value of C e–e obtained from an RPA<br />

calculation in an homogeneous electron gas with the same parameter r s is<br />

also shown.


V.M. Silkin et al. / Surface Science 601 (2007) 4546–4552 4551<br />

point below. C e–e is plotted as a function of the cluster radius<br />

R. For small cluster sizes, the discretization of levels is<br />

clearly observable in the large oscillations found. For clusters<br />

of nanometer size, however, these oscillations are<br />

damped <strong>and</strong> the value of the linewidth is C e–e 150 meV.<br />

Let us also mention that an RPA calculation of the electron<br />

lifetime in a homogeneous electron gas, with the same<br />

parameters that we fix in our calculation (E 0 = 1 eV <strong>and</strong><br />

r s = 4) gives a value C e–e = 60 meV [45], smaller than the<br />

ones obtained for these big clusters.<br />

One of the reasons not often discussed in the literature<br />

for the difference between the value of C e–e in bulk <strong>and</strong> in<br />

nanometer-sized clusters is the difference in the electron–<br />

electron screened interaction W(r,r 0 ,x) between an infinite<br />

electron gas <strong>and</strong> a finite jellium system, even for values of r<br />

<strong>and</strong> r 0 far from the cluster surface. In Fig. 7, we quantify<br />

this statement by showing the imaginary part of<br />

ImW(r,r 0 ,x) for a value of r nearby the cluster center<br />

(r = 0.31) <strong>and</strong> as a function of r 0 . The dimension of the<br />

chosen is sufficiently big (R = 38.8 a.u.) to consider that<br />

no surface effects are relevant. ImW is plotted for two different<br />

values of the energy, namely x = 0.815 meV <strong>and</strong><br />

x = 54.4 eV. The former is representative of actual values<br />

of x that enter the calculation of C e–e in clusters. Fig. 7<br />

shows that ImW(0, r 0 ,x = 0.815 eV) is clearly different in<br />

a cluster as compared to the bulk counterpart. The dynamic<br />

<strong>screening</strong> between electrons is thus modified due to<br />

the different mobility of electrons in a confined system.<br />

However, for higher values of x, the response of the system<br />

is similar to that calculated in the independentelectron<br />

approximation v 0 <strong>and</strong> finite-size effects are not<br />

relevant.<br />

Confinement effects are also relevant in the choice of the<br />

initial state in the decay process. The discretization of levels<br />

in the cluster makes it impossible to exactly define the same<br />

initial energy E 0 for all the cluster sizes considered. An estimate<br />

of the variation that this feature can introduce in the<br />

linewidth value is also shown in Fig. 6. For a cluster size of<br />

R = 31 a.u., the value of C e–e for two initial states very<br />

close in energy is plotted. One of them is the first l =13<br />

state with E 0 = 1 eV, <strong>and</strong> the second one is a l = 4 state<br />

with E 0 = 0.99 eV. In spite of the proximity of the two energy<br />

levels, the value of C e–e varies as much as 35%. The<br />

main reason for that is the different coupling between the<br />

initial state wave function <strong>and</strong> the available final states<br />

for the decay. In Fig. 8, we show the normalized radial distribution<br />

of the KS wave functions rR l;E0 ðrÞ, where<br />

w 0 ðrÞ ¼R l;E0 ðrÞY lm ðX r Þ, for the two cases considered. The<br />

integrated radial distributions<br />

QðrÞ ¼<br />

Z r<br />

0<br />

<br />

dr 0 r 02 R l;E0 ðr 0 2<br />

Þ<br />

ð5Þ<br />

are plotted in Fig. 8 as well. While the l = 13 wave function<br />

does not have any nodes, oscillations in the l = 4 wave<br />

function are conspicuous. A proper definition of the initial<br />

excitation in the decay process can thus be crucial when<br />

calculating the linewidth.<br />

Fig. 7. Imaginary part of the self-energy ImW(r 0 ,r 0 ,x) (in eV) for a<br />

cluster of N = 912 <strong>and</strong> r s =4(R = 38.8 a.u.), as a function of the radial<br />

coordinate r 0 (in a.u.), for two different values of the energy x = 0.815 eV<br />

<strong>and</strong> 54.4 eV. The coordinate r 0 is fixed to 0.31 a.u. The corresponding<br />

values for ImW in an homogeneous electron gas obtained from Mermin<br />

response functions are also plotted for comparison.<br />

Fig. 8. Normalized radial distribution of wave functions with l =4, 13<br />

<strong>and</strong> E 0 = 0.99, 1 eV, respectively, corresponding to a cluster with N = 440<br />

<strong>and</strong> r s =4 (R = 30.42 a.u.). E 0 is measured from the Fermi level of the<br />

cluster. The integrated value of the wave functions is plotted in the upper<br />

panel.


4552 V.M. Silkin et al. / Surface Science 601 (2007) 4546–4552<br />

4. Summary <strong>and</strong> conclusions<br />

Low-dimensional systems bear electronic properties that<br />

can be very different from those well studied in bulk systems.<br />

Here, we have shown that the modification of the dynamic<br />

<strong>screening</strong> in the electron–electron interaction <strong>and</strong><br />

the change in the structure of energy levels in these systems<br />

can combine to drastically modify the linewidth of electronic<br />

excitations as compared to the bulk reference. In<br />

particular, we have analyzed the screened interaction between<br />

electrons <strong>and</strong> the phase space availability for the decay<br />

in two different cases: quantum well states in overlayers<br />

of Na over Cu(111) <strong>and</strong> gas-phase metal clusters of sizes<br />

up to few nanometers. In the former case, the amount of<br />

alkali atoms coverage on the metal surface leads to a shift<br />

of the QWS energy b<strong>and</strong>, that produces an appreciable<br />

change in the linewidth. In the latter case, the variation<br />

in size of the cluster is responsible for modifications in<br />

the decay rate. Our detailed analysis of these properties<br />

shows that a proper choice of the system characteristics<br />

should allow to tune the excitation linewidths due to electron–electron<br />

<strong>scattering</strong> processes.<br />

Acknowledgements<br />

Financial support by the Basque Departamento de Educación,<br />

Universidades e Investigación, the University of<br />

the Basque Country UPV/EHU (Grant No. 9/UPV<br />

00206.215-13639/2001), the Spanish MEC (Grant No.<br />

FIS2004-06490-C03-00), <strong>and</strong> the EU Network of Excellence<br />

NANOQUANTA (Grant No. NMP4-CT-2004-<br />

500198) is acknowledged.<br />

References<br />

[1] M. Aeschlimann, M. Bauer, S. Pawlik, W. Weber, S. Burgermeister,<br />

D. Oberli, H.C. Siegmann, Phys. Rev. Lett. 79 (1997) 5158.<br />

[2] H. Petek, S. Ogawa, Prog. Surf. Sci. 56 (1997) 239.<br />

[3] E. Knoesel, A. Hotzel, M. Wolf, Phys. Rev. B 57 (1998) 12812.<br />

[4] R. Knorren, K.H. Bennemann, R. Burgermeister, M. Aeschlimann,<br />

Phys. Rev. B 61 (2000) 9427.<br />

[5] E.V. Chulkov, A.G. Borisov, J.P. Gauyacq, D. Sánchez-Portal, V.M.<br />

Silkin, V.P. Zhukov, P.M. Echenique, Chem. Rev. 106 (2006) 4160.<br />

[6] U. Höfer, I.L. Shumay, Ch. Reuß, U. Thomann, W. Wallauer, T.<br />

Fauster, Science 277 (1997) 1480.<br />

[7] M. Weinelt, J. Phys.: Condens. Matter 14 (2002) R1099.<br />

[8] P.M. Echenique, R. Berndt, E.V. Chulkov, Th. Fauster, A. Goldmann,<br />

U. Höfer, Surf. Sci. Rep. 52 (2004) 219.<br />

[9] J. Kröger, L. Limot, H. Jensen, R. Berndt, S. Crampin, E. Pehlke,<br />

Prog. Surf. Sci. 80 (2005) 26.<br />

[10] J. Güdde, U. Höfer, Prog. Surf. Sci. 80 (2005) 49.<br />

[11] A.M. Bradshow, H.P. Bonzel, G. Ertl, Physics <strong>and</strong> Chemistry of<br />

Alkali Metal Adsorption, Elsevier, Amsterdam, 1989.<br />

[12] R.D. Diehl, R. McGrath, Surf. Sci. Rep. 23 (1996) 43.<br />

[13] We define one Na monolayer as the most densely packed structure of<br />

the first Na layer, namely a (3/2 · 3/2) unit mesh with 4 Na atoms per<br />

9 Cu surface atoms: D. Tang, D. McIlroy, X. Shi, C. Su, D. Heskett,<br />

Surf. Sci. 255 (1991) L497;<br />

J. Kliewer, R. Berndt, Surf. Sci. 477 (2001) 250.<br />

[14] S.-Å. Lindgren, L. Walldén, Solid State Commun. 34 (1980)<br />

671.<br />

[15] S.-Å. Lindgren, L. Walldén, Phys. Rev. Lett. 59 (1987) 3003.<br />

[16] N. Fischer, S. Schuppler, Th. Fauster, W. Steinmann, Surf. Sci. 314<br />

(1994) 89.<br />

[17] E.V. Chulkov, J. Kliewer, R. Berndt, V.M. Silkin, B. Hellsing, S.<br />

Crampin, P.M. Echenique, Phys. Rev. B 68 (2003) 195422.<br />

[18] C. Corriol, V.M. Silkin, D. Sánches-Portal, A. Arnau, E.V. Chulkov,<br />

P.M. Echenique, T. von Hofe, J. Kliewer, J. Kröger, R. Berndt, Phys.<br />

Rev. Lett. 95 (2005) 176802.<br />

[19] C. Voisin, D. Christofilos, N. Del Fatti, F. Vallée, B. Prével, E.<br />

Cottancin, J. Lermé, M. Pellarin, M. Broyer, Phys. Rev. Lett. 85<br />

(2000) 2200.<br />

[20] N. Pontius, G. Lüttgens, P.S. Bechthold, M. Neeb, W. Eberhardt, J.<br />

Chem. Phys. 115 (2001) 10479.<br />

[21] P. Gerhardt, M. Niemietz, Y. Dok Kim, G. Ganteför, Chem. Phys.<br />

Lett. 382 (2003) 454.<br />

[22] M. Drescher, M. Hentschel, R. Kienberger, M. Uiberacker, V.<br />

Yakovlev, A. Scrinzi, Th. Westerwalbesloh, U. Kleineberg, U.<br />

Heinzmann, F. Krausz, Nature 419 (2002) 803.<br />

[23] A. Baltus˘ka, Th. Udem, M. Uiberacker, M. Hentschel, E. Goulielmakis,<br />

Ch. Gohle, R. Holzwarth, V.S. Yakovlev, A. Scrinzi, T.W.<br />

Hänsch, F. Krausz, Nature 421 (2003) 611.<br />

[24] A. Borisov, D. Sánchez-Portal, R. Díez Muiño, P.M. Echenique,<br />

Chem. Phys. Lett. 387 (2004) 95.<br />

[25] L. Hedin, S. Lundqvist, Solid State Phys. 23 (1969) 1.<br />

[26] P.M. Echenique, J.M. Pitarke, E.V. Chulkov, A. Rubio, Chem. Phys.<br />

251 (2000) 1.<br />

[27] M. Quijada, R. Díez Muiño, P.M. Echenique, Nanotechnology 16<br />

(2005) S176.<br />

[28] J.M. Carlsson, B. Hellsing, Phys. Rev. B 61 (2000) 13973.<br />

[29] E.V. Chulkov, V.M. Silkin, P.M. Echenique, Surf. Sci. 437 (1999)<br />

330.<br />

[30] J. Kliewer, R. Berndt, Phys. Rev. B 65 (2001) 035412.<br />

[31] L. Vitali, P. Wahl, M.A. Schneider, K. Kern, V.M. Silkin, E.V.<br />

Chulkov, P.M. Echenique, Surf. Sci. 523 (2003) L47.<br />

[32] M.G. Vergniory, J.M. Pitarke, S. Crampin, Phys. Rev. B 72 (2005)<br />

193401.<br />

[33] V.M. Silkin, A. García-Lekue, J.M. Pitarke, E.V. Chulkov, E.<br />

Zaremba, P.M. Echenique, Europhys. Lett. 66 (2004) 260.<br />

[34] V.M. Silkin, J.M. Pitarke, E.V. Chulkov, P.M. Echenique, Phys. Rev.<br />

B 72 (2005) 115435.<br />

[35] M. Brack, Rev. Mod. Phys. 65 (1993) 677.<br />

[36] J.A. Alonso, L.C. Balbás, Topics in Current Chemistry, vol. 182,<br />

Springer-Verlag, Berlin/Heidelberg, 1996, p. 119.<br />

[37] W. Kohn, L.J. Sham, Phys. Rev. 140 (1965) A1133.<br />

[38] O. Gunnarsson, B.I. Lundqvist, Phys. Rev. B 13 (1976)<br />

4274.<br />

[39] P. Fulde, Electron Correlations in Molecules <strong>and</strong> Solids, Springer-<br />

Verlag, Berlin, 1995 (Section 9.2.3).<br />

[40] W. Ekardt, Phys. Rev. B 31 (1985) 6360.<br />

[41] M.J. Puska, R.M. Nieminen, M. Manninen, Phys. Rev. B 31 (1985)<br />

3486.<br />

[42] R. Díez Muiño, A. Arnau, A. Salin, P.M. Echenique, Phys. Rev. B 68<br />

(2003) 041102(R).<br />

[43] J. Kliewer, R. Berndt, E.V. Chulkov, V.M. Silkin, P.M. Echenique, S.<br />

Crampin, Science 288 (2000) 1399.<br />

[44] P.M. Echenique, J. Osma, V.M. Silkin, E.V. Chulkov, J.M. Pitarke,<br />

Appl. Phys. A 71 (2000) 503.<br />

[45] J.M. Pitarke, Private Communication, 2004.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!