11
This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail. Powered by TCPDF (www.tcpdf.org) This material is protected by copyright and other intellectual property rights, and duplication or sale of all or part of any of the repository collections is not permitted, except that material may be duplicated by you for your research use or educational purposes in electronic or print form. You must obtain permission for any other use. Electronic or print copies may not be offered, whether for sale or otherwise to anyone who is not an authorised user. Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study of adsorbed metallic quantum dots: Na on Cu(111 Published in: Physical Review B DOI: 10.1103/PhysRevB.66.235420 Published: 30/12/2002 Document Version Publisher's PDF, also known as Version of record Please cite the original version: Torsti, T., Lindberg, V., Puska, M. J., & Hellsing, B. (2002). Model study of adsorbed metallic quantum dots: Na on Cu(111. Physical Review B, 66(23), 1-10. [235420]. https://doi.org/10.1103/PhysRevB.66.235420

Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

This is an electronic reprint of the original article.This reprint may differ from the original in pagination and typographic detail.

Powered by TCPDF (www.tcpdf.org)

This material is protected by copyright and other intellectual property rights, and duplication or sale of all or part of any of the repository collections is not permitted, except that material may be duplicated by you for your research use or educational purposes in electronic or print form. You must obtain permission for any other use. Electronic or print copies may not be offered, whether for sale or otherwise to anyone who is not an authorised user.

Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B.Model study of adsorbed metallic quantum dots: Na on Cu(111

Published in:Physical Review B

DOI:10.1103/PhysRevB.66.235420

Published: 30/12/2002

Document VersionPublisher's PDF, also known as Version of record

Please cite the original version:Torsti, T., Lindberg, V., Puska, M. J., & Hellsing, B. (2002). Model study of adsorbed metallic quantum dots: Naon Cu(111. Physical Review B, 66(23), 1-10. [235420]. https://doi.org/10.1103/PhysRevB.66.235420

Page 2: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

PHYSICAL REVIEW B 66, 235420 ~2002!

Model study of adsorbed metallic quantum dots: Na on Cu„111…

T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and B. Hellsing41Laboratory of Physics, Helsinki University of Technology, P.O. Box 1100, FIN-02015 HUT, Finland

2CSC - Scientific Computing Ltd., P.O. Box 405, FIN-02101 Espoo, Finland3Department of Physics, Va¨xjo University, SE-35195 Va¨xjo, Sweden

4Department of Experimental Physics, Chalmers and Go¨teborg University, SE-412 96 Go¨teborg, Sweden~Received 25 July 2002; published 30 December 2002!

We model electronic properties of the second-monolayer Na adatom islands~quantum dots! on the Cu~111!surface covered homogeneously by the first Na monolayer. An axially symmetric three-dimensional jelliummodel, taking into account the effects due to the first Na monolayer and the Cu substrate, has been developed.The electronic structure is solved within the local-density approximation of the density-functional theory usinga real-space multigrid method. The model enables the study of systems consisting of thousands of Na atoms.The results for the local density of states are compared with differential conductance (dI/dV) spectra andconstant current topographs from scanning tunneling microscopy.

DOI: 10.1103/PhysRevB.66.235420 PACS number~s!: 68.65.Fg, 68.37.Ef, 68.65.Hb

inid

troeestu

p

avca

inuc

snst

ceali

boc-

lleo

,s

arown

neasethenges

lec--y a

enftfor

Naent

ofi-

he

vepro-uc-

le

ortonin

umpari-VI

yer

I. INTRODUCTION

At certain faces of metals, such as the~111! face of noblemetals, the surface electron states are confined to the vicof the top layer by the vacuum barrier on the vacuum sand the bandgap on the substrate side.1 The electrons in thesesurface states form a two-dimensional nearly-free-elecgas.2–5 It has also been observed that when adsorbing onseveral monolayers of alkali-metal atoms on these surfacmanifold of discrete standing-wave states, so called quanwell states, perpendicular to the surface are formed.6,7 Thesestates can be detected, for instance, in photoemission stroscopy ~PES!,8 inverse photoelectron spectroscopy,9 two-photon photoemission spectroscopy10 ~2PPES! and scanningtunneling microscopy~STM!.11 A large amount of experi-mental data is available for the system Na on Cu~111!.6–11

The electronic structure and dynamics for this system halso been investigated by first-principles theoreticalculations.12,13

These localized surface states are of great interest sthey play an important role in many physical processes sas epitaxial growth,14 surface catalysis,15,16 molecularordering,17 and adsorption.18 Experimental tools such aSTM and PES play an important role in the investigatiosince they enable spatial and spectroscopic resolving ofelectron states.

One important discovery is the confinement of surfastate electrons in so called quantum corrals. These man-mnanoscale structures are formed by deliberately assembadatoms to enclosed structures by STM.19 Due to the smallsize of the corrals, quantum effects are present, andspatial and spectroscopic properties of the confined statesbe studied experimentally. A natural way of forming lowdimensional structures on metal surfaces is by controgrowth of epitaxial layers. With an appropriate choicedeposition and annealing temperatures small islandscalled quantum dots~QD!, with variable shapes and sizemay form.20 The advantage of these structures, in compson with the corrals, is that they are relatively stable at ltemperatures. This enables the imaging and investigatio

0163-1829/2002/66~23!/235420~10!/$20.00 66 2354

itye

nto, am

ec-

el

ceh

,he

-deng

than

dfso

i-

of

their properties without inducing structural damage. Oquantum-mechanical effect of the confinement is the increof the surface-state energy, which in turn may lead todepopulation of the surface-state band and thereby chain the surface properties.

In this paper, we present calculated results for the etronic structure of Na on Cu~111!, with the emphasis on describing the real-space resolved density of states nearbsodium QD adsorbed on a sodium-covered Cu~111! surface.Previously, an all-electron density-functional theory~DFT!study of a free-standing Na layer in vacuum has bepresented.21 More recently, a DFT calculation using ultrasopseudopotentials for the free-standing Na layer as well asthe layers adsorbed on Cu~111! have been presented.12 Asimple free-electron model calculation for a free-standingQD has already been published by two of the presauthors.22

The calculations in this work are made in the contextthe DFT.23,24 More specifically, the Rayleigh-quotient multgrid ~RQMG! method in axial symmetry26,25 is used for thenumerical solution of the ensuing Kohn-Sham equations. Telectron-ion interaction is simplified using the jelliummodel,23 where the ions are replaced by a rigid positibackground charge of constant density. This model hasvided basic physical understanding of the electronic strtures of simple metal surfaces,23 thin metal films,28 vacanciesand voids inside metals,29 and finite clusters of simple metaatoms.30 Recently, also uniform cylindrical nanowires havbeen studied within the jellium model.31–33

The paper is organized as follows: Sec. II gives a shreview over experimental results for the system NaCu~111!, Sec. III describes the computational method usedthe calculations, Sec. IV discusses the details of our jellimodel, and in Sec. V the results are presented and comsons are made with experimental findings. Finally, Sec.gives the conclusions.

II. Na ON Cu „111…

Alkali metals adsorbed on the closed-packed~111! surfaceof metals form hexagonal structures at saturated monola

©2002 The American Physical Society20-1

Page 3: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

at

m

of

thmdche

nt

thithstanth

te

ow.ddioetenkyin

oe

up-

ea

recu--the

leof

iclexi-,

el

inthis

-po-

m-ac-s

Kus,

m-

ntis-

ible

thelf-

nce.

ityan

T. TORSTIet al. PHYSICAL REVIEW B 66, 235420 ~2002!

coverages, following approximately the underlying substrstructure.34 The first monolayer of Na on Cu~111! is observedto saturate at the coverage ofQ54/9'0.44,35 correspondingto four Na atoms per nine surface-Cu atoms. The Na atothus, form a hexagonal (3/233/2) structure and the Na atomspacing of 7.43a0 is comparable to the atomic distance6.92a0 in bulk Na.

The adsorption of Na atoms on the Cu~111! surface willinduce a charge redistribution at the interface betweenadlayer and the substrate. It has been seen from photoesion experiments10,36that when the Na coverage is increasethe Cu Shockley surface state decreases in energy. Forerages aboveQ'0.11, the surface state is shifted below tlower band edge of the local band gap of the Cu~111! sur-face, and is no longer visible in photoemission experimeTwo-photon photoemission experiments10 indicate other un-occupied Na-induced states in the local band gap atCu~111! surface, which will also decrease in energy wincreasing Na coverage. For higher coverages, the lowethese states will be downshifted below the Fermi energy,thus get occupied. At the saturated monolayer coverage,state will be located about 0.1 eV~Refs. 10,12, and 36! be-low the Fermi energy at theG-point of the surface Brillouinzone. The corresponding next lowest state will be loca2.1 eV above the Fermi level10 at theG point. The lower ofthe two states has one node in thez direction, while thehigher has two.

If the Na-atom deposition continues after the first monlayer ~ML ! is completed, a second layer will start to groRecent STM measurements37,38 indicate that the seconmonolayer of Na grows via the formation of compact islanwith hexagonal atomic arrangement. Normal photoemissexperiments8,39 indicate that when the second monolaygrows the emission intensity due to one-monolayer stadecreases gradually, and for coverages above 1.3 ML, apeak'0.1 eV above the Fermi energy appears. This peaascribed to the two-monolayer thick parts, and the energshifted to somewhat lower values as the coverage iscreased.

Some theoretical attention has also been paid to NaCu~111!, including the island growth. Free-electron modcalculations have been performed for circular Na~Ref. 22! aswell as hexagonal Ag~Refs. 40 and 41! and Na~Ref. 42!free-standing islands. All-electron calculations for an unsported monolayer21 of Na and first-principles slab calculations for one-atomic-Na layer in (232) and (3/233/2) ad-sorbate structures12 on Cu~111! have been presented. In thpresent paper, we report jellium model calculations forunsupported monolayer of Na and a cylinder shaped fstanding Na QD. We also present two-density-jellium callations for the system Na on Cu~111!, where we have modeled the underlying Cu~111! substrate by using a lowerdensity slab to mimic the decay of the surface states intosubstrate. Comparison is made with experiments and prous theoretical calculations.

III. COMPUTATIONAL METHODS

In the Kohn-Sham scheme of DFT, one solves the etron densityn(r ) of the system self-consistently from a set

23542

e

s,

eis-,ov-

s.

e

ofdis

d

-

sn

rs

ewisis-

nl

-

ne--

evi-

c-

equations. One of these equations is the single-partSchrodinger equation. The models used in this work are aally symmetric. Thus, the Schro¨dinger equation is separableand the wave functions can be written as products,

cmkn~r ,z,f!5eimfUmkn~r ,z!. ~1!

Above, m is the azimuthal quantum number implied by thaxial symmetry whilen differentiates between orthogonastates with samem andk. In the calculations involving theinfinite monolayer, twok-vectors are used as explainedSec. IV. The external potential of the systems studied inwork is caused by the positive background chargen1(r ).The effective potentialVeff includes also the Hartree potential of the electron density and the exchange-correlationtential VXC , which we treat in the local-densityapproximation.49 The electron densityn(r ) is obtained bysumming single-electron densities with the occupation nubers f mkn . The degeneracies of the states are taken intocount by the factor (22d0m) and the occupation numberf mkn obey the Fermi-Dirac statistics with a Fermi level (EF)so that the system is neutral. A finite temperature of 1200is used to stabilize the solution of the set of equations. Thin the present axial symmetry@r5(r ,z)#, the Kohn-Shamequations read as

21

2 S 1

r

]

]r1

]2

]r 22

m2

r 2

]2

]z212Ve f fD Umkn~r !

5«mknUmkn~r ! ~2!

n~r !52(mkn

~22d0m! f mknuUmkn~r !u2, ~3!

Veff~r !5F~r !1VXC~r !, ~4!

S 1

r

]

]r1

]2

]r 21

]2

]z2D F524p@n2~r !2n1~r !#. ~5!

The Schro¨dinger equation~2! is solved using the RQMGmethod,26 which has been implemented in various geoetries, including the axial symmetry.26,25 In the RQMGmethod, the Rayleigh quotient^cuHuc&/^cuc& on the finestlevel grid is directly minimized, the orthogonality constraibeing taken into account by a penalty functional. The Poson equation~5! is solved for the electrostatic potentialF(r )using a standard multigrid method.43

To obtain self-consistency, we use the simplest posspotential mixing scheme,

Vini 115AVout

i 1~12A!Vini . ~6!

The largest system of this work contains 2550 electrons,diameter of the supercell being 170 Å. Obtaining seconsistency in such a system requires a very smallA value of0.005. Otherwise, the charge sloshing results in divergeMore sophisticated mixing strategies27 will be indispensablein the future calculations. However, because of the simplicof our model systems, we can very accurately estimate

0-2

Page 4: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

gs

nddrd

al

cr-raFnalthton

eso

asiskcihairethhthaoo

N

oxc

otrnel-b

sngeraca

low.andno-the

or.urlis-theted

resese

nsl abe

ap-itzs,ringof

in

MODEL STUDY OF ADSORBED METALLIC QUANTUM . . . PHYSICAL REVIEW B66, 235420 ~2002!

initial guess for the self-consistent effective potential of larsystems using the more easily convergent smaller systemreference.

In our largest calculation, a grid of 319395 points is usedfor the presentation of the wave functions, potential, adensity. Taking into account the unoccupied states needethe modeling, up to 2400 different states have to be solveevery self-consistency iteration. Luckily, it is straightforwato parallelize the calculation over the 65 differentm valuesand the twok points ~see below!. Moreover, the RQMGmethod26 handles this part of the calculation with optimefficiency.

The two-jellium model for the surface described in SeIV results in an asymmetric density distribution with a suface dipole. Thus, the electrostatic potential on the substside is a constant different from that on the vacuum side.the Poisson equation, we thus use the boundary conditiozero derivative on the substrate side, and that of zero von the vacuum side. Solving the Poisson equation,boundaries above and below the system are extendeddistance five times greater than in the case of the wave futions.

In this work, we calculate the local density of stat~LDOS! above a surface at distances corresponding to thtypical in STM measurements~of the order of 20a0). Atsuch distances, the amplitude of the wave function decreby several orders of magnitude. This kind of modelingthus a serious test for the RQMG method. We have checthe accuracy of our method for the spherical harmonic oslator and a model hydrogen atom potential, for which twave-functions are known analytically. The evanescent tof the wave functions solved with the RQMG method agwith the analytical ones, even when the amplitude ofwave functions has dropped by 20 orders of magnitude. Tlevel of accuracy is beyond the reach of plane-wave meods, where periodic boundary conditions are necessary,which provide a uniform accuracy across the calculation vume, resulting in spurious oscillations in the vacuum partsthe system.

IV. MODELING THE SYSTEM

We are interested in the system of a monolayer-thickQD on the complete Na monolayer on the Cu~111! surface.We know from experiments that these islands are apprmately hexagonal in shape, following the underlying struture of the Na monolayer.

In order to interpret recent STM data for these typessystems, mapping the energy resolved real-space elecdensity near a QD is necessary. First, it is of interest to fiwhat level of theoretical modeling is required. It has beshown previously40–42 that simple two-dimensiona‘‘particle-in-a-box’’ calculations give qualitatively good results, in the sense that the peak structure of LDOS resemspectra obtained in the STMdI/dV measurements. In thiwork, we improve the theoretical description by performiself-consistent three-dimensional DFT calculations, whthe effects of the underlying monolayer and substrateintroduced. The hexagonal QD is modeled by a cylindri

23542

eas

dinat

.

teorofueea

c-

se

es

edl-elseeis-

ndl-f

a

i--

fondn

les

erel

jellium QD, and the underlying Na monolayer and Cu~111!substrate by the two-density-jellium slab, as described beComparisons between calculations of free-standing QD’sQD’s on a substrate show indeed that the underlying molayer and substrate induce a new type of states thatsimple particle-in-a-box calculations cannot account fSince thez dependence of wave functions is included in ocalculations, we can calculate the tunneling current at reatic STM-tip distances above the system, and estimateenergy dependence of the step height from the calculaconstant current topographs.

The different model systems studied in this work ashown in Fig. 1. Our model is readily applicable to the caof a free-standing cylindrical quantum dot, where we uzero Dirichlet boundary conditions for the wave functioand for the Coulomb potential. The next step is to modefree-standing monolayer. A uniform planar system cannotexactly reproduced in the axial symmetry. We adopt anproximation scheme analogous to the Wigner-Semethod.44 We imagine the plane being filled by hexagonand then approximate these hexagons by area-covecircles. In order to sample the Brillouin-zone of the latticecircles we use twok-points,k50 andk at the Brillouin-zone

FIG. 1. Profile of the axially symmetric background chargethe case of~a! jellium model for free-standing Na quantum dot~b!jellium model for free-standing Na monolayer~c! two-jelliummodel for Na monolayer on Cu~111! ~d! two-jellium model for Naquantum dot on Na monolayer on Cu~111!

0-3

Page 5: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

le

th

e

thu

hl isineb

ers

rsak

e

eeoreio

ofr-theird

fthetheeri-rmienicther of

Nato

ap-f asen

re-

as afnly

n be1.1

mnd

eroare

ofrre-ith

ing

theza-

ple

me

tothe

ec-

ge

T. TORSTIet al. PHYSICAL REVIEW B 66, 235420 ~2002!

boundary. The wave functions withk50 are required tohave a vanishing radial derivative at the radius of the circwhereas the wave functions withk at the Brillouin-zoneboundary vanish there. According to our calculations,model gives a uniform (r independent! charge distributionfor the monolayer. It also minimizes the interactions betwea QD inside a circle with its periodic images.45

The next step is to place the Na monolayer on top ofCu~111! substrate. The effect of the substrate is modeleding the two-jellium model, which is illustrated in Figs. 1~c!and 2. We do not model the electrons of the bulk Cu. Tdensity of electrons per unit area in the two-jellium modekept the same as in the jellium model for a free-standmonolayer. We add a layer of lower-density jellium, in ordto mimic the different wave-function decays into the sustrate and into the vacuum. The thicknessw2 and density~viar s2) give two free parameters of the lower-density jelliumwhich we adjust in order to reproduce the relevant expmental values of the first and second surface band bottomthe coverages of 1 ML and 2 ML, respectively. Here the fiand second bands correspond to wave functions with onetwo nodes in the vertical direction, respectively. The thicness w18 of the higher-density jellium in the two-jelliummodel is given by

w185w12S r s1

r s2D 3

w2 , ~7!

wherew1 is the thickness of the free-standing Na monolay

A. The underlying monolayer and substrate

To test our model, we first study the systems of a frstanding Na monolayer and that of a Na monolayerCu~111!, and compare the results with other theoreticalsults and experimental findings. The Na jellium densitydetermined from the bulk nearest-neighbor distance

FIG. 2. One complete monolayer of Na on Cu~111! surfacewithin the two-jellium model. The positive background char~shaded areas!, electron density~dashed line!, effective potential~solid line!, and electrostatic potential~dash-dotted line! are shown.The shading corresponds to Fig. 1.

23542

,

e

n

es-

e

gr-

,i-attnd-

r.

-n-

sf

6.92a0 and the experimental height of 5.5a0 ~2.9 Å! of 1 MLof Na on Cu~111!.38,42 The resulting density parameterr s53.79a0 gives a slightly higher density than its bulk value3.93a0 for Na. The thickness and the density of the lowedensity slab have been chosen by fitting the bottom ofsecond band for the 1-ML-Na coverage and that of the thband for the 2-ML-Na coverage on Cu~111! to the experi-mental values.10,36,39 The values of r s256.0a0 and w256.3a0 give ~using the unit cell of radius 72.8a0 containing400 electrons per monolayer! in the 1-ML case the bottom othe second band at 75 meV below the Fermi level and in2-ML case, the bottom of the third band at 50 meV aboveFermi level. These values are reasonably close to the expmental values of about 100 meV below and above the Felevel, respectively.8,36,39The correct positions relative to thFermi level are important because we solve for the electrostructures self-consistently, so that the occupancies ofsingle-electron states affect the potential and the charactethe states themselves.

B. The quantum dot

We start by studying a free-standing monolayer-thickQD @Fig. 1~a!#, since this system shows close resemblancethe simple particle-in-a-box system often used as a firstproximation when describing the electronic structure oQD on a surface. The number of atoms in the QD is choto 100, which corresponds to a QD radius of aboutR536.40a0. The uppermost panel in Fig. 3 shows the corsponding energy spectrum relative to the Fermi energyEF .Note that the the discrete energy eigenvalues are plottedfunction of the quantum numberm and not as a function ok. There are three bands below the vacuum level, but othe first band~no horizontal nodal planes! is occupied. Theemergence of the succeeding second and third bands caseen as the condensation of the energy levels at aroundeV and 2.5 eV, respectively.

The LDOS calculated at the cylinder axis at the jelliuedge and at 8a0 above the edge are shown in the lowest athe middle panel of Fig. 3, respectively. Only states withm50 contribute, since they are the only ones with nonzcontributions at the axis. The discrete energy levelsbroadened to Lorenzians with the widthG58 meV. TheLDOS at the jellium edge can easily be resolved in termsthe contributions from the different bands: The peaks cosponding to first, second, and third bands form series wquadratically increasing intervals and smoothly increaspeak amplitudes. At the distance of 8a0 above the QD, thecontribution due to the first-band states is diminished andcontribution due to the third-band states with high quantition in thez direction~two horizontal nodal planes! is domi-nating the LDOS. Comparison with the results of the simparticle-in-a-box calculation by Lindberg and Hellsing22

shows that these two calculations give qualitatively the saresults.

We now compare these results for the free-standing QDthe system of the QD adsorbed on a Na monolayer onCu~111! surface@Fig. 1~d!#. In our calculation, the substratis a two-jellium cylindrical supercell containing 400 ele

0-4

Page 6: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

anegn

heiu

l-

ypcul

a

ntireve as.n

er-cal

onle

theOS

edD.

QDre

nge

thes in

ithadsing

ne

en-iththe

st-lab

s in

ingate,

bi-

Na

Shes aghtly

s-e

airsand

fith a

ee

f tey

he

MODEL STUDY OF ADSORBED METALLIC QUANTUM . . . PHYSICAL REVIEW B66, 235420 ~2002!

trons. The energy spectrum is shown in the uppermost pof Fig. 4, in which the differentm levels corresponding to thk50 points are given with thek dispersion calculated usinthe k point at the Brillouin zone boundary. In comparisowith the spectrum of the monolayer-thick QD in Fig. 3, tbands are shifted downwards because of the larger jellthickness at the QD. The lowest-energy states~in the bulb ofthe level diagram! have nok-dispersion and they are locaized at the QD and the substrate slab below it~see Fig. 5!.Their dispersion as a function ofm is similar to that in Fig. 3.Introducing the underlying substrate gives rise to a new tof states, which are not localized to the QD region. In fathese states form the overwhelming majority. As a resnew bands are induced in the energy spectrum and they

FIG. 3. The uppermost panel: Energy spectrum of the frstanding cylindrical jellium QD@Fig. 1~a!# containing 100 elec-trons. The discrete eigenenergies are shown as a function oquantum numberm. The top of the figure corresponds to thvacuum level. The middle panel: The LDOS calculated at the cinder axis at 8a0 above the jellium edge. The lowest panel: TLDOS calculated at the cylinder axis at the jellium edge.

23542

el

m

et,t,re

less dispersive as a function ofm. The reducedm dispersionreflects the fact that the states are extended over the ecircular supercell. For the same reason, these bands halarger dispersion in thek space than the localized QD bandA simple particle-in-a-box or free-standing QD calculatiocannot provide states of this kind, and it is, therefore, intesting to see up to what extent they contribute to the loelectronic structure above the QD.

The LDOS for the QD adsorbed on a Na monolayerthe Cu~111! surface is given in the lowest and the middpanels of Fig. 4 at the jellium edge and at 14a0 above theedge, respectively. In order to avoid complications due tointeractions between the supercells, we calculate the LDusing only thek50 states and the LDOS is then calculatas in the case of the free-standing monolayer-thick Na QTo enable a thorough comparison with the free-standingresults, we have to study first the wave functions in modetail.

Figure 5 gives all the wave functions in the interestienergy region for them50 states of the QD adsorbed on thNa monolayer. The first three states,n51,2, and 3, corre-spond to states within the first band. They are localized toQD and the substrate slab below it and they have no nodethe z direction. Then54 state shows another character wa density no longer localized to the QD region but sprealso to the slab region around. This is the first state belongto the new type of bands induced by the slab. Staten58 isthe beginning of the next band consisting of states with onode in thez direction~second band in the QD!. These statesare resonance states, the amplitude of which is stronglyhanced in the QD region, but due to the hybridization wthe delocalized slab states they are actually delocalized towhole system. Then514 state starts the next band consiing of delocalized states with one horizontal node in the sregion, i.e., it is a second band in the slab. Finally, staten516 represents the first-resonance state with two nodethe z direction ~third band in the QD!. In Fig. 5, one notesthat the statesn516 and 17 and also the statesn519 and 20form pairs. The state lower in energy in the pair is a bondcombination of a QD state and a surrounding slab stwhereas the state higher in energy is an antibonding comnation.

The electronic structure of the QD adsorbed on themonolayer, discussed above in terms of them50 wave func-tions, is reflected in the LDOS in Fig. 4. First, in the LDOat the jellium edge, we notice that in comparison with tfree-standing QD model, the underlying slab introducenew type of bands and squeezes the other bands more titogether to fit more states below the Fermi level~the lowestpanels of Figs. 3 and 4!. Therefore, the present LDOS lookqualitatively different from that of the monolayer-thick unsupported QD in Fig. 3. Moreover, the hybridization of thQD and surrounding slab states to bonding-antibonding pcauses the splitting of peaks seen clearly for the third-bstates at the distance of 14a0 above the jellium edge~themiddle panel of Fig. 4!. If we had a continuous spectrum oslab states, we would have a single resonance peak wfinite energy width.

-

he

l-

0-5

Page 7: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

eethdts

ioc

celaeist

ina

e is

ithofhe

facecur-nceode,the

hentheur-

he

Na

lofCucal

ns

dis-as

ureL

soed

the

td inthe

sys-ance

la

icl:

th

T. TORSTIet al. PHYSICAL REVIEW B 66, 235420 ~2002!

Having the modeling of the STM results in mind, thinteresting question arising is whether or not the localizstates calculated by the free-standing QD model givesame LDOS far above the QD as the states in the moincluding the substrate slab. Studying the LDOS plomatching each peak with the corresponding wave functwe notice that the resonance states with strongly enhanamplitude in the QD region are dominant at large distanabove the QD. The contribution of the more delocalized sstates is small. Therefore, the free-standing QD model ispected to preserve validity in predicting LDOS at large dtances above the QD. The too broad energy spectrum infree-standing QD model can be corrected for by increasthe dot height with a monolayer of Na jellium or withtwo-jellium layer.

FIG. 4. The uppermost panel: Energy spectrum of the systema QD containing 100 electrons on top of a two-density-jellium sdescribed by a supercell of 400 electrons@Fig. 1~d!#. The k50eigenenergies are given by thick horizontal bars. The thin vertbars indicate the dispersion in thek-space. The middle paneLDOS calculated at the cylinder axis at 14a0 above the jelliumedge. The lowest panel: LDOS calculated at the cylinder axis atjellium edge.

23542

deel,

n,edsbx--heg

V. COMPARISON WITH EXPERIMENT

One of the most useful instruments in surface sciencthe scanning tunneling microscope~STM!.46,47It can be usedfor measuring the real-space electronic distribution watomic resolution, as well as the local energy distributionelectrons and the lifetimes of excited electron states. Treal-space distribution is achieved when scanning the sureither in the constant current mode, where the tunnelingrent is kept constant by changing the tip-surface distausing a feedback mechanism, or in the constant height mwhere the tunneling current is measured when scanningsurface at a constant tip height. The resulting image tdisplays the topography of the surface. Information aboutlocal electronic structure is obtained by measuring the crent variation with the applied voltage. This quantity, tdifferential conductancedI/dV, is proportional to the prod-uct of the local density of states~LDOS! and the transmis-sion coefficientT.48,50 However, if the applied voltage issmall, the bias dependence ofT is small, and with Eq. 3, weapproximately have

dI

dV}(

mkn~22d0m!uUmkn~r !u2d~emkn2eV!. ~8!

In the STM study by Kliewer and Berndt,42 constant currenttopographs anddI/dV measurements are presented for aisland on Na monolayer on Cu~111!. The size of the island is2303170a0

2 (120390 Å2). We have studied a cylindricajellium dot with similar dimensions, i.e., having the radius85a0 and containing thus about 550 electrons. The Na/substrate is described in our calculations by a cylindritwo-density-jellium supercell with the radius of 160a0 andcontaining 2000 electrons. The radius of 85a0 is actuallyfixed to reproduce the peak structure ofdI/dV spectra byKliewer and Berndt42 as well as possible~See Fig. 6 anddiscussion below!. For comparison, Kliewer and Berndt42

used in their modeling two-dimensional hard-wall hexagowith the radius of 83a0.

We show in Fig. 6 the LDOS at 18a0 above the jelliumedge, both at thez axis of the QD, and atr 520a0 away fromthe axis. The height corresponds to a typical tip-sampletance in the STM experiments. The LDOS is calculatedfor the smaller systems in Sec. IV B. The peaks in the figcorrespond to states with two horizontal nodes in the 2-Mpart of the system~third band in the QD!. At the axis, onlythe m50 states contribute, while away from the axis alpeaks withmÞ0 occur. The LDOS peaks can be labelwith the ‘‘quantum number’’N by counting for the numberof radial nodes of the corresponding wave functions in2-ML part ~See Fig. 6!. For m50, the N51 state has noradial nodes in the dot region, whereasN52 has one radialnode and so on for largerN. As in the case of the smaller dodiscussed above in Sec. IV B, the states strongly peakethe QD are resonance states due to the hybridization withstates of the surrounding monolayer and span the wholetem. Besides the delocalization of the states, the resoncharacter causes the fact that in the LDOS~Fig. 6!, several

ofb

al

e

0-6

Page 8: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

sth

on

er

ai

Dthroce

lche

d

akstita-al-ner-in

t of

at. 6.

y inthellerhem-

ishe-

hehestleartate-

l-

le

ugwckTax

m

e

MODEL STUDY OF ADSORBED METALLIC QUANTUM . . . PHYSICAL REVIEW B66, 235420 ~2002!

peaks may correspond to the same resonance state acussed earlier for the smaller system. We have identifiedLDOS peaks by examining the wave functions. The horiztal lines below the quantum numbersm and N connect thepeaks belonging to the resonance in question.

The relative positions of peaks appearing in the expmentaldI/dV spectra by Kliewer and Berndt42 are shown inFig. 6 as arrows pointing downwards. The experimental dis shifted so that the lowest experimental peak coincides wthe lowest calculated peak. The experimental spectrumrecorded slightly off from the center of the hexagonal Qwhich should be taken into account when comparing withcalculated results. Previous two-dimensional free-electcalculations for hexagonal potential boxes have reproduwell the experimental peak positions.41,42In the present mod-eling, the experimental peak positions agree with the calated m50 resonance positions with the exception of t

FIG. 5. Wave functions for the 21 lowest lyingm50, k50states for the system of a monolayer-thick QD, containing 100 etrons, on top of a Na/Cu slab~400 electrons per supercell!. Thewave functions are plotted in a plane parallel to the z axis throthe center of the QD. In each subfigure, the cylinder axis is shoby the solid vertical line. The shading indicates the positive baground charge and the dashed vertical line points its QD edge.upper, lower and right-hand subfigure borders and the cylinderlimit the computation volume with the dimension of (73360)a0

2.

23542

dis-e-

i-

tathis,end

u-

third and fifth experimental peak. The calculatedm50 reso-nances obey the pattern«n5E01AN2 as would be expectedfor a free particle in a hard-wall cylinder. We have fitteE0'22 meV,A'15 meV. It is gratifying to note that in theLDOS recorded off the cylinder axis, strongm51 resonancepeaks appear so that the third and fifth experimental pecan be explained. Thus, our model can reproduce quantively the experimental peak positions. According to our cculations, the resonance width increases toward higher egies. The increase is maybe slightly stronger thanexperiment, indicating a somewhat too weak confinementhe resonance states in our model.

We have also calculated the isosurfaces of the LDOSthe energies corresponding to the dominant peaks in FigThe results are shown along with the total electron densitFig. 7. In order to see clearly the nodal structures ofdifferent states, the LDOS is calculated using a smaLorenzian width of 0.8 meV. The density is smooth in tinterior of the QD and shows minor oscillations at the perieter of the QD. The development of the nodal structureclear and compares qualitatively well with that found in texperimentaldI/dV maps.38 It can also be seen that the isosurfaces corresponding to the two (m,N)5(0,4) peaks differfrom each other mainly near the perimeter of the dot. Thigher peak shown at the energy corresponding to the hig~0,3! state, on the other hand, does not show equally c~0,3! character. The explanation is the appearance of a swith quantum numbers~2,2! at almost exactly the same energy.

From Eq.~8!, we obtain a simple formula for the tunneing current

I ~U,r ,z!}EEF

EF1U

r~E,r ,z!dE, ~9!

c-

hn-heis

FIG. 6. Cylindrical QD containing 550 electrons on two-jelliusubstrate. The local density of states is shown at 18a0 above jelliumedge at the axis~solid line! and atr 520a0 ~dashed line! away fromthe axis. Lorenz broadening withG58 meV has been used. Threlative experimental peak positions42 are given by vertical arrowspointing downwards. The peaks are identified with (m,N) reso-nance states having two horizontal node planes in the QD.

0-7

Page 9: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

rmcaQ

s aas

tetu-

cu-OSthetantrs-

g. 8be

are

ys-teparttheteri-lowiaseralthat

ightingro-ac-the

-

sgs

is

mined

d

T. TORSTIet al. PHYSICAL REVIEW B 66, 235420 ~2002!

wherer(E,r ,z) is the LDOS at the heightz and distancerfrom the axis. Calculated isosurfaces of this quantity fonumerical constant current topographs, from which weestimate the apparent step height at the perimeter of the

FIG. 7. Cylindrical QD of 550 electrons on two-jellium substrate. Isosurfaces of the electron density~top! and the LDOS~withG50.8 meV) at energies corresponding to the dominant peakFig. 6. The quantum numbers of the dominant states contributineach energy is indicated. The isovalue for each plot is chosen avalue of the corresponding quantity at 18a0 above the jellium edgeand 30a0 off from the axis. The height-to-radius ratio in the plotsexaggerated.

23542

nD.

The inset of Fig. 8 shows the constant LDOS height afunction of the distance from the cylinder axis for the bivoltages of2400, 0, and1400 meV. The value of LDOS is10211 arbitrary units on the scale of Fig. 6. The absoluheight of the isosurface from the jellium edge depends narally on the LDOS value chosen but according to our callations, the relative changes are insensitive to the LDvalue over a wide range of values. In order to constructapparent step height, we first obtain the numerical conscurrent topographs@Eq. ~9!#, then average the profiles ovethe oscillations above the 2-ML and 1-ML parts of the sytem and take the difference. The results are shown in Fias a function of the bias voltage. The trends seen canexplained by studying the LDOS isosurfaces~see the inset ofFig. 8!, and then noting that the constant current surfacesobtained by simple integation@Eq. ~9!#. At 2100 meV, thesecond band starts to contribute in the 1-ML part of the stem raising the height there and thereby lowering the sheight. Then the onset of the third band in the 2-ML praises the step again. This rise is similar to that seen inexperiment by Kliewer and Berndt38 as well as the decline ahigher voltages. However, the comparison with the expmental result shows differences: our step height is tooby a factor of 2, and the raising of the step at negative bvoltages is not seen in the experiment. There may be sevreasons for the differences in the step heights. One isthe experimental step height of 5.5a0 ~2.9 Å!, which isdetermined at a voltage just before the rise in the step heis directly used as the thickness of the jellium describthe second monolayer of Na. A more consistent pcedure might be to take the voltage dependence intocount. Moreover, the apparent step height depends on

ofatthe

FIG. 8. Cylindrical QD containing 550 electrons on two-jelliusubstrate. The step height of the second Na monolayer determfrom calculated constant current surfaces@Eq. ~9!# is shown as afunction of the bias voltage~energy relative to the Fermi level!. Theinset shows the LDOS isosurface profiles~height-to-radius ratio ex-aggerated! at energies2400 meV, ~dashed line! 0 meV ~solidline!, and 400 meV~dash-dotted line!. The height is measured fromthe jellium edge of the second-ML QD. The width of the plotteregion is 160a0, i.e., the radius of the circular supercell.

0-8

Page 10: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

aneipl

nbesy-ri

icaCf topon

thru

nt

is-tes,enteat thetheion

ESrousut-

y of

MODEL STUDY OF ADSORBED METALLIC QUANTUM . . . PHYSICAL REVIEW B66, 235420 ~2002!

relative vacuum decay rates of the second and third-bstates of the 1-ML and 2-ML systems, respectively. Thcorrect description may be too demanding for our simmodel.

VI. CONCLUSIONS

In this paper, we have presented a model for the electrostructures of alkali-metal islands or quantum dots adsoron metal surfaces. In particular, we have focused on thetem of Na on the Cu~111! surface, where approximately hexagonal Na quantum dots have been observed to form duthe epitaxial growth of the second-Na monolayer.

We have modeled the quantum dots as small cylindrjellium islands, and the underlying Na monolayer andsubstrate as a two-density-jellium slab. The parameters omodel have been chosen to fit experimental spectroscdata and calculated first-principles band structures forand two completed monolayers of Na on the Cu~111! sur-face. The calculations were performed in the context ofdensity-functional theory using a real-space electronic stture calculation method.

The calculated results are compared with experime

,

ei

e,

.

23542

dre

icds-

ng

luheice

ec-

al

findings from scanning tunneling microscope and photoemsion experiments. The model gives local densities of stawhich are in a quantitative agreement with constant currtopographs anddI/dV spectra and maps. Thereby, the idof surface states, which are localized as resonances aquantum dots is supported. The future applications ofmodel will include studies of the adsorption and dissociatof molecules in the vicinity of alkali-metal quantum dots.

ACKNOWLEDGMENTS

We would like to thank Lars Wallde´n and S. Å. Lindgrenfor sharing their knowledge on the system based on Pexperiments. We would like to thank R.M. Nieminen fomany useful discussions. We acknowledge the genercomputer resources from the Center for Scientific Comping, Espoo, Finland. One of the authors~T.T.! acknowledgesfinancial support by the Vilho, Yrjo¨ and Kalle Vaisala foun-dation. This research has been supported by the AcademFinland through its Centers of Excellence Program~2000–2005!.

ure

J.

ys.

aperB

i.

1H. Luth, Surfaces and Interfaces of Solids~Springer-Verlag, Ber-lin, 1993!.

2P.O. Gartland and B.J. Slagsvold, Phys. Rev. B12, 4047~1975!.3P. Heimann, H. Neddermeyer, and H.F. Roloff, J. Phys. C10, L17

~1977!.4A. Zangwill Physics at Surfaces~Cambridge University Press

Cambridge, 1988!.5N. W. Ashcroft and N. D. Mermin,Solid State Physics~Saunders,

Philadelphia, 1976!.6S.-A. Lindgren and L. Wallde´n, Solid State Commun.34, 671

~1980!.7S.-A. Lindgren and L. Wallde´n, Phys. Rev. Lett.59, 3003~1987!.8A. Carlsson, B. Hellsing, S.-A˚ . Lindgren, and L. Wallde´n, Phys.

Rev. B56, 1593~1997!.9R. Dudde, L.S.O. Johansson, and B. Reihl, Phys. Rev. B44, 1198

~1991!.10N. Fischer, S. Schuppler, R. Fischer, Th. Fauster, and W. St

mann, Phys. Rev. B43, 14 722~1991!.11J. Kliewer, R. Berndt, E.V. Chulkov, V.M. Silkin, P.M. Echeniqu

and S. Crampin, Science288, 1399~2000!.12J.M. Carlsson and B. Hellsing, Phys. Rev. B61, 13 973~2000!.13B. Hellsing, J. Carlsson, L. Wallde´n, and S.-A˚ . Lindgren, Phys.

Rev. B61, 2343~2000!.14N. Memmel and E. Bertel, Phys. Rev. Lett.75, 485 ~1995!.15E. Bertel, P. Roos, and J. Lehmann, Phys. Rev. B52, R14 384

~1995!.16J.V. Lauritsen, S. Helveg, E. Lægsgaard, I. Stengaard, B

Clausen, H. Topso”, and F. Besenbacher, J. Catal.197, 1 ~2001!.17S.J. Stranick, M.M. Kamna, and P.S. Weiss, Science266, 99

~1994!.18E. Bertel, Phys. Status Solidi A159, 235 ~1997!.

n-

S.

19M.F. Crommie, C.P. Lutz, and D.M. Eigler, Science262, 218~1993!.

20H. Roder, E. Hahn, H. Brune, J.-P. Bucher, and K. Kern, Nat~London! 366, 141 ~1993!.

21E. Wimmer, J. Phys. F: Met. Phys.13, 2313~1983!.22V. Lindberg and B. Hellsing, Surf. Sci.506, 297 ~2002!.23N. D. Lang, inTheory of the Inhomogeneous Electron Gas, edited

by S. Lundqvist and N. H. March~Plenum, New York, 1983!, p.309.

24R.O. Jones and O. Gunnarsson, Rev. Mod. Phys.61, 689 ~1989!.25T. Torsti, M. Heiskanen, M. J. Puska, and R. M. Nieminen, Int.

Quantum Chem.,~to be published!; e-print cond-mat/0205056.26M. Heiskanen, T. Torsti, M.J. Puska, and R.M. Nieminen, Ph

Rev. B63, 245106~2001!.27For a short review and a promising new idea, see the recent p

by D. Raczkowski, A. Canning, and L.W. Wang, Phys. Rev.64, 121101~2001!.

28F.K. Schulte, Surf. Sci.55, 427 ~1976!.29M. Manninen, R.M. Nieminen, P. Hautoja¨rvi, and J. Arponen,

Phys. Rev. B12, 4012~1975!.30W.A. de Heer, Rev. Mod. Phys.65, 611 ~1993!; M. Brack, ibid.

65, 677 ~1993!.31N. Zabala, M.J. Puska, and R.M. Nieminen, Phys. Rev. Lett.80,

3336 ~1998!.32N. Zabala, M.J. Puska, and R.M. Nieminen, Phys. Rev. B59,

12 652~1999!.33M.J. Puska, E. Ogando, and N. Zabala, Phys. Rev. B64, 033401

~2001!.34R. Diehl and R. McGrath, Surf. Sci. Rep.23, 43 ~1996!.35D. Tang, D. McIlroy, X. Shi, C. Su, and D. Heskett, Surf. Sc

Lett. 255, L497 ~1991!.36S.-A. Lindgren and L. Wallde´n, Phys. Rev. B38, 3060~1988!.

0-9

Page 11: Torsti, T.; Lindberg, V.; Puska, M.J.; Hellsing, B. Model study ......Model study of adsorbed metallic quantum dots: Na on Cu—111– T. Torsti,1,2 V. Lindberg,3 M. J. Puska,1 and

ev

ss

. B

T. TORSTIet al. PHYSICAL REVIEW B 66, 235420 ~2002!

37J. Kliewer and R. Berndt, Surf. Sci.477, 250 ~2001!.38J. Kliewer and R. Berndt, Phys. Rev. B65, 035412~2001!.39A. Carlsson, S.-A˚ . Lindgren, C. Svensson, and L. Wallde´n, Phys.

Rev. B50, 8926~1994!.40J. Li, W.-D. Schneider, R. Berndt, and S. Crampin, Phys. R

Lett. 80, 3332~1998!.41J. Li, W.-D. Schneider, S. Crampin, and R. Berndt, Surf. Sci.422,

95 ~1999!.42J. Kliewer and R. Berndt, Appl. Phys. A: Mater. Sci. Proce

A72, S155~2001!.43A. Brandt, Math. Comput.31, 333 ~1977!.44C. Kittel, Introduction to Solid State Physics, 7th ed.~Wiley, New

York, 1996!, pp. 248–252.

23542

.

.

45G. Makov, R. Shah, and M.C. Payne, Phys. Rev. B53, 15 513~1996!; T. Korhonen, M.J. Puska, and R.M. Nieminen,ibid. 54,15 016~1996!.

46G. Binnig and H. Rohrer, Helv. Phys. Acta55, 726 ~1982!.47J. Tersoff and D.R. Hamann, Phys. Rev. B31, 805 ~1985!.48A. Selloni, P. Carnevali, E. Tosatti, and C.D. Chen, Phys. Rev

31, 2602~1985!.49D.M. Ceperley and B.J. Alder, Phys. Rev. Lett.45, 566

~1980!; J.P. Perdew and A. Zunger, Phys. Rev. B23, 5048~1981!.

50J. Li, W.-D. Schneider, and R. Berndt, Phys. Rev. B56, 7656~1997!.

0-10