density-functional theory with screened van der waals ......density-functional theory with screened...

5
Density-Functional Theory with Screened van der Waals Interactions for the Modeling of Hybrid Inorganic/Organic Systems Victor G. Ruiz 1 , Wei Liu 1 , Egbert Zojer 2 , Matthias Scheffler 1 , and Alexandre Tkatchenko 1 1 Fritz–Haber–Institut der Max–Planck–Gesellschaft, Faradayweg 4-6, 14195, Berlin, Germany 2 Institute of Solid State Physics, Graz University of Technology, Petersgasse 16, A–8010 Graz, Austria The electronic properties and the function of hybrid inorganic/organic systems (HIOS) are in- timately linked to their interface geometry. Here we show that the inclusion of the many–body collective response of the substrate electrons inside the inorganic bulk enables us to reliably predict the HIOS geometries and energies. This is achieved by the combination of dispersion-corrected density-functional theory (the DFT+vdW approach) [Phys. Rev. Lett. 102, 073005 (2009)], with the Lifshitz-Zaremba-Kohn theory for the non–local Coulomb screening within the bulk. Our method yields geometries in remarkable agreement (0.1 ˚ A) with normal incidence x–ray standing wave measurements for the 3,4,9,10–perylene–tetracarboxylic acid dianhydride (C24O6H8, PTCDA) molecule on Cu(111), Ag(111), and Au(111) surfaces. Similarly accurate results are obtained for xenon and benzene adsorbed on metal surfaces. Hybrid inorganic/organic systems (HIOS) are essential ingredients of a wide range of emerging devices. Appli- cations in which such interfaces play a crucial role in- clude (opto)–electronics (e.g., transistors, light–emitting diodes), organic photovoltaics, and sensors [1]. The in- terface geometry of HIOS plays a crucial role in the deter- mination of their electronic properties, and the accurate prediction of interface structure and stability is essential for controling the function and quality of these highly sought-after technologies. Within the possible variety of HIOS, 3,4,9,10– perylene–tetracarboxylic acid dianhydride (C 24 O 6 H 8 , PTCDA) on coinage metals is one of the best experi- mentally and theoretically characterized systems. The adsorption geometry of PTCDA on Cu(111), Ag(111), and Au(111), has been accurately determined by means of the normal incidence x–ray standing wave (NIXSW) technique [2–4], making these systems a suitable choice for testing the predictive power of theoretical methods. The reliable prediction of the equilibrium structure and dynamic properties of HIOS from first principles repre- sent a great challenge to the state–of–the–art theoretical methods due to the interplay of covalent interactions, electron transfer processes, van der Waals (vdW) inter- actions, and Pauli repulsion. In particular, vdW interac- tions are fundamental in determining the structure and stability of organic molecules on solid surfaces [5–10]. Romaner et al. studied the adsorption of PTCDA on coinage metals [11], and concluded that a more accurate approach to include vdW interactions is needed. To illustrate the problem, we start by comparing the performance of existing theoretical methods for the in- teraction of PTCDA with Ag(111) in Fig. 1. The av- erage adsorption distance (2.86 ˚ A) is reliably known from a number of NIXSW studies for the ordered mono- layer at room temperature (see Ref. 4 and references therein). There is also an estimate of the adsorption energy of 2.4 eV by extrapolation from the NTCDA molecule, which is closely related to PTCDA and ex- perimentally accessible to temperature-programmed des- orption [10, 12]. Using standard density-functional the- ory (DFT) with the Perdew–Burke–Ernzerhof (PBE) [13] functional resuls in no visible minimum in the binding curve [10]. The local–density approximation (LDA) un- derestimates the binding distance (2.65 ˚ A), although it (fortuitously) yields a better adsorption energy [14]. The inclusion of vdW interactions using the vdW-DF non- local functional [15] leads to results far off both for the binding distance and the binding energy [10, 11]. In- teratomic pairwise PBE-D correction by Grimme [16] appears to be closer to the experimental binding dis- tance; the calculated binding energy is, however, by 1 eV larger than the above mentioned estimate. The in- teratomic pairwise PBE+vdW correction scheme [17], where the vdW coefficients and radii are determined non- empirically from the electron density, leads to a better prediction for the energy but it still overestimates the equilibrium distance by about 0.25 ˚ A. Even the com- putationally most expensive calculations using exact ex- change with electron correlation treated in the random- phase approximation (with some additional approxima- tions to make these calculations feasible) yield 0.2 ˚ A overestimation in the equilibrium distance of PTCDA on Ag(111) [18]. In this Letter, we propose a method that extends stan- dard pairwise vdW corrections [16, 17] to the model- ing of adsorbates on surfaces. Here, this is achieved by combining the DFT+vdW scheme [17] with the Lifshitz- Zaremba-Kohn (LZK) theory for the vdW interaction be- tween an atom and a solid surface [19, 20]. In our ap- proach (DFT+vdW surf ), the vdW energy correction to the DFT total energy is given by a sum of C ab 6 R -6 ab terms, where R ab are the distances between atoms a and b, in analogy to the DFT-D and DFT+vdW methods. By em- ploying the LZK theory, however, we include the many– body collective response (screening ) of the substrate elec- trons in the determination of the C 6 coefficients and vdW radii, going effectively beyond the pairwise description.

Upload: others

Post on 14-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Density-Functional Theory with Screened van der Waals Interactions for theModeling of Hybrid Inorganic/Organic Systems

    Victor G. Ruiz1, Wei Liu1, Egbert Zojer2, Matthias Scheffler1, and Alexandre Tkatchenko11Fritz–Haber–Institut der Max–Planck–Gesellschaft, Faradayweg 4-6, 14195, Berlin, Germany

    2Institute of Solid State Physics, Graz University of Technology, Petersgasse 16, A–8010 Graz, Austria

    The electronic properties and the function of hybrid inorganic/organic systems (HIOS) are in-timately linked to their interface geometry. Here we show that the inclusion of the many–bodycollective response of the substrate electrons inside the inorganic bulk enables us to reliably predictthe HIOS geometries and energies. This is achieved by the combination of dispersion-correcteddensity-functional theory (the DFT+vdW approach) [Phys. Rev. Lett. 102, 073005 (2009)],with the Lifshitz-Zaremba-Kohn theory for the non–local Coulomb screening within the bulk. Ourmethod yields geometries in remarkable agreement (≈ 0.1 Å) with normal incidence x–ray standingwave measurements for the 3,4,9,10–perylene–tetracarboxylic acid dianhydride (C24O6H8, PTCDA)molecule on Cu(111), Ag(111), and Au(111) surfaces. Similarly accurate results are obtained forxenon and benzene adsorbed on metal surfaces.

    Hybrid inorganic/organic systems (HIOS) are essentialingredients of a wide range of emerging devices. Appli-cations in which such interfaces play a crucial role in-clude (opto)–electronics (e.g., transistors, light–emittingdiodes), organic photovoltaics, and sensors [1]. The in-terface geometry of HIOS plays a crucial role in the deter-mination of their electronic properties, and the accurateprediction of interface structure and stability is essentialfor controling the function and quality of these highlysought-after technologies.

    Within the possible variety of HIOS, 3,4,9,10–perylene–tetracarboxylic acid dianhydride (C24O6H8,PTCDA) on coinage metals is one of the best experi-mentally and theoretically characterized systems. Theadsorption geometry of PTCDA on Cu(111), Ag(111),and Au(111), has been accurately determined by meansof the normal incidence x–ray standing wave (NIXSW)technique [2–4], making these systems a suitable choicefor testing the predictive power of theoretical methods.

    The reliable prediction of the equilibrium structure anddynamic properties of HIOS from first principles repre-sent a great challenge to the state–of–the–art theoreticalmethods due to the interplay of covalent interactions,electron transfer processes, van der Waals (vdW) inter-actions, and Pauli repulsion. In particular, vdW interac-tions are fundamental in determining the structure andstability of organic molecules on solid surfaces [5–10].Romaner et al. studied the adsorption of PTCDA oncoinage metals [11], and concluded that a more accurateapproach to include vdW interactions is needed.

    To illustrate the problem, we start by comparing theperformance of existing theoretical methods for the in-teraction of PTCDA with Ag(111) in Fig. 1. The av-erage adsorption distance (2.86 Å) is reliably knownfrom a number of NIXSW studies for the ordered mono-layer at room temperature (see Ref. 4 and referencestherein). There is also an estimate of the adsorptionenergy of 2.4 eV by extrapolation from the NTCDAmolecule, which is closely related to PTCDA and ex-

    perimentally accessible to temperature-programmed des-orption [10, 12]. Using standard density-functional the-ory (DFT) with the Perdew–Burke–Ernzerhof (PBE) [13]functional resuls in no visible minimum in the bindingcurve [10]. The local–density approximation (LDA) un-derestimates the binding distance (≈ 2.65 Å), although it(fortuitously) yields a better adsorption energy [14]. Theinclusion of vdW interactions using the vdW-DF non-local functional [15] leads to results far off both for thebinding distance and the binding energy [10, 11]. In-teratomic pairwise PBE-D correction by Grimme [16]appears to be closer to the experimental binding dis-tance; the calculated binding energy is, however, by 1eV larger than the above mentioned estimate. The in-teratomic pairwise PBE+vdW correction scheme [17],where the vdW coefficients and radii are determined non-empirically from the electron density, leads to a betterprediction for the energy but it still overestimates theequilibrium distance by about 0.25 Å. Even the com-putationally most expensive calculations using exact ex-change with electron correlation treated in the random-phase approximation (with some additional approxima-tions to make these calculations feasible) yield 0.2 Åoverestimation in the equilibrium distance of PTCDA onAg(111) [18].

    In this Letter, we propose a method that extends stan-dard pairwise vdW corrections [16, 17] to the model-ing of adsorbates on surfaces. Here, this is achieved bycombining the DFT+vdW scheme [17] with the Lifshitz-Zaremba-Kohn (LZK) theory for the vdW interaction be-tween an atom and a solid surface [19, 20]. In our ap-proach (DFT+vdWsurf), the vdW energy correction tothe DFT total energy is given by a sum of Cab6 R

    −6ab terms,

    where Rab are the distances between atoms a and b, inanalogy to the DFT-D and DFT+vdW methods. By em-ploying the LZK theory, however, we include the many–body collective response (screening) of the substrate elec-trons in the determination of the C6 coefficients and vdWradii, going effectively beyond the pairwise description.

  • FIG. 1. Adsorption energy Eads as a function of verticaldistance d for PTCDA on Ag(111) employing different the-oretical approaches. The estimated adsorption energy for thesystem (2.4±0.1 eV [10]) and the experimental adsorptiondistance (2.86±0.05 Å [4]) are indicated by shaded intervals.These error bars correspond to typical experimental error es-timates.

    Interface polarization effects are accounted for via the in-clusion of semi–local hybridization due to the dependenceof the Cab6 interatomic coefficients on the electron den-sity in the DFT+vdW method. We show that non–localscreening reduces the vdW C6 coefficients by up to a fac-tor of four for coinage metals. As indicated by the solidblue line in Fig. 1 for the PTCDA/Ag(111) interface,DFT+vdWsurf indeed provides the most accurate resultsfor the binding distance and energy when compared toall other theoretical methods. Similarly accurate resultsare found for PTCDA adsorbed on other coinage met-als as well as for Xe and benzene on a range of metallicsurfaces.

    Beyond the region of orbital overlap, the vdW in-teraction of an atom a with a solid B is given byEvdW ' −CaB3 /(Z − Z0)3, where Z is the distance fromthe vdW image plane Z0, and the C

    aB3 is the correspond-

    ing vdW interaction coefficient [19–21]. Zaremba andKohn derived this formula starting from quantum me-chanics, obtaining the following expression for the CaB3coefficient [19, 20]

    CaB3 =~

    ∫ ∞0

    dω α(iω)εB(iω)− 1εB(iω) + 1

    , (1)

    where α(iω) is the polarizability of the atom a, andεB(iω) is the dielectric function of the solid B. Equa-tion 1 incorporates screening effects inside the bulk as itis evident from its dependence on the dielectric functionεB(iω).

    A convenient approach to recover the LZK formula [21,22], is to determine the vdW interaction of an atom awith a solid B by a summation of −Cab6 /R6ab pair poten-

    tials between the atom a and the atoms b in the infinitehalf-space of the solid B. In this case, the image-planeposition is given as Z0 = d/2, where d is the interlayerdistance, and the CaB3 coefficient is obtained as

    CaB3 = ns

    (π6

    )Cab6,LZK , (2)

    where ns is the number of atoms per unit volume in thebulk.

    Combining Eqs. (1) and (2), the screened Cab6,LZK coef-ficient is derived. It effectively “inherits” the many–bodyscreening effects contained in CaB3 from Eq. (1). The nextstep is to disentangle the heteronuclear Cab6,LZK interac-

    tion coefficient into homonuclear Caa6 and Cbb6,LZK coef-

    ficients. For this purpose, we adopt a Pade-approximantmodel for the atomic polarizability as in the DFT+vdWmethod [17]. The Pade-approximant model allows to per-form analytic frequency integration, leading to a simplecombination rule for Cab6,LZK

    Cab6,LZK =2Caa6 C

    bb6,LZK

    αbLZKαa0

    Caa6 +αa0

    αbLZKCbb6,LZK

    . (3)

    The αa0 and αbLZK correspond to the adsorbate atom and

    the effective “atom-in-a-solid” static polarizabilities, re-spectively. In the context of LZK theory, the frequency-dependent polarizability of the adsorbed atom is equiv-alent to the one of the free atom in the gas phase.Since accurate αa0 and C

    aa6 values are known for the

    free atoms [23], the two parameters to be determinedin Eq. (3) are αbLZK and C

    bb6,LZK . We obtained them by

    solving a system of two equations (Eq. (3)) using any twoatoms from the list: H, C, Ne, Ar and Kr for a given solidB. The dielectric function εB(iω) was calculated in termsof the absorptive part of the dielectric function ε2 at realfrequencies by the Kramers–Kronig relation. The datawas taken from the reflection energy–loss spectroscopy(REELS) experiments by Werner et al. [24]. The LZKdispersion coefficients along with αbLZK for an atom b in-side the bulk corresponding to Cu, Ag, Au, Pd, and Ptmetals, obtained with the above procedure, are shownin Table I. For comparison, the free atomic parametersare also displayed. The screened dispersion coefficientsfor atoms in the bulk are reduced by a factor of 1.5, 2,and 4 for Pd, Au, and Cu, respectively, and 3 for Agand Pt compared to their free atom counterparts. Thisclearly illustrates the sensitive dependence of the dielec-tric screening on the inorganic substrate. We note thatboth LZK parameters are essentially invariant to the na-ture of the adsorbed atom. Thus, they can be consideredas intrinsic properties of the bulk.

    The LZK theory is exact for atom–surface distances be-yond orbital overlap. However, it does not include effectsdue to rapid spatial variations (interface polarization) inthe dielectric function close to the surface. We effec-tively include these contributions through the linkage of

    2

  • TABLE I. Screened C6 coefficients (hartree · bohr6), polar-izability (in bohr3), and vdW radii (in bohr). Free atomicparameters as used in the DFT+vdW method are also shownfor comparison.

    Substrate Cbb6,LZK αbLZK R

    bLZK C

    bb6,free α

    bfree R

    bfree

    Cu 59 10.9 2.40 253 42.0 3.76Ag 122 15.4 2.57 339 50.6 3.82Au 134 15.6 2.91 298 36.5 3.86Pd 102 13.9 3.06 158 23.7 3.66Pt 120 14.5 2.80 347 39.7 3.92

    the LZK theory with the DFT+vdW method. Once theαbLZK , C

    bb6,LZK , and R

    bLZK are computed, the reference

    DFT+vdWsurf polarizability, C6, and vdW radius for thebulk atoms are set to those values. Nevertheless, theseparameters change with the definition of a dimensionlesseffective volume vieff for species i, in terms of the Hirsh-feld partitioning of the electron density [25, 26],

    vieff =

    (∫r3wi(r)n(r)d

    3r∫r3nrefi (r)d

    3r

    ), (4)

    where wi(r) is the Hirshfeld atomic partitioning weightof the species i, r3 is the cube of the distance fromthe nucleus of an atom i, n(r) is the total electrondensity, and nrefi (r) is the reference electron density.For the solid, the reference corresponds to the spher-ical electron density of an atom in the bulk, and fora molecule, it corresponds to the free atom electrondensity. The effective Cii6,eff coefficient is determined

    as Cii6,eff = (vieff)

    2Cii6,ref , the effective polarizability as

    αieff = vieffα

    iref , and the effective vdW radius, which deter-

    mines the onset of the damping function, is calculated asRieff = (α

    ieff/α

    iref)

    1/3Riref [17]. Significant interface polar-ization is observed in the DFT+vdWsurf method whenorganic molecules are adsorbed on metals. For exam-ple, for the PTCDA molecule located at the experimen-tal equilibrium position on Ag(111), the molecular C6,effcoefficient is increased by 9% compared to the gas-phasePTCDA, while for the surface atoms it is decreased by4% compared to the atoms inside the bulk (C6,ref).

    To further benchmark the DFT+vdWsurf approach,we next calculated the binding energy curves for a rigidPTCDA molecule on Cu(111) and Au(111) surfaces.These are shown together with the experimentally deter-mined average adsorption distances in Fig. 2 (Calculationdetails are included in the supplemental material [27].).The data for PTCDA on Ag(111) from Fig. 1 is alsoincluded for the sake of comparison. The first interest-ing observation comes from comparing the PBE+vdWand PBE+vdWsurf binding curves for PTCDA/Ag(111)in Fig. 1. The LZK correction leads to a reduced C6coefficient (see Table I), and yields significantly smallerlong-range vdW energy. However, the vdW radii of

    FIG. 2. Adsorption energy Eads as a function of verticaldistance d for PTCDA on Cu(111), Ag(111) and Au(111)employing the DFT+vdWsurf method. The distance d isevaluated with respect to the position of the unrelaxed top-most metal layer. The experimental adsorption distancesare 2.66±0.05, 2.86±0.05, and 3.34±0.05 Å corresponding toCu [2], Ag [4], and Au [3], respectively. They are indicatedby shaded regions. Calculation details are included in thesupplemental material [27].

    Ag(111) atoms are also reduced, amounting to a largervdW contribution at shorter range. These two effectsmodify the binding curve for PTCDA/Ag(111), leadingto a reduced adsorption energy and a reduced adsorptiondistance compared to PBE+vdW. The same conclusionshold for the other coinage metals. As shown in Fig. 2, theequilibrium distances agree very well with the NIXSWresults and also the adsorption energies reproduce thetrend in the binding strength inferred from experiment:Eads(Cu) > Eads(Ag) > Eads(Au) [4].

    It is well known that the PTCDA molecule is sig-nificantly distorted upon adsorption on Ag(111) andCu(111). Seeking to understand whether PBE+vdWsurf

    can also accurately reproduce the vertical distor-tion, we carried out structural relaxations of thePTCDA/Me(111) systems, where we allowed the atomsin the PTCDA molecules and in the top metal layer torelax. Figure 3 shows the average bonding distancesof the carbon atoms and the carboxylic and anhydrideoxygens for the relaxed geometries and compares themto experimental data where available. The adsorptiondistance of the C–backbone after relaxation agrees re-markably well with the experimental data to better than0.1 Å for Cu(111) and Ag(111), while an underestima-tion of ≈ 0.1 Å is observed for Au(111). The slightlylarger discrepancy observed for Au can be attributed tothe Au(111) surface reconstruction [28, 29]. The posi-tion of the carboxylic oxygen (Oc) for PTCDA/Ag(111)agrees with experiment to better than 0.1 Å. Our calcu-lations reproduce a distortion in the Oc of 0.13 Å with

    3

  • FIG. 3. Comparison between the experimental binding dis-tances [3, 4] (with typical error bars shown in parenthesis)and the calculated distances in the present work for PTCDAon Cu(111) (a), Ag(111) (b) and Au(111) (c). The distancesare measured as an average over the constituent atoms ofthe two molecules in the unit cell (carbon, anhydride oxygen,and carboxylic oxygen) with respect to the unrelaxed topmostmetallic layer. The structure of PTCDA is also displayed in(d). PTCDA/Cu(111) shows a distortion in the carboxylicoxygen, which is in agreement with a low coherent fraction forthe carboxylic oxygen as reported for the NIXSW experimentin [2]. Calculation details are included in the supplementalmaterial [27].

    respect to the C–backbone due to the attraction to thesurface, while the position of the anhydride oxygen (Oa)is underestimated by around 0.15 Å. With respect toPTCDA/Cu(111), we observe that the oxygen atoms areplaced at an averaged position of 2.62 Å i.e., they residebelow the C–backbone, opposite to the findings of Ger-lach et al. [2]. In addition, the Oc is located below theC–backbone, while the Oa is located above it. Our calcu-lations for different monolayer structures show that thevertical position of the oxygen atoms in PTCDA/Cu(111)is sensitive to the lateral placement of the molecules inthe unit cell, and we will present further analysis in forth-coming work [30]. Nonetheless, the overall agreementbetween the geometries derived from NIXSW data andDFT+vdWsurf is within 0.1 Å.

    In order to test the applicability of DFT+vdWsurf

    method to systems other than PTCDA/Me(111), we cal-culated the relaxed geometry for Xe and benzene on avariety of metallic surfaces. Also for these systems, theDFT+vdWsurf approach leads to a good agreement in theadsoption distances and interaction energies comparedto available data as shown in Table II. In the case ofXe/Me(111), we find that the top and fcc hollow adsorp-tion sites are practically degenerate. For Xe/Pt(111),measurements show that, at low coverage, the difussion

    TABLE II. Comparison of equilibrium distance dAds−Sub andadsorption energy between PBE+vdWsurf and experiment forXe/Me(111) [31, 32] and C6H6/Me(111) systems [33–37]. Weare not aware of experimental results for the equilibrium dis-tance of the C6H6/Me(111) systems. For the Xe/Me(111)systems, we report the dAds−Sub distance for the top adsorp-tion site. The distances are referenced to the relaxed topmostmetallic layer. Calculation details are included in the supple-mental material [27].

    SubstratedAds−Sub [Å] Eads [eV]

    This work Exp. This work Exp.Xe/Cu 3.46 3.60 ± 0.08 0.24 0.17–0.20Xe/Ag 3.57 3.60 ± 0.05 0.21 0.20–0.23Xe/Pt 3.46 3.40 ± 0.10 0.20 0.26–0.28Xe/Pd 3.10 3.07 ± 0.06 0.29 0.31–0.33

    C6H6/Cu 2.68 – 0.91 0.68–0.81C6H6/Ag 2.93 – 0.81 0.66–0.80C6H6/Pt 2.08 – 2.18 1.84–2.25C6H6/Pd 2.10 – 2.11 –

    barrier for lateral movement of the Xe atoms on thesurface is less than 10 meV [38]. We report the Xe-surface distance for the top site because it is experimen-tally determined as the most stable one. For Xe/Pt(111),the PBE+vdWsurf method also yields an accurate vibra-tional energy of 3.4 meV compared to the experimentalvalue of 3.7 meV [39]. Detailed analysis of this problemwill be presented in a forthcoming publication [40].

    In summary, we have combined the DFT+vdWmethod with the Lifshitz-Zaremba-Kohn theory, lead-ing to a promising method (DFT+vdWsurf) that can ac-curately describe the structure and energetics of HIOS.For non-metals, the DFT+vdWsurf theory is equivalentto using screened reference parameters for “atom-in-a-solid” as recently presented by some of us [41]. TheDFT+vdWsurf method has the same cost as the under-lying DFT calculation, and does not depend on the na-ture of the surface, being equally applicable to insulators,semiconductors, and metals.

    We are grateful for support from the FP7 Marie CurieActions of the European Commission, via the InitialTraining Network SMALL (MCITN-238804). A.T. issupported by a grant from the European Research Coun-cil (ERC Starting Grant VDW-CMAT). The authors ac-knowledge Dr. Lorenz Romaner and Dr. Oliver Hofmannfor helpful discussions.

    [1] L. Kronik and N. Koch, MRS Bull. 35, 417 (2010).[2] A. Gerlach et al., Phys. Rev. B 75, 045401 (2007).[3] S. K. M. Henze et al., Surf. Sci. 601, 1566 (2007).[4] A. Hauschild et al., Phys. Rev. B 81, 125432 (2010).[5] N. Atodiresei, V. Caciuc, P. Lazić, and S. Blügel, Phys.

    Rev. Lett. 102, 136809 (2009).[6] G. Mercurio et al., Phys. Rev. Lett. 104, 036102 (2010).

    4

  • [7] D. Stradi et al., Phys. Rev. Lett. 106, 186102 (2011).[8] T. Olsen, J. Yan, J. J. Mortensen, and K. S. Thygesen,

    Phys. Rev. Lett. 107, 156401 (2011).[9] E. McNellis, Ph.D. thesis, Fritz–Haber–Institut der MPG

    (2010).[10] A. Tkatchenko et al., MRS Bull. 35, 435 (2010).[11] L. Romaner et al., New J. Phys. 11, 053010 (2009).[12] U. Stahl et al., Surf. Sci. 414, 423 (1998).[13] J. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett.

    77, 3865 (1996).[14] M. Rohlfing, R. Temirov, and F. Tautz, Phys. Rev. B 76,

    115421 (2007).[15] M. Dion et al., Phys. Rev. Lett. 92, 246401 (2004).[16] S. Grimme, J. Comput. Chem. 27, 1787 (2006).[17] A. Tkatchenko and M. Scheffler, Phys. Rev. Lett. 102,

    073005 (2009).[18] M. Rohlfing and T. Bredow, Phys. Rev. Lett. 101, 266106

    (2008).[19] E. M. Lifshitz, Soviet Physics JETP 2, 73 (1956).[20] E. Zaremba and W. Kohn, Phys. Rev. B 13, 2270 (1976).[21] L. W. Bruch, M. W. Cole, and E. Zaremba, Physical

    Adsorption: Forces and Phenomena (Oxford UniversityPress, 1997).

    [22] S. H. Patil, K. T. Tang, and J. P. Toennies, J. Chem.Phys. 116, 8118 (2002).

    [23] X. Chu and A. Dalgarno, J. Chem. Phys. 121, 4083(2004).

    [24] W. Werner, K. Glantschnig, and C. Ambrosch-Draxl, J.

    Phys. Chem. Ref. Data 38, 1013 (2009).[25] F. L. Hirshfeld, Theor. Chim. Acta 44, 129 (1977).[26] E. R. Johnson and A. D. Becke, J. Chem. Phys. 123,

    024101 (2005).[27] Details of the DFT calculations and the employed struc-

    tures for HIOS can be found in the supplemental mate-rial.

    [28] S. Mannsfeld et al., Org. Electron. 2, 121 (2001).[29] L. Killian, E. Umbach, and M. Sokolowski, Surf. Sci. 600,

    2633 (2006).[30] L. Romaner et al., J. Phys.: Condens. Matter., to be

    submitted.[31] R. D. Diehl et al., J. Phys.: Condens. Matter 16, S2839

    (2004).[32] G. Vidali et al., Surf. Sci. Rep. 12, 133 (1991).[33] S. Lukas et al., J. Chem. Phys. 114, 10123 (2001).[34] M. Xi et al., J. Chem. Phys. 101, 9122 (1994).[35] R. Caputo et al., J. Phys. Chem. A 111, 12778 (2007).[36] J. Jalkanen et al., J. Phys. Chem. B 110, 5595 (2006).[37] J. Gottfried et al., J. Phys. Chem. B 110, 17539 (2006).[38] J. Ellis, A. P. Graham, and J. P. Toennies, Phys. Rev.

    Lett. 82, 5072 (1999).[39] L. W. Bruch, A. P. Graham, and J. P. Toennies, Mol.

    Phys. 95, 579 (1999).[40] W. Liu et al., to be submitted.[41] G. X. Zhang et al., Phys. Rev. Lett. 107, 245501 (2011).

    5