+ All Categories
Home > Documents > The role of the tip symmetry on the STM topography of ...jcc2161/documents/Siegert_et_al...and...

The role of the tip symmetry on the STM topography of ...jcc2161/documents/Siegert_et_al...and...

Date post: 24-Jan-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
8
Phys. Status Solidi B 250, No. 11, 2444–2451 (2013) / DOI 10.1002/pssb.201350002 p s s basic solid state physics b status solidi www.pss-b.com physica Part of Special Issue on Quantum Transport at the Molecular Scale The role of the tip symmetry on the STM topography of -conjugated molecules Benjamin Siegert * , Andrea Donarini, and Milena Grifoni Institute of Theoretical Physics, University of Regensburg, 93040 Regensburg, Germany Received 4 June 2013, revised 23 September 2013, accepted 30 September 2013 Published online 30 October 2013 Keywords electronic transport, molecular electronics, reduced density matrix, scanning tunneling microscopy Corresponding author: e-mail [email protected], Phone: +49 941 943 2044, Fax: +49 941 943 2038 We present an STM theory based on the reduced density matrix (RDM) formalism which is able to describe transport proper- ties of an STM junction for -conjugated molecules on thin insulating films. It combines a very popular derivation of STM tunneling matrix elements [1], based on Bardeen’s tunneling formalism [2], with a generalized master equation approach for interacting molecular systems. We show that this method allows the efficient implementation of different tip symmetries in STM simulations. With the example of hydrogen phthalocyanine (H 2 Pc), we study the influence of s- and p-wave tip sym- metries on the constant-height current maps of -conjugated molecules. Constant-height STM images evaluated at the cationic reso- nance of H 2 Pc. Left image computed by using an s-wave tip and right image for a linear combination of p x and p y states. © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1 Introduction Since its invention in 1982 by Binning and Rohrer [3], scanning tunneling microscopy has been gathering enormous attention in- and outside the physics community. There already exist many theories describing STM junctions [4–11], and a comprehensive review can be found in Ref. [12]. One of the most famous theories of STM was initially proposed by Tersoff and Hamann [13–15] based on the theory of tunneling of Bardeen [2]: They modeled the tip as a spherical potential well and showed that the tunneling current is proportional to the local density of states of the sam- ple. Although not capable of explaining atomic resolution in STM experiments, this theory laid the foundations for the understanding of STM images. A more sophisticated descrip- tion of the tip-sample tunneling was proposed by Chen [1] which took into account higher angular momentum states and therefore different tip symmetries. A major benefit of his proposal is of rather technical nature; he concluded in his “derivative rule” that tunneling matrix elements involving a given tip symmetry are proportional to corresponding spatial derivatives of the sample wavefunction evaluated at the apex of the tip. He further stated that higher corrugation amplitudes and thus atomic resolution can be ascribed to d z 2 -tip symme- tries [16]. In recent years, STM experiments with molecules on thin insulating films [17–19] are enjoying increasing inter- est. The insulating layers, only few ˚ angstr¨ om thin and grown on the conductive substrate, are effectively decoupling the molecular electrons from the electrons in the substrate. Thus, there is essentially no hybridization of substrate and sample wavefunction, which allows to study the electronic properties of the pristine molecule at low temperatures. In this work we combine the reduced density matrix (RDM) formalism [20] with the STM tunneling theory of Chen [1]. The RDM formalism is a powerful tool to inves- tigate the transport properties of interacting nanosystems weakly coupled to leads. It is applicable also to molecular quantum dots [21–23] where it can be used to study many- body quantum interference in three-terminal [24] and STM setups [25, 26]. Its combination with Chen’s theory allows © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
Transcript
  • Phys. Status Solidi B 250, No. 11, 2444–2451 (2013) / DOI 10.1002/pssb.201350002 p s sbasic solid state physics

    b

    statu

    s

    soli

    di

    www.pss-b.comph

    ysi

    ca

    Part of Special Issue onQuantum Transport at the Molecular Scale

    The role of the tip symmetry on theSTM topography of �-conjugatedmoleculesBenjamin Siegert*, Andrea Donarini, and Milena Grifoni

    Institute of Theoretical Physics, University of Regensburg, 93040 Regensburg, Germany

    Received 4 June 2013, revised 23 September 2013, accepted 30 September 2013Published online 30 October 2013

    Keywords electronic transport, molecular electronics, reduced density matrix, scanning tunneling microscopy

    ∗ Corresponding author: e-mail [email protected], Phone: +49 941 943 2044, Fax: +49 941 943 2038

    We present an STM theory based on the reduced density matrix(RDM) formalism which is able to describe transport proper-ties of an STM junction for �-conjugated molecules on thininsulating films. It combines a very popular derivation of STMtunneling matrix elements [1], based on Bardeen’s tunnelingformalism [2], with a generalized master equation approach forinteracting molecular systems. We show that this method allowsthe efficient implementation of different tip symmetries in STMsimulations. With the example of hydrogen phthalocyanine(H2Pc), we study the influence of s- and p-wave tip sym-metries on the constant-height current maps of �-conjugatedmolecules. Constant-height STM images evaluated at the cationic reso-

    nance of H2Pc. Left image computed by using an s-wave tipand right image for a linear combination of px and py states.

    © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

    1 Introduction Since its invention in 1982 by Binningand Rohrer [3], scanning tunneling microscopy has beengathering enormous attention in- and outside the physicscommunity. There already exist many theories describingSTM junctions [4–11], and a comprehensive review can befound in Ref. [12]. One of the most famous theories of STMwas initially proposed by Tersoff and Hamann [13–15] basedon the theory of tunneling of Bardeen [2]: They modeled thetip as a spherical potential well and showed that the tunnelingcurrent is proportional to the local density of states of the sam-ple. Although not capable of explaining atomic resolution inSTM experiments, this theory laid the foundations for theunderstanding of STM images. A more sophisticated descrip-tion of the tip-sample tunneling was proposed by Chen [1]which took into account higher angular momentum statesand therefore different tip symmetries. A major benefit ofhis proposal is of rather technical nature; he concluded in his“derivative rule” that tunneling matrix elements involving agiven tip symmetry are proportional to corresponding spatial

    derivatives of the sample wavefunction evaluated at the apexof the tip. He further stated that higher corrugation amplitudesand thus atomic resolution can be ascribed to dz2 -tip symme-tries [16]. In recent years, STM experiments with moleculeson thin insulating films [17–19] are enjoying increasing inter-est. The insulating layers, only few ångström thin and grownon the conductive substrate, are effectively decoupling themolecular electrons from the electrons in the substrate. Thus,there is essentially no hybridization of substrate and samplewavefunction, which allows to study the electronic propertiesof the pristine molecule at low temperatures.

    In this work we combine the reduced density matrix(RDM) formalism [20] with the STM tunneling theory ofChen [1]. The RDM formalism is a powerful tool to inves-tigate the transport properties of interacting nanosystemsweakly coupled to leads. It is applicable also to molecularquantum dots [21–23] where it can be used to study many-body quantum interference in three-terminal [24] and STMsetups [25, 26]. Its combination with Chen’s theory allows

    © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

    JulianHighlight

    JulianHighlight

    JulianHighlight

    JulianHighlight

    JulianHighlight

  • Original

    Paper

    Phys. Status Solidi B 250, No. 11 (2013) 2445

    Figure 1 (Left) Artistic view of the STM setup. (Right) Sketch ofthe potential profile of the STM setup along the direction perpen-dicular to the molecular plane.

    us to consider STM junctions containing weakly coupledmolecules and tips with different symmetries.

    We especially focus how s-wave, p-wave, and mixed sp-wave tip states affect the structural pattern of nodal planesand orbital lobes (as shown in the figure in the abstract) inSTM images.

    The outline of this paper is as follows: in Section 2 weintroduce our transport theory for �-conjugated moleculeson thin insulating films in STM junctions together with theHamiltonian of the system. We derive the tunneling matrixelements entering the equations of motion for the reduceddensity matrix of the system and show how to compute thetunneling current. In Secion 3, the results of our calculationsare provided and discussed. Finally, in Section 4 we presentour conclusions.

    2 Model Our system consists of a �-conjugatedmolecule on a thin insulating film coupled to an STM tipand can be described by the Hamiltonian

    H = Hmol + HS + HT + Htun. (1)

    The presence of the insulating layer allows to neglecthybridization of the molecular wavefunction with the wave-function of the substrate and thus to work in the weakcoupling sequential tunneling regime. The Hamiltonian ofthe molecule Hmol = H0 + Vee is given by

    Hmol =∑

    ν

    �νd†νd

    ν+

    ∑〈ν,ν′〉

    bνν′d†νd

    ν′ + Vee, (2)

    where the multi-indices ν, ν′ include the atomic site rα =(xα, yα, zα), the orbital index lm ∈ {1s, 2s, 2px, 2py, 2pz} of ahydrogen-like atomic orbital and the spin index. We set thez-coordinate of the molecule to d, where z = 0 defines themetal–insulator interface (see Fig. 1).

    The operator d†ν

    creates an electron in the molecule withthe set of quantum numbers ν with the onsite energy �ν. Thevalues for the different onsite energies were taken from Ref.[27]. We consider the set of all 2s (1s for hydrogen),

    2px and 2py orbitals as the �-system, and consequentlythe set of 2pz orbitals as the �-system. The hopping parame-ters bνν′ are calculated using the Slater–Koster LCAO method[28] with the overlap parameters taken from Refs. [27, 29]and the geometrical parameters taken from Refs. [30–32]. Forthe electron–electron interaction Vee we are using a constant-interaction term [33]:

    Vee = U2

    (N − N0)2 , (3)

    where U quantifies the strength of the Coulomb interaction.N = ∑

    νd†

    νd

    νand N0 is the number of electrons occupying

    the neutral molecule. Under consideration that the grand-canonical energy of the molecule,

    EG = 〈HG〉 = 〈H − μGN〉, (4)

    with μG being the chemical potential of the leads in equilib-rium, has to be minimal in its neutral state, we adjusted theparameter U and renormalized the single-particle eigenener-gies by a constant value. This criterion only fixes a rangeof possible U and single particle shifts. The final valueshave been obtained by fitting our transport calculations toexperimental results [34, 35]. Image charge effects are knownto produce such renormalizations of onsite energies andCoulomb interactions [36]. A microscopic evaluation of thiseffect lies outside the scope of this paper and will be consid-ered for later work.

    The substrate and the tip are assumed to act as reservoirsof noninteracting electrons, Hη =

    ∑kσ

    �η

    kc†ηkσcηkσ , where η =

    S, T denotes substrate or tip. The tunneling Hamiltonian Htundescribes tunneling between the molecule and substrate andbetween the tip and the molecule:

    Htun =∑ηkjσ

    kjc†ηkσdjσ + h.c., (5)

    where η again denotes substrate or tip, k = (kx, ky, kz) is thewavevector of an electron in lead η, σ is the spin of the elec-tron and the index j denotes a molecular orbital (MO) ofthe �-system. The tunneling amplitudes tηkj depend on thewavevector k and the MO index j.

    The single-particle MOs of the system are obtained bynumerical diagonalization of H0. Due to the planar geometryof H2Pc [30, 37, 38] the �-system is decoupled from the �-system and thus not influenced by the presence of the centralhydrogen atoms. The reduced symmetry can be recoveredalso in the �-system if we introduce a spatial modulationof the on-site energies ��

    αassociated to the electron density

    of the filled �-orbitals. For this purpose, we introduce theMulliken charge [39] relative to all occupied �-orbitals at

    www.pss-b.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

    JulianHighlight

  • ph

    ysic

    a ssp status solidi b

    2446 B. Siegert et al.: Tip symmetry and STM topography

    each atomic site:

    QMα

    = 2e∑j, lm

    |c(α,lm)j|2, (6)

    where e is the absolute value of the electron charge and c(α,lm)jis the LCAO coefficient of the orbital lm at the atomic site αin the MO j. The Mulliken charge reflects the non-uniformdelocalization of the MOs over all atomic valence orbitalsin the �-system. If we exclude the electron occupying the pzorbital, an isolated carbon (nitrogen) atom has 3(4) �-likevalence electrons. We denote the associated charge Q0

    αthe

    expected Mulliken charge. From the difference between theexpected and the actual Mulliken charge one can deduce alocal variation of the onsite energy for the valence electronin the pz orbital at site α:

    Δ�α

    = 2QMα

    − Q0α

    Qeffα

    �r��

    α, (7)

    where �r is the dielectric constant of the underlying insulator.This variation is derived as follows: We assume ��

    αto be the

    energy of a second shell orbital of a hydrogen-like atom withcharge Qeff

    α,

    ��α

    =(

    Qeffα

    e

    )2EH1 (e)

    22= EH2 (Qeffα ), (8)

    where EH1 (e) = −13.6 eV. The non-uniform delocalization ofthe �-orbitals introduces a variation of the expected Mullikencharge δα =

    (QM

    α− Q0

    α

    )and thus a screening of Qeff

    α:

    EH2

    (Qeff

    α+ δα

    �r

    )− EH2 (Qeffα ) ≈ 2

    EH1 (e)

    22e2�rQeff

    αδα (9)

    = 2QMα

    − Q0α

    Qeffα

    �r��

    α. (10)

    Note that the variation of the charge has been screened bythe underlying insulator. Using these energy corrections, wecan establish an effective Hamiltonian for the �-system:

    H�eff =∑

    α

    (��

    α+ Δ�

    α

    )d†

    αd

    α. (11)

    Diagonalizing this effective Hamiltonian yields the MOs ofthe �-system:

    〈r|jσ〉 =∑

    α

    cαj〈r|ασ〉 =∑

    α

    cαjpz(r − rα) uσ, (12)

    where �j is the energy of the MO j and 〈r|ασ〉 denotes apz-orbital with spinor component uσ centered at the atomicsite rα,

    pz(r − rα) = Nα(z − zα) e−Qeffα +

    δα�r

    2a0|r−rα|, (13)

    with normalization Nα and a0 being Bohr’s radius. The tun-neling region between the substrate and the molecule ismodelled as a one-dimensional potential barrier. Due to thez-localization of the molecular orbitals one can prove that forthe substrate tunneling amplitudes it holds [25]:

    tSkj = �j〈Skσ|jσ〉 = �j∑

    α

    cαj〈Skσ|ασ〉, (14)

    where |Skσ〉 is an eigenstate of the substrate Hamiltonian.The latter is modelled as a quantum well, only confined inz-direction, see Fig. 1. A product ansatz for the substratewavefunction yields plane waves in the xy-plane and the solu-tion of a one-dimensional finite potential well in z-direction:

    〈r|Skσ〉 = 1√S

    eikxx+ikyy Ψkz (z) uσ (15)

    with uσ being its spinor part and S the surface area of thesubstrate. The wavefunction in z-direction reads

    Ψkz (z) = Ω

    ⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

    − sin(kzz0)κ2 + k2

    z

    2κkzeκ(z−z0), z < z0,

    cos(kzz) − κkz

    sin(kzz), z0 ≤ z ≤ 0,e−κz, z > 0.

    (16)

    For the wavevectors, we have k2x/y/z

    = 2m�2

    �x/y/z and κ2 =2m�2

    (�SF + φS0 − �z) for the decay constant of the evanescentparts of the piecewise-defined wavefunction.

    Consequently, the matrix element between substratewavefunction and a pz orbital in Eq. (14),

    〈Skσ|ασ〉 =∫

    d3re−ikxx−ikyy√

    SΨkz (z)pz(r − rα), (17)

    can be calculated as

    〈Skσ|ασ〉 = Vα(kz, k‖, κ) e−ik‖·rα , (18)

    where the scalar product k‖ · rα = kxxα + kyyα results fromthe shifting of the integrand in Eq. (17) by +rα and thefunction Vα(kz, k‖, κ), equivalent to OS(k) in Ref. [25], con-denses the results of the integration. Ultimately, the tunnelingamplitudes for the substrate read:

    tSkj = �j∑

    α

    cαjVα(kz, k‖, κ) e−ik‖·rα . (19)

    For an even more detailed description of the derivation of theeigenstates of the substrate Hamiltonian and the calculationof the tunneling amplitudes for the substrate, we refer to Refs.[25, 26], respectively.

    For the derivation of the tip tunneling amplitudes, wefollow in this paper the derivation proposed in Ref. [1] rather

    © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com

  • Original

    Paper

    Phys. Status Solidi B 250, No. 11 (2013) 2447

    Table 1 Tunneling matrix elements for the tip for l = 0 (s state), l = 1 (p states), and l = 2 (d states).

    tip state (lc) tunneling amplitude Mlckj tip state (lc) tunneling amplitude Mlckj

    s2π�2Cs

    mκψ(rtip) dz2

    2π�2Cz2

    mκ3

    (∂2

    ∂z2tip− 1

    3κ2

    )ψ(rtip)

    pz2π�2Cz

    mκ2

    ∂ztipψ(rtip) dxz

    2π�2Cxzmκ3

    ∂2

    ∂xtip∂ztipψ(rtip)

    px2π�2Cx

    mκ2

    ∂xtipψ(rtip) dyz

    2π�2Cyzmκ3

    ∂2

    ∂ytip∂ztipψ(rtip)

    py2π�2Cy

    mκ2

    ∂ytipψ(rtip) dxy

    2π�2Cxymκ3

    ∂2

    ∂xtip∂ytipψ(rtip)

    dx2−y22π�2Cx2−y2

    mκ3

    (∂2

    ∂x2tip− ∂

    2

    ∂y2tip

    )ψ(rtip)

    than the one in Ref. [26]. The advantage of this approach isthat it is independent of the explicit form of the wavefunc-tion inside the tip; it rather models the exponentially decayingpart of the tip wavefunction and allows to consider differentsymmetries of tip wavefunctions, which are reflected in theform of their vacuum tails. It is based on Bardeen’s tunnel-ing theory [2], which states that a tunneling matrix elementbetween two wavefunctions χ and ψ living in two differentspatial regions can be represented by the integral

    M(χ, ψ) = − �2

    2m

    ∫Σ

    dS (χ∗∇ψ − ψ∇χ∗) , (20)

    where Σ is the separation surface of the two regions. Fol-lowing Ref. [1], we write the wavefunction of the tip χk(r)in a separation ansatz as a linear combination of real-valuedspherical harmonics (cubic harmonics) Ylc(θ, φ) and spheri-cal modified Bessel functions of the third kind kl(κρ) [40]:

    χk(r) =∑

    l,cClcχ

    lck (r), (21)

    χlck (r) := Nlckl(κρ)Ylc(θ, φ), (22)

    where κ =√

    2m�2

    (�TF + φT0 − �k), ρ = |r − rtip|. The Clc aredimensionless mixing coefficients and the Nlc are ensur-ing that the resulting tunneling matrix element has the rightdimension of energy. They are calculated using a cutoff of κ−1

    for the lower integration limit, since the kl(x) are not regularat x = 0 for l > 1. In the present work we are consideringan idealized choice of the coefficients Clc, restricting to pures, pure p, or equally mixed s and p states. In order to betterdescribe realistic tips the Clc can be determined from firstprinciple calculations, following, e.g., the lines of Ref. [41].By using the recurrence relations of the spherical modifiedBessel functions and relating k0(κρ) to the Green’s functionof the spherical Helmholtz equation, Ref. [1] arrives at his“derivative rule,” stating that a tunneling amplitude involvingtip states with symmetries different from s-wave symmetriescan be related to the respective spatial derivatives of the sam-

    ple wavefunction, evaluated at the position of the tip. Finally,by inserting Eq. (22) into Eq. (20) and following the pro-cedure in Ref. [1], the tunneling amplitudes for the tip areobtained as:

    tTkj =∑

    l,c

    ClcMlc

    kj =∑

    l,c

    ClcM(χlc

    k , ψj). (23)

    Tunneling amplitudesMlckj for s states (l = 0), p states (l = 1),and d states (l = 2) are listed in Table 1.

    The dynamics of our system is calculated using thereduced density matrix approach [20]. We give in this workonly a short overview over the derivation of the equa-tion of motion for the reduced density matrix (RDM). Amore detailed discussion can be found for example in Refs.[23–25]. The starting point is the density matrix of the com-plete system in the interaction picture. Its time evolution isdescribed by the Liouville–von Neumann equation,

    i�dρI(t)

    dt= [HItun(t), ρI(t)] (24)

    with the tunneling Hamiltonian acting as a perturbation. Eq.(24) has the formal solution

    ρI(t) = ρI(t0) − i�

    ∫ tt0

    dt1 [HI

    tun(t1), ρI(t1)]. (25)

    By reinserting Eq. (25) into Eq. (24) and tracing out thedegrees of freedom of the substrate and the tip, we obtainthe generalized master equation:

    ρ̇Ired(t)

    = − i�

    trS,T{

    [HItun(t), ρI(t0)]

    }

    +(

    i

    )2 ∫ tt0

    dt1 trS,T{ [

    HItun(t), [HI

    tun(t1), ρI(t1)]

    ] },

    (26)

    www.pss-b.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

    JulianHighlight

    JulianHighlight

    JulianHighlight

    JulianHighlight

  • ph

    ysic

    a ssp status solidi b

    2448 B. Siegert et al.: Tip symmetry and STM topography

    which describes the time evolution of the RDM up to secondorder in the tunneling Hamiltonian. In the secular approxima-tion, after retransforming the reduced density matrix to theSchrödinger picture we project it onto subblocks of particlenumber N and energy E as

    ρNE = PNE ρSred PNE, (27)

    where we used the projection operator

    PNE =∑

    j

    |NEj〉〈NEj|. (28)

    Equation (26) finally becomes:

    ρ̇NE = −∑

    η

    ∑i, j

    ∑E′

    {

    Γ η+ij

    (E′ − E)PNE di PN+1E′ d†j ρNE

    + Γ η−ij

    (E − E′)PNE d†i PN−1E′ dj ρNE

    − Γ η−ij

    (E′ − E)PNE di ρN+1E′ d†j NE

    − Γ η+ij

    (E − E′)PNE d†i ρN−1E′ dj NE

    }. (29)

    Furthermore, we restrict our transport calculations to themany-body ground states of the molecule. The rate Γ η±ij (�E)in Eq. (29) is given by

    Γ η±ij

    (�E) = 2π�

    ∑k

    (tη

    ki

    )∗tη

    kjf±η

    (�E)δ(�ηk − �E), (30)

    where f +η

    (�k) is the Fermi distribution of an electron inlead η = (S, T ) with chemical potentials μT = −φT0 + ceVb,μS = −φS0 − (1 − c)eVb and f −η (�k) = 1 − f +η (�k). Notethat in STM setups there is an asymmetric bias drop c acrossthe junction.

    Its value can be estimated as follows: Let us con-sider an empty capacitor (substrate–tip) with capacitanceC0. After replacing half of the vacuum inside the capaci-tor by a dielectric with relative permittivity εr (substrate +insulatinglayer − tip), the effective capacitance of this sys-tem then reads Ceff = εr1+εr C0, yielding the parameter for thebias drop as c = Ceff

    C0= εr

    1+εr . This estimation relies on theassumption of equal tip–molecule and molecule–substratedistances. They are comparable in typical STM setups withinsulating films. Taking the relative permittivity of NaClεNaClr = 5.92 [42] yields c = 0.86. At present we neglect apossible dependence of the bias drop on the tip and substrateworkfunction. We believe that this effect would be capturedby a self-consistent solution of the Poisson equations for thejunction that goes beyond the scope of this publication.

    Equation (29) can be put in a shorter form,

    ρ̇red = Lρred. (31)

    The operator L is called the Liouvillian of the system. Wecalculate the stationary solution ρstatred which is given by thenullspace of L. Ultimately, the stationary current throughcontact η is calculated as

    〈Iη〉 = trsys{

    Iηρstatred

    }, (32)

    where the current operator Iη can be obtained from the time-derivative of the average charge on the system and is definedas [25, 26]:

    Iη = e∑i, j

    ∑N

    ∑E, E′

    PNE{

    diPN+1E′ d†j Γ η+ij

    (E′ − E)

    − d†iPN−1E′ dj Γ η−ij

    (E − E′) }PNE. (33)

    3 Results In all calculations we used a substrate–molecule distance of d = 7 Å and the workfunctions of thesubstrate and the tip were chosen to be equal: φT0 = φS0 =−μG, see also Fig. 1. In Fig. 2, we show the I–V characteris-tics and a stability diagram of H2Pc. From Fig. 2a, we are ableto inhere the state of the system according to the values of theworkfunction and the bias voltages. In analogy with stabilitydiagrams associated to single electron transistors, what wesee in the central part of this figure is a Coulomb diamond.Varying the workfunction acts as ramping an effective gatevoltage. Transitions occur if the chemical potential of thesubstrate μS = μG − (1 − c)eVb or the chemical potential ofthe tip μT = μG + ceVb matches the chemical potential of themolecule μN = EN − EN−1. At zero bias, going left throughthe diamonds increases the particle number in the system,while going right decreases it. Due to the asymmetric biasdrop c, the Coulomb diamond is tilted. Its steeper bordersmark transitions where the chemical potential of the substratealigns with the chemical potential of the system, while its flat-ter borders indicate a resonance with the chemical potentialof the tip.

    The bias trace in Fig. 2b, taken at μG = −2.9 eV,shows two resonances at Vb = 0.13 V and Vb = −0.8 V. Thecorresponding constant-height STM images in both casesresemble the LUMO orbital of H2Pc. The reason for thisis the transition from an N to an N + 1-particle state (seeFig. 2a), which is mediated by the LUMO single particleMO. The I–V characteristic in Fig. 2c corresponds instead toμG = −4 eV. The constant-height STM images correspond-ing to the two resonances now resemble different orbitals.The current map at positive bias still reflects the LUMO asthe system undergoes an N to N + 1 transition, while at neg-ative bias it reflects the shape of the HOMO consistently withthe corresponding N − 1 to N transition. This regime, wherethere is an anionic resonance at positive bias and a cationicresonance at negative bias is usually the most common inSTM experiments with molecules on thin insulating films[17, 18]. In the following we will discuss the influence ofdifferent tip symmetries on the STM topography of H2Pc,

    © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com

  • Original

    Paper

    Phys. Status Solidi B 250, No. 11 (2013) 2449

    Figure 2 Stability diagram (a) and I–V characteristics (b, c) for H2Pc on a thin insulating film of thickness d = 7 Å and a tip height of4.3 Å. In the central region in panel a transport is blocked due to Coulomb blockade and the molecule is neutral, with N = 42 electronspopulating the �-system. The current traces shown in panels b and c correspond to the two dashed lines in panel a. The small insets showconstant-height STM images computed at the bias voltage of the respective resonance, indicated by the arrows.

    with the workfunctions of the tip and the substrate chosen tobe φT0 = φS0 = −μG = 4 eV.

    For the s-wave tip symmetry the recorded constant-height STM images, Fig. 3a and d, are showing the

    characteristic patterns of nodal planes and orbital lobes ofthe corresponding MOs.

    A striking difference to the STM images with the s-wavetip can be seen in Fig. 3b and e, where constant-height STM

    Figure 3 Constant-height STM images of H2Pc for different tip configurations at a tip-sample distance of 5 Å and a thickness of theinsulating layer d = 7 Å. Panels a–c show current maps recorded at the cationic resonance at Vb = −1.1 V, while panels d–f show currentmaps recorded at the anionic resonance at Vb = 1.4 V. The white bar indicates a length of 5 Å.

    www.pss-b.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

  • ph

    ysic

    a ssp status solidi b

    2450 B. Siegert et al.: Tip symmetry and STM topography

    images are shown for a tip with px and py symmetry. Thenodal plane patterns of the MOs disappeared; instead thecurrent now has maxima and minima at the positions ofthe nodal planes and of the orbital lobes, respectively. Byconsidering the tunneling matrix elements as the derivativesof the molecular wavefunction in the x- and y-directions,this result is not surprising. However, as put forward inRef. [19], there exists a nice empirical explanation of thesefindings, namely that if the tip is directly placed above apz-orbital of the molecule, the contributions of the p-orbitalin the tip are cancelling each other due to their phases,differing in their signs. By placing the tip above a nodalplane, that is between two pz orbitals which differ in theirsigns, there is no such cancellation of the contributions ofthe p-orbital in the tip and tunneling is occuring.

    Finally, in Fig. 3c and f we show the results of a simu-lated mixed s–p wave tip, which we modeled as an equallyweighted linear combination of s, px, and py orbitals. Theyshow the non-uniform competition of the contributions ofthe s and the p orbitals in the tip. In the inner parts of themolecule there are still maxima in the current occuring atthe positions of the nodal planes. In the outer parts, how-ever, the contribution of the s-wave tip state outweighs thep-wave contribution, and there are maxima at orbital lobes.Additionally, in contrast to Fig. 3b and e, the overall imageappears more smeared out, what also can be attributed to thes-wave part of the tip state.

    4 Conclusions In this work we have presented a semi-quantitative transport theory for STM-setups involving �-conjugated molecules on thin insulating films. To this endwe derived tunneling matrix elements that account for theconditions under which tunneling between the molecule andthe contacts takes place. For the substrate we started froma microscopic point of view [26], using the solutions of aSchrödinger equation for a potential well in the z-direction.For the tunneling matrix elements of the tip we used theapproach in Ref. [1], which only considers the shape of thetip wavefunction in the vacuum between tip and sample, butallows for the inclusion of spatially different tip states.

    Finally, we derived the generalized master equation forthe reduced density matrix in secular approximation withthe tunneling Hamiltonian treated in second order. The cur-rent through the system was calculated using the stationarysolution for the reduced density matrix.

    In order to model the investigated molecule we used theSlater–Koster LCAO approach [28] to get its single-particleeigenstates and a constant-interaction model to calculate theenergies of the many-particle ground states of the molecule.

    We presented our numerical calculations for the case ofhydrogen phthalocyanine. As in a recent experimental work[19], which addressed the same question, we confirm thatdifferent tip symmetries have strong influence on the topog-raphy of STM images.

    All in all we showed that our formalism is able tomirror experimental findings quite well, despite the limi-tations induced by our applied approximations. A natural

    improvement of our model is a better treatment of screeningand image charge effects induced by the contacts, as theycould be responsible for negative differential conductanceand the breaking of molecular symmetries [36]. Also therole of vibronic excitations would be an interesting topic toaddress. For example, it has been experimentally shown inRef. [18] for the case of a naphthalocyanine that vibrons canbe crucial in the current-induced switching of the positionsof the central hydrogen atoms of the molecule.

    Acknowledgements Sincere thanks are due to Jascha Reppfor fruitful discussions. We also acknowledge financial support bythe DFG within the research programs SPP 1243 and SFB 689.

    References

    [1] C. J. Chen, Phys. Rev. B 42, 8841 (1990).[2] J. Bardeen, Phys. Rev. Lett. 6, 57 (1961).[3] G. Binning and H. Rohrer, Phys. Rev. Lett. 49, 57 (1982).[4] N. García, C. Ocal, and F. Flores, Phys. Rev. Lett. 50, 2002

    (1983).[5] N. García and F. Flores, Physica B + C 127, 137 (1984).[6] A. Baratoff, Physica B + C 127, 143 (1984).[7] E. Stoll, A. Baratoff, A. Selloni, and P. Carnevali, J. Phys. C

    17, 3073 (1984).[8] C. Noguera, J. Microsc. 152, 3 (1988).[9] G. Doyen, E. Koetter, J. P. Vigneron, and M. Scheffler, Appl.

    Phys. A 51, 281 (1990).[10] J. Buker and G. Kirczenow, Phys. Rev. B 72, 205338 (2005).[11] C. Toher, I. Rungger, and S. Sanvito, Phys. Rev. B 79, 205427

    (2009).[12] W. A. Hofer, A. S. Foster, and A. L. Shluger, Rev. Mod. Phys.

    75, 1287 (2003).[13] J. Tersoff and D. R. Hamann, Phys. Rev. Lett. 50, 1998 (1983).[14] J. Tersoff and D. R. Hamann, Phys. Rev. B 31, 805 (1985).[15] P. K. Hansma and J. Tersoff, J. Appl. Phys. 61, R1 (1987).[16] C. J. Chen, Phys. Rev. Lett. 65, 448 (1990).[17] J. Repp and G. Meyer, Phys. Rev. Lett. 94, 026803 (2005).[18] P. Liljeroth, J. Repp, and G. Meyer, Science 317, 1203 (2007).[19] L. Gross, N. Moll, F. Mohn, A. Curioni, G. Meyer, F. Han-ke,

    and M. Persson, Phys. Rev. Lett. 107, 086101 (2011).[20] K. Blum, Density Matrix Theory and Applications (Plenum

    Press, New York, 1996).[21] M. H. Hettler, W. Wenzel, M. R. Wegewijs, and H. Schoeller,

    Phys. Rev. Lett. 90, 076805 (2003).[22] C. Romeike, M. R. Wegewijs, and H. Schoeller, Phys. Rev.

    Lett. 96, 196805 (2006).[23] C. Timm, Phys. Rev. B 77, 195416 (2008).[24] D. Darau, G. Begemann, A. Donarini, and M. Grifoni, Phys.

    Rev. B 79, 235404 (2009).[25] S. Sobczyk, A. Donarini, and M. Grifoni, Phys. Rev. B 85,

    205408 (2012).[26] A. Donarini, B. Siegert, S. Sobczyk, and M. Grifoni, Phys.

    Rev. B 86, 155451 (2012).[27] W. A. Harrison, Elementary Electronic Structure (World Sci-

    entific, Singapore, 1999).[28] J. C. Slater and G. F. Koster, Phys. Rev. 94, 1498 (1954).[29] W. A. Harrison, Phys. Rev. B 24, 5835 (1981).[30] P. N. Day, Z. Wang, and R. Pachter, J. Mol. Struct. (Theochem)

    455, 33 (1998).

    © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-b.com

    JulianHighlight

    JulianHighlight

  • Original

    Paper

    Phys. Status Solidi B 250, No. 11 (2013) 2451

    [31] M. S. Liao and S. Scheiner, J. Chem. Phys. 114, 9780 (2001).[32] D. Dini and M. Hanack, in: The Porphyrin Handbook, Vol. 17,

    Phthalocyanines: Properties and Materials, edited by K. M.Kadish, K. M. Smith, and R. Guilard, (Academic Press, SanDiego, 2003).

    [33] L. P. Kouwenhoven, D. G. Austing, and S. Tarucha, Rep. Prog.Phys. 64, 701 (2001).

    [34] M. Takada and H. Tada, Jpn. J. Appl. Phys. 44, 5332 (2005).[35] D. R. T. Zahn, G. N. Gavrila, and M. Gorgoi, Chem. Phys.

    325, 99 (2006).[36] K. Kaasbjerg and K. Flensberg, Phys. Rev. B 84, 115457

    (2011).

    [37] G. Baffou, A. J. Mayne, G. Comtet, G. Dujardin, L. Stauffer,and P. Sonnet, J. Am. Chem. Soc. 131, 3210 (2009).

    [38] K. Nilson, J. Åhlund, B. Brena, E. Göthelid, J. Schiessling,N. Mårtensson, and C. Puglia, J. Chem. Phys. 127, 114702(2007).

    [39] R. S. Mulliken, J. Chem. Phys. 23, 1833 (1955).[40] M. Abramovitz and I. A. Stegun, Handbook of Mathematical

    Functions (Dover Publications, New York, 1972) p. 443ff,[41] K. Palotás, G. Mándi, and L. Szunyogh, Phys. Rev. B 86,

    235415 (2012).[42] R. P. Lowndes and D. H. Martin, Proc. R. Soc. Lond. A 316,

    351 (1970).

    www.pss-b.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Recommended