Brought to you by:
Paper

Multi-orbital non-crossing approximation from maximally localized Wannier functions: the Kondo signature of copper phthalocyanine on Ag(100)

and

Published 18 August 2011 © 2011 IOP Publishing Ltd
, , Citation Richard Korytár and Nicolás Lorente 2011 J. Phys.: Condens. Matter 23 355009 DOI 10.1088/0953-8984/23/35/355009

0953-8984/23/35/355009

Abstract

We have developed a multi-orbital approach to compute the electronic structure of a quantum impurity using the non-crossing approximation. The calculation starts with a mean-field evaluation of the system's electronic structure using a standard quantum chemistry code; here we use density functional theory (DFT). We transformed the one-electron structure into an impurity Hamiltonian by using maximally localized Wannier functions. Hence, we have developed a method to study the Kondo effect in systems based on an initial one-electron calculation. We have applied our methodology to a copper phthalocyanine molecule chemisorbed on Ag(100), and we have described its spectral function for three different cases where the molecule presents a single spin or two spins with ferro- and anti-ferromagnetic exchange couplings. We find that the use of broken-symmetry mean-field theories such as Kohn–Sham DFT cannot deal with the complexity of the spin of open-shell molecules on metal surfaces and extra modeling is needed.

Export citation and abstract BibTeX RIS

1. Introduction

Since the study of the Kondo features of Ce adatoms on Ag(111) [1] and Co adatoms on Au(111) [2], the scanning tunneling microscope (STM) has become a privileged tool in the study of surface Kondo physics. The STM is a non-intrusive probe that can address adsorbed objects at very low bias with very small currents. Hence, the STM basically explores the equilibrium properties of the adsorbed systems. Besides adatoms, Kondo physics has been revealed in objects of increasing complexity, from single adsorbates [13] to large organic molecules [48], ordered nanostructures [9] and adatom–ligand structures [10, 11].

Several recent reports show that organic molecules display a Kondo state due to a spin in their extended π-orbitals even when they are adsorbed on a metal surface [7, 8, 12, 13]. This is somewhat of a surprise for two reasons: (i) some of these molecules are magnetic in the gas phase because they have a magnetic atom, (ii) an extended π-orbital is expected to have a large overlap with the metal surface likely driving the π-system into a mixed valence regime instead of a Kondo one. Hence, the appearance of Kondo physics in these systems depends on a series of parameters where common wisdom is likely to fail. It is then of great interest to perform calculations to rationalize the particular features of adsorbed large molecules that depend as little as possible on adjustable parameters.

We have implemented an impurity solver for the Anderson Hamiltonian [14] that reads the one-electron structure from a density functional theory (DFT) calculation for a given spin configuration of the impurity (in the present case the molecular orbitals involved in the spin configuration), uses the DFT hybridization to compute the dynamical electron exchange with the substrate, and assumes a single-electron fluctuation which corresponds to the U limit of the Anderson Hamiltonian. The impurity solver uses the non-crossing approximation NCA [1518] in the multi-orbital formalism by Kuramoto [16]. The multi-orbital aspects of the Kondo problem must be correctly taken into account in realistic accounts of the Kondo problem as shown by Kroha and collaborators [19, 20].

The NCA is a reliable approximation [21] away from very low temperatures where it fails to reproduce the Fermi-liquid behavior in fully screened Kondo systems [22] and where a spurious spectral feature appears at the Fermi energy [22, 23]. These deficiencies can be avoided by using schemes that go beyond NCA such as the coherent T-matrix approximation [24]. At typical experimental temperatures NCA is a good choice for this type of calculation because it retains all the electronic structure of the one-electron part of the Hamiltonian, while keeping a correct description of the main Kondo features. In fact, the main limitation of our theory comes from the use of the customary local or semi-local approximations to DFT. Indeed, DFT calculations on electronic gaps lead to discrepancies of a factor of 2 off the experimental gap [25]. This is definitely a big drawback when evaluating Kondo temperatures (TK) because they depend exponentially on the molecular level value. The Kondo temperature depends on the ratio of the molecular eigenvalue to its broadening, due to the molecular hybridization with the metal substrate. We will use this ratio as a parameter. Yet, our procedure retains the symmetry and the relative strengths of the one-electron Hamiltonian as given by DFT. A second important approximation of our work is that of the infinite intramolecular Coulomb energy, U. Extensions of NCA to treat finite U [2628] lead to considerable improvement of the TK. However, since the LDA-based electronic structure impedes the calculation of the TK, the infinite U approximation largely suffices for our purposes.

In NCA, the Kondo physics is modeled by virtual fluctuations of the impurity occupancy. These fluctuations are made possible by coupling the Kondo impurity to the substrate. In principle, NCA can treat all possible configurations. The U approximation is admissible, provided that the self-energy contribution due to configurations with two electrons or more is negligible. One can generalize this idea and indeed consider two sets of configurations which differ by the addition of one electron, energetically separated from other configurations by some large value 'U' [16]. Hence, with NCA we can treat impurities of increasing complexity, where the physics involved will correspond to virtual fluctuations among configurations differing in electron number by one.

In this way, Roura Bas and Aligia have treated the singlet–triplet quantum transition of an Anderson impurity within NCA [29, 30]. We use a similar approach to describe the configurations of a copper phthalocyanine (CuPc) molecule on Ag(100). On this surface, CuPc captures one electron from the substrate [31] while maintaining a very localized spin on the copper atom at the center of the molecule. Hence, this system is properly characterized by a Hamiltonian describing singlet–triplet transitions. In order to perform our multi-orbital NCA calculations, we first simulate CuPc/Ag(100) with DFT; we transform to a maximally localized Wannier function (MLWF) basis set [32], and solve the DFT Hamiltonian expressed in this basis set with our multi-orbital NCA code, selecting molecular orbitals that take part in the Kondo physics.

The use of MLWF is mandatory to be able to unambiguously transform the DFT Hamiltonian into an Anderson-like one as we have shown in [32]. This is perhaps the biggest difference with other works using impurity solvers based on DFT calculations [3335]. Our approach is thus algorithmic, except for the tuning of the molecular orbital levels with respect to their hybridization with the substrate, since these are quantities where current DFT approaches fail.

In section 2 we give more details regarding the implementation and execution of the calculations on CuPc/Ag(100). In section 3, we first show the DFT results and their conversion to the MLWF basis set. The MLWF results permit us to explore the electronic structure that is involved in Kondo physics. We also evaluate the spectral functions for the lowest unoccupied molecular orbital (LUMO) that is actively involved in the generation of Kondo spectral features. Finally, we analyze our results and conclude this work.

2. Method

The NCA is particularly adequate for describing the electronic structure at different energy scales. Our multi-orbital implementation [16] uses the mean-field results of a local density approximation (LDA) calculation, transforms the LDA Hamiltonian into a MLWF basis which permits us to write an Anderson Hamiltonian, U, while keeping the full multi-configurational aspects of the problem [32]. The NCA is applied on the obtained Anderson Hamiltonian. In this way, the calculated electronic structure contains all molecular + substrate information in the presence of Kondo physics. In this section, we give details on how this is achieved, with special care in the choice of the configurations that will determine the Kondo physics of CuPc on Ag(100).

2.1. Density functional calculations

Density functional calculations of gas phase copper phthalocyanine (CuPc) capture the relevant electronic and geometric properties of the molecule [36]. Briefly, CuPc is a D4h molecule, figure 1, where the d electrons of the central copper atom are split by the ligand field of the surrounding atoms. The ligands capture ∼2 electrons of the Cu atom rendering the molecule in a magnetic d9 configuration [36]. The d manifold is split such that the dx2y2 orbital is singly occupied (SOMO, the singly occupied molecular orbital). The following empty orbital (LUMO, lowest unoccupied molecular orbital) has π character and double degeneracy, because it constitutes the eg representation [36] of the point group.

Figure 1.

Figure 1. Ball-and-stick scheme of a copper phthalocyanine molecule. The central atom is the copper one, and the rest of atoms are nitrogen, carbon and hydrogen in increasing distance from the central one. The free molecule has a D4h symmetry which is reduced to C4 upon adsorption on Ag(100) [31].

Standard image

We have used the Siesta code [37], relaxing the molecule and first two surface layers to forces below 0.04 eV  Å−1using the geometry of [31]. The super-cell contains five layers of 7 × 7 Ag atoms. Norm-conserving pseudo-potentials [38] and strictly localized DZP basis set [39] are used, including an improved noble-metal surface description [40]. The Kohn–Sham Bloch functions were calculated on a Monkhorst–Pack grid 2 × 2 × 1 and transformed to the basis of maximally localized Wannier functions [4143] (MLWF). Our method for obtaining MLWF from Siesta is described in [32].

2.2. Maximally localized Wannier functions

The main reasons for working with a MLWF basis set are: (i) the reduction of the computational problem by projecting the M-dimensional Hilbert space of Kohn–Sham states, where M is the full dimension of the CuPc plus Ag(100) electronic problem, onto a smaller N-dimensional space which faithfully represents [42] the energy bands around the Fermi level taking part in the Kondo effect; (ii) the orthogonality of the MLWF basis set that permits us to use standard multi-orbital NCA [16, 22]; and (iii) the extreme localization of MLWF plus their orthogonality gives us a natural way to partition the problem into impurity and substrate subspaces [32].

Maximally localized Wannier functions have been used in the study of strongly correlated matter (for instance see [4447]). In some cases, the spread minimization is skipped and the Kohn–Sham electronic structure is projected directly onto trial orbitals. This strategy has been used in the study of transition metal oxides [48]. In our system, we found that the projection onto trial orbitals gives unreliable results. The disentanglement method [42] can be contrasted with a direct selection of certain bands of interest. This is often done in the study of transition metal oxides where bands bear strong orbital character [48, 49]. In the adsorbate problem, the band structure is far more complex, which impedes direct band selection. We conclude that the MLWF are an optimal choice for the problem of a large molecule on a metallic substrate.

For the substrate's MLWF, we choose to describe the s–p bands only. In order to achieve this, the s–p bands of the slab are generated by interstitial Wannier functions, in the same way as the surface state of Cu(111) was achieved in [32]. This is a reasonable selection, as long as Kondo physics is considered, because the d bands start at ∼−3 eV below the Fermi energy. This is away from the relevant region in energy with an energy scale several orders of magnitude larger than the typical Kondo scale. Additionally, d states are rather localized, presenting small coupling with molecular states. For these two reasons d states can be safely omitted in the description of the substrate electronic structure taking place in the Kondo effect. The choice of the interstitial centers for the MLWF is delicate because CuPc on Ag(100) displays a C4 symmetry [31] and it is crucial to ensure that MLWF do not reduce the symmetry of the problem. An unexpected problem of the obtention of MLWF for an extended surface is that convergence to MLWF is more difficult as the lateral dimensions of the super-cell are increased. The surface had to be repeatedly tested to achieve a good description of its electronic bands within 2 eV of the Fermi energy.

The molecule's MLWF are obtained jointly with the substrate ones by using the disentanglement method [42]. In order to achieve the clear partitioning between molecule and substrate, the initial set of trial functions consists of 5 d orbitals of the central copper atom (see figure 1), 16 s states for C–H bonds, 32 pz orbitals for every carbon atom, 24 orbitals for the nitrogen atoms and 40 s orbitals for C–C and C–N bonds. This Wannier function description was tested in the two separate systems: the clean Ag(100) surface and the gas phase copper phthalocyanine molecule.

2.2.1. One-particle Hamiltonian.

In [32], we showed how to obtain an Anderson-like Hamiltonian from a Kohn–Sham Hamiltonian in a MLWF basis set. There, the evaluation of the main quantities of an atomic magnetic impurity in a non-magnetic host was described. A special emphasis was put on obtaining the intra-atomic Coulomb energy. In the present section, we extend that study to the case of molecules. The Hamiltonian terms have to be expressed in a suitable way to be able to apply the NCA scheme to account for Kondo physics on molecular orbitals.

Thanks to the localization and orthogonality of the MLWF basis, the Wannier-projected Kohn–Sham Hamiltonian can be organized into four blocks

Equation (1)

where Hmol contains matrix elements between molecular Wannier functions, the second block following the diagonal, Hslab, refers to Wannier functions of the silver slab, while the off-diagonal blocks, V', involve couplings between molecule and slab.

The eigenstates of Hmol are the molecular orbitals now obtained for the MLWF transformed Hamiltonian. As we shall see below, CuPc is a magnetic molecule because one electron is kept in copper's dx2y2 orbital. Furthermore, upon adsorption CuPc captures one more electron in the first π-orbital, the doubly degenerated LUMO. Hence, this open-shell structure, when hybridized with the substrate, gives rise to the electron fluctuations of the Kondo effect. These states are singled out of the problem and we define a special subspace for them, Script Himp, the 'impurity' subspace. The rest of the states are lumped together into the substrate's subspace, Script Hsubs.

Hence, we first apply to (1) a unitary transformation that diagonalizes Hmol and leaves Hslab untouched. In this way, we obtain the molecular orbitals of the molecular part of the Hamiltonian, so that the first block of (1) becomes diagonal. Explicitly,

Equation (2)

The unitary matrix T defines the transformation from MLWF of the molecule to molecular orbitals. The second step is to choose the molecular orbitals that play a role in the Kondo effect. These orbitals are placed in the first rows and columns of (2) and define the subspace Script Himp.

After this rearrangement, the matrix of the Wannier-projected Kohn–Sham Hamiltonian reads

Equation (3)

where the first block is diagonal and contains selected eigenenergies of Hmol, the block Hsubs contains Hamiltonian matrix elements in Script Hsubs and the block V contains couplings between states in Script Himp and Script Hsubs. The molecular orbitals that do not directly intervene in the Kondo effect are included in the substrate. Hence, the full mean-field structure of the DFT calculation is preserved within the inner window of the MLWF transformation which is 2 eV around the Fermi energy in the present calculation.

In the present case, the obtention of an Anderson-like Hamiltonian is facilitated by the small values of the intrinsic U term in LDA. As shown in [32], one can extract the intrinsic U term when writing the full Kohn–Sham Hamiltonian in a Wannier basis set. In that reference, the Co atom impurity had values of U around 1 eV. For the present method to work, one should subtract the intrinsic U term such as is done in the LDA + U method [50, 51] or in recent parameterizations of model Hamiltonians [52]. Anisimov and co-workers show that this intrinsic U term is related to the Hund's rule exchange and, thus, is much smaller than the actual Coulomb U values. This is particularly true in the case of molecules. In the case of CuPc, we have evaluated the intrinsic U-term to be ∼30 meV, and hence negligible in front of the molecular level values and hybridization function.

2.2.2. Ab initio calculation of a hybridization function.

The hybridization function has a fundamental role in the impurity physics [53]. It is of uttermost importance to make accurate evaluations of this function for numerical applications. In order to achieve this, we diagonalize Hsubs, (3), obtaining Bloch states |kn〉 and Bloch energies epsilonkn, n is the band index and k is from the discretized Brillouin zone with 50 × 50 × 1 k-points. Convergency on k-points was checked by a four-fold increase of the k-point sampling. The couplings V of (3) are transformed to the |kn〉 basis accordingly; we label them Vkn,m, where m indices states in Script Himp. After these manipulations, we can calculate

Equation (4)

from the one-particle Hamiltonian (3). It is a matrix in Script Himp.

We emphasize that the Hamiltonian (3) is obtained by projecting the Kohn–Sham electronic structure onto a set of N MLWF. The smaller N-dimensional subspace spanned by MLWF is designed in order to represent the selected energy bands around the Fermi level, with the same level of accuracy as the original bands of the M-dimensional space of Kohn–Sham states. The total number of Wannier functions N is the sum of the MLWF of the molecule and of the substrate. The Wannier description has been tested in separate CuPc and Ag(100) systems. M usually means the highest band output from the ab initio calculation. Interestingly, we found that it is very important to verify the convergence in M. Inclusion of bands with energy of even tens of eV above the Fermi level is vital and preconditions the correct projection onto the Wannier space. As a consequence, the important value of Γmm(ω) at the Fermi level turns incorrect if M is too small. In the present case, convergence was attained for N = 705 and M = 3000.

Formally, (3) becomes the one-particle part of the Anderson Hamiltonian on bringing Hsubs to a diagonal form. Although the procedure we have just introduced is straightforward and essentially algorithmic, it is not correct in principle, because the on-site energies of molecular orbitals in Himp contain a certain part of the Coulomb energy that is included in the mean-field-like LDA framework [32]. Furthermore, the inadequacy of Kohn–Sham energies to describe true one-electron energies yields an important uncertainty when evaluating the Kondo temperature. At the same time, the hybridization is strongly dependent on the distance of the molecule to the surface. Small variations of these two quantities (one-electron energies and molecule–surface hybridization) exponentially propagate in the Kondo temperature. Present ab initio methods cannot then predict the Kondo temperature of a molecular system.

2.3. Impurity electronic configurations

In this work, the one-particle Hamiltonian Himp is replaced by a Hamiltonian with many-body interactions. This subsection discusses the physical grounds on which is designed.

The experimental analysis on the Kondo features in CuPc/Ag(100) [13] indicates that the two-fold degenerate eg LUMO has the main role in Kondo fluctuations. However, DFT calculations [31] show that the spin in the SOMO orbital is not quenched by charge transfer from the molecule. Hence, the observed Kondo resonance is interpreted [13] as due to a spin S = 1 of the molecule, formed by an electron in LUMO and one electron in SOMO. The LUMO electron is captured from the substrate upon adsorption.

We adopt this interpretation and model the molecule as an impurity with a two-electron S = 1 ground state. Furthermore, the lifetime of the SOMO will be assumed an order of magnitude longer than the LUMO's, as will be confirmed by ab initio calculation in section 3.1.2.

The electron in the SOMO will be represented by a local magnetic moment which interacts via exchange with the hybridized LUMO. The multi-orbital structure of the problem appears in the two-fold degeneracy of the LUMO. The configurations that we will study are then the ones coming from the filling of the LUMO, figure 2.

Figure 2.

Figure 2. Considered electronic configurations for the two-fold LUMO (eg): (a) The empty orbital (denoted by eg 0), (b) singly occupied orbital eg 1, (c) doubly occupied configuration eg 2, (d) doubly occupied configuration eg 11.

Standard image

We choose the relevant configurations by estimating their energetic accessibility in the charge fluctuation process. Then, the relevant parameters determining the electronic configuration are: the LUMO on-site energy epsilonL and the charging energy U. The lowest-energy molecular configuration for the adsorbed molecule is singly occupied, as suggested by DFT calculations [31]. We can estimate the free molecule U by evaluating the energies of the neutral (E0), singly (EI) and doubly (EII) negatively charged molecule. Assuming a simple impurity Hamiltonian, then the first affinity is given by

Equation (5)

and the second affinity by

Equation (6)

From these equations and the total energy calculations for the free molecular species we obtain that U = 2.08 eV. However, U is substantially screened on the metallic surface. A constrained DFT calculation for π-systems on silver surfaces leads to a reduction of U to values between 0.5 and 1.0 eV [54]. In the present case, the CuPc molecule lies at ∼3 Å from the surface, reducing the metal screening. Hence, since epsilonL = −0.35 eV (see section 3.2), which leads to |epsilonL| |< U and the doubly occupied configurations eg 2, eg 11, figures 2(c) and (d), are higher in energy than the neutral one, figure 2(a). Thus, we will consider eg 1–eg 0 fluctuations.

Hence, the impurity Hamiltonian contains the eigenvalues of the two LUMO as given by the diagonalization of the MLWF Hamiltonian (2), plus the exchange interaction I with the spin of the SOMO:

Equation (7)

We introduce Hubbard operators which automatically restrict the LUMO occupancy to zero or one. The sums are over spin and orbital degrees of freedom, the latter indexed by a. The second term in the Hamiltonian is the direct exchange interaction involving the spin S2 of the SOMO and the spin operator of the LUMO expressed through Pauli matrices τ as

If I = 0 the Schrieffer–Wolff projection of the model onto the eg 1 manifold leads to a SU(4) Coqblin–Schrieffer model [55]. The evolution of spectral features with I leads to a quantum phase transition of the singlet–triplet impurity problem, that has recently received much attention experimentally [56] and theoretically [29, 30]. Here, we will study three cases: I < 0, I = 0 and I>0.

2.4. Multi-orbital non-crossing approximation

The above impurity Hamiltonian (7) commutes both with the LUMO occupancy operator, na = ∑σ|aσ〉〈aσ|, and the total spin operator S = S1 + S2 showing that it properly describes the molecular properties. Due to the hybridization with the substrate, the molecule will exchange electrons. Here, we assume just fluctuations between the two above configurations, figures 2(a) and (b) plus the fluctuations of the SOMO spin, as included in the definition of .

The full Anderson-like Hamiltonian reads

Equation (8a)

Equation (8b)

The hybridization part is written with Hubbard operators which change the state of LUMO from empty to occupied or vice versa. Substrate electrons are described by fermionic operators ckσ and energies epsilonkσ. For the sake of brevity, all substrate electronic degrees of freedom are encapsulated in the k symbol. The band index will be introduced when necessary.

Let . Following Bickers [18], we write down a resolvent operator of the impurity

as a result of averaging over eigenstates |Ω〉 of the non-interacting substrate with energies EΩ. The substrate partition function is denoted by Zsubs. Since does not mix different occupancies, it can be written as a block matrix

where refers to one-electron occupancy (i.e. LUMO empty) and acts on two-electron occupancies of the impurity.

These quantities are calculated via self-energies defined by

Following Kuramoto [16], we introduce basis states labeled by α for the configurations with one electron and β for the two-electron configurations. In the non-crossing approximation the self-energies for the fixed occupations are given by all diagrams without crossings of substrate electron lines,

The hybridization vertex Vkσ(α|β) comes when emitting an electron from LUMO to the band state kσ which is accompanied by impurity transition from β to the state with label α. Similarly, Vkσ(β|α) is brought about when annihilating a one-electron state α of the impurity and creating a two-electron state by absorbing a substrate electron. Explicitly,

Equation (9a)

Equation (9b)

It is convenient to re-express the fixed-occupation self-energies (also called bosonic and fermionic self-energies in slave-boson approaches [17, 22, 30]) in terms of a hybridization function that is directly related to the level broadening of the impurity configurations

Equation (10)

Please, notice that we have not included π or 2π factors as sometimes is done in the literature. In order to evaluate lifetimes a factor 2π will be added to the diagonal terms of the hybridization function (10).

We can now represent the fixed-occupation self-energies in the form

Equation (11a)

Equation (11b)

This permits us to efficiently perform all computations using fast Fourier transforms.

Let us write β = (a,S,Sz), where a = 1,2 indicates the two orbitals of LUMO, S the total spin and Sz one of its components. Similarly, α denotes the projection of the spin of SOMO on the z-axis, the only degree of freedom of the one-electron configurations.

In the present case, the substrate is non-magnetic, epsilonkσ and Vkσ,a do not depend on the electron spin, σ, and the expression for the hybridization factorizes into spin and orbital parts (see appendix). This considerably simplifies the equations (11). Introducing

Equation (12a)

Equation (12b)

Equation (12c)

and defining the orbital part of the hybridization function by (see (A.1))

Equation (13)

we can write

Equation (14a)

Equation (14b)

The resolvent for the two-electron configurations now depends on the total spin and is a matrix in the orbital space. The equations (12) and (14) constitute a self-consistent system.

The main quantity of interest, the Green's function of LUMO, will be calculated from the general time-ordered correlation function of Hubbard operators, defined through the thermal average

Equation (15)

Starting from the latter expression, we average over the singly occupied configurations αα' and take the Fourier transform [16] to obtain the Green's function of LUMO for the given spin multiplet

Equation (16)

The quantities A2,aa' and A1 are the spectral functions of the resolvents and Zi is the impurity partition function. We follow Kuramoto [16] and use defect propagators in the numerical implementation of (16), see appendix.

The spectral function of LUMO is calculated by tracing over orbital and spin degrees of freedom,

Equation (17)

3. Results: Kondo physics of CuPc on Ag(100)

Full detail on the adsorption of CuPc on Ag(100) is given in other publications [31, 60]. Here, we build on those results and apply our methodology to obtain the electronic structure in the presence of the Kondo effect. First, we build the Anderson-like Hamiltonian following the recipes of section 2.2.1 and next, we solve the multi-orbital NCA equations, section 2.4, presenting the spectral function or PDOS onto CuPc doubly degenerated LUMO. Special care is given to the evaluation and presentation of the hybridization function, Γmm'(ω), of section 2.2.2 because most of the relevant Kondo physics is associated with the symmetry and values of this function.

3.1. Hamiltonian for CuPc on Ag(100)

3.1.1. Molecular orbitals.

Thanks to the partitioning made possible by the MLWF basis, we can extract the molecular Hamiltonian Hmol (1) computed from the LDA calculation of CuPc on Ag(100) [31, 60]. Table 1 presents eigenenergies of Hmol the molecular orbitals close to the substrate's Fermi level. The orbitals are then closely related to the gas phase CuPc ones and hence, we denote them by SOMO, LUMO and highest occupied molecular orbital (HOMO) labels. For comparison, the results for the LDA Siesta calculation of the gas phase molecule of the same molecular geometry are also given. The MLWF qualitatively reproduce the gas phase molecular levels. We emphasize that the epsilon (MLWF) refer to molecular orbitals screened by the metal but without direct hybridization and the epsilon (Siesta) are for a true gas phase molecule without any screening or coupling with an external metal.

Table 1.  Eigenenergies of the molecular orbitals for the impurity Hamiltonian evaluated with MLWF using (2), compared with the gas phase LDA calculation using Siesta, for the same geometry of the CuPc molecule. All energies in the third column were shifted, so that the energies of SOMO (↓) coincide. Energies of the second LUMO state are in parenthesis. The comparison is only indicative that the MLWF capture the molecular properties since the two calculations are performed for different systems: the MLWF refer to molecular orbitals screened by the metal but without direct hybridization and the Siesta column is for a true gas phase molecule without any screening or coupling with an external metal.

  epsilon (MLWF) (eV) epsilon (Siesta) (eV)
HOMO (↑) −1.110 −0.888
HOMO (↓) −1.112 −0.890
SOMO (↑) −0.549 −0.714
SOMO (↓) 0.040 0.040
LUMO (↑) 0.185 (0.190) 0.496 (0.498)
LUMO (↓) 0.211 (0.216) 0.520 (0.522)

Figure 3 shows the PDOS onto these four molecular orbitals. The calculation is performed for the LDA k-point grid (2 × 2 × 1) which is clearly insufficient to account for the continuum character of the substrate electronic states. Hence, the peaks have been slightly broadened using a Gaussian broadening of 50 meV. Despite the qualitative character of this figure, much information can be gleaned from it. The difference in widths between SOMO on the one hand and LUMO and HOMO on the other hand is substantial. The SOMO peak basically presents the numerical width, and is a featureless peak, while the LUMO shows oscillations and spreads over several hundreds of meV. Then, the latter orbitals hybridize more strongly with the substrate. From this figure we can conclude that while the LUMO width is in the range of a few hundreds of meV, the SOMO is significantly less. This is in agreement with the move involved calculations of the hybridization function that we present in section 3.1.2.

Figure 3.

Figure 3. Density of states projected onto molecular orbitals (PDOS). The projection onto the SOMO is given by bold lines, onto the HOMO by dashed lines and onto the LUMO by dash-dotted lines. Energies are with respect to the Fermi energy. Majority spin is shown in the upper panel and minority in the lower one. The PDOS are convoluted with a 50 meV Gaussian.

Standard image

The PDOS are in excellent agreement with the PDOS on molecular orbitals from DFT results [60]. Both the eigenenergies and the PDOS give support to our method of transforming the DFT Hamiltonian to a MLWF and selecting the impurity Hamiltonian using (3).

It is difficult to conclude on the actual occupancies from the PDOS. Indeed, the PDOS numerical broadening thwarts any precise calculation of orbital occupancies, and the definition itself of occupancy of a molecular orbital in a chemisorbed system is somewhat arbitrary. As we show in the next section, the actual peak width of SOMO is negligible; the SOMO then remains singly occupied as in the gas phase. The LUMO peaks in figure 3 cross the Fermi level, hence these orbitals capture charge from the substrate. The LUMO are then the only orbitals that participate in charge transfer from the surface.

3.1.2. Hybridization function.

The calculation of (4) proceeds along the lines described in section 2.2.2. The main ingredients of this calculation are the Bloch eigenvalues and the molecule–substrate couplings. The Bloch eigenstates span the substrate continuum, and this task is much simplified thanks to the reduced computational effort that the Wannier basis set introduces. Indeed, this allows us to use a very dense k-point sampling mesh and hence, accurate results. The couplings are obtained after rotation into the molecular orbital basis set (3) of the couplings obtained after the Wannierization process. This procedure is numerically very reliable because it assures perfect agreement with the LDA results within an energy window of a few eV about the Fermi energy. Hence, the shortcomings of our procedure to compute the hybridization function are the ones inherent to the LDA in the present case. The molecule–substrate couplings depend exponentially on the molecule–surface distance. This means that the determination of the system's geometry is very important to obtain accurate hybridization values. Otherwise, we expect LDA values of the hybridization function to be accurate because the hybridization function evaluates the transparency of the barrier between the molecule and the substrate and good numerical results have been previously obtained  [61].

The off-diagonal parts of Γmm'(ω) are very small (<1 meV). This is a very strong result that shows that the substrate does not mix the different molecular orbitals among themselves. Hence, every molecular orbital defines an electronic channel of the system.

The diagonal elements for SOMO and LUMO orbitals are presented in figure 4, along with the density of states of the substrate. The three curves are very similar in shape. By comparing the hybridization functions with the substrate DOS, we notice that the hybridization function grows more slowly than the DOS as the electron energy increases. This is an effect due to the couplings, V, between the molecular orbitals and the substrate electronic structure. However, at higher energy, it is the DOS that controls the behavior of the hybridization function with energy. Figure 4 shows that the SOMO width has to be multiplied by 20 to be comparable to the LUMO one. The SOMO orbital has then a very small mixing with the substrate as compared to the LUMO. Finally, in the U = 0 picture, the FWHM for the LUMO is 2πΓaa(ω = epsilonL) = 440 meV and is comparable to the FWHM of the PDOS peak, figure 3, obtained with an insufficient k-point sampling.

Figure 4.

Figure 4. Diagonal elements of the hybridization function matrix (in eV) as a function of the electron energy, ω, with respect to the Fermi level. Only the LUMO and the SOMO (note the factor 20) are considered. For comparison, the substrate-projected density of states is given in arbitrary units. The delta function of (4) has been replaced by a 5 meV wide Gaussian; the k-point sampling is 50 × 50 × 1.

Standard image

We have also evaluated the hybridization function of the LUMO with d orbitals by preparing a Wannier basis set with d orbitals. In the region of interest here (some 2 eV about the Fermi energy), this hybridization is strictly zero due to the lack of d states at these energies. However, when resonant with Ag d-band (3 eV below the Fermi energy), the hybridization function becomes larger than the corresponding values for the sp Wannier functions. At 4 eV below the Fermi energy, the d-electron hybridization has a maximum of 0.032 eV, while the sp one is at 0.035 eV. The consequence of this is that the spectral function at −4 eV will not be just a simple Lorentzian tail. However, the LUMO orbital is some hundreds of meV away from the Fermi energy, and the effect of the d-band contribution to the overall shape of the LUMO spectral function is negligible both for Kondo physics and for the one-electron spectral shape.

Finally, we comment on the problem of broken spin symmetry in DFT. Spin-polarized LDA implies two subsystems: minority and majority spin. This in turn says that we have two sets of Wannier functions, two distinct substrates, hybridization functions, etc. The main effect of breaking the spin invariance in our DFT calculation is that the SOMO occupancy is nSOMO≈1. As a secondary effect, the substrate and molecule become slightly spin-polarized as well. However, this effect is perturbational [14] and is not an intrinsic property of the bare substrate and impurity of the Anderson model. In what follows we drop the minority spin data of the hybridization function and restore the spin symmetry.

3.2. Multi-orbital NCA results

Our LDA calculations yield a hybridization function for the SOMO, ΓS, ten times smaller than the one for the LUMO levels. From these values we obtain ΓS≈4.5 meV. We use the standard expression for the Kondo temperature [62]

Equation (18)

and take the ab initio value for the on-site energy from the table 1. The bandwidth is 2D, where a rectangular DOS is typically assumed. Here we have taken D as 10 eV because the DOS of figure 3 integrates to 703 states, while the calculated DOS at the Fermi energy is 32.28 eV −1. Hence 2D = 703/32.28 = 21.8 eV, and D turns out roughly 10 eV. We get TK,S of the range 10−26 eV . Thus, the energy scale that would correspond to a Kondo effect on the SOMO orbital (without exchange coupling to LUMO) is unobservable.

These arguments show that the Kondo coupling in this system is indeed given by virtual charge fluctuations of the two-fold degenerate LUMO. Hence, we are dealing with the impurity problem described in section 2.3, for which we can calculate the spectral function according to the section 2.4.

However, the on-site energies of LUMO in (7), as given by their LDA values in the table 1, lie very close to the Fermi level, which would correspond to a fluctuating valence regime. That has not been observed in experiment [13]. This is related to the fact that the LDA energies of LUMO do not reflect the considerable Coulomb repulsion in their spin splitting. We fix these deficiencies by rescaling the epsilonLaa(ω) ratio in order to achieve a SU(4) Kondo temperature, TK,L = Dexp(−|epsilonL|/4Γaa), of ∼20 K as observed in the measurements [13]. The results presented in this section are calculated with epsilonL = −0.35 eV and Γaa(epsilonL) rescaled to 0.01 eV.

For temperatures below TK,L the model Hamiltonian (7) and (8) exhibits a rich variety of physics as the value of I changes. Four regimes can be identified: (i) I>0, I>TK,L, (ii) I < 0, |I|>TK,L, (iii) I>0, I < TK,L and (iv) I < 0, |I| |< TK,L.

In the first case, taking the limit I, the Kondo physics is that of the underscreened Kondo effect. Hence, the model will show singular Fermi-liquid characteristics [6365] in the low-energy domain. The physics in (ii) is dominated by the molecular spin zero states. In the limit I→− we obtain a spin-less molecule with orbital pseudo-spin, which is over-screened by two substrate channels. The cross-over temperature is, however, exponentially smaller than TK,L, due to the reduced degeneracy. In the regime (iii) the SOMO is effectively decoupled from the LUMO and we recover the case of two degenerate LUMO on the same footing as the electron spin: we have an SU(4) system as described above. Finally, in (iv), the negative I allows for the Kondo screening of the SOMO by a two-stage Kondo effect [66] at very low temperatures. Here we present NCA spectral functions in the I = 0 case and in the intermediate regimes Γaa>|I|>TK,L which are dominated by inelastic spin transitions.

Figure 5 shows the spectral function of LUMO at T = 7 K when the exchange interaction between SOMO and LUMO is turned off, I = 0. The ground state of is four-fold degenerate in this case. Since each of the two orbitals of LUMO defines an independent scattering channel in the substrate (i.e. Γaa'(ω) is diagonal), the LUMO is subject to a SU(4) Kondo effect. The spectral function shows a broad charge-excitation peak of the FWHM given by 4×2πΓLL(epsilonL) and a narrow Kondo peak.

Figure 5.

Figure 5. LUMO spectral function for the I = 0 case. The wide peak corresponds to a charge-excitation and the narrow peak is the Kondo peak (shown also in the inset).

Standard image

When the SOMO and LUMO spins are subject to a ferromagnetic interaction (I>0) the ground state of the molecule without couplings to the substrate is an orbitally degenerate S = 1. The excited state is a S = 0 orbital doublet. Now, two satellites develop at ±I from the Kondo peak, figure 6. These satellites are inelastic replicas of the Kondo peak since the spin-excitation energy is exactly I. Decomposition into spin channels yields that the Stokes ( + I) satellite is in the S = 0 channel, while the anti-Stokes peak as well as the Kondo peak are in the S = 1 channel. This result can be understood by recalling the meaning of the spectral function. The spectral function at T = 0 yields the probability density to inject one electron in the system when ω>0 or to inject a hole when ω < 0. Hence, the positive energy satellite corresponds to an excited Kondo effect triggered by the injection of an electron to the S = 0 state, which is an excited state of . Similar results have been found by Roura Bas and Aligia [29, 30].

Figure 6.

Figure 6. LUMO spectral function for I = 25 meV (bold line). The spin zero and spin one channel contributions are presented. The positive energy satellite corresponds to injection of one electron in the excited electronic structure of the system: the spin zero channel. The negative energy peak corresponds to removing one electron from the molecular ground state in the S = 1 channel.

Standard image

When the interaction is anti-ferromagnetic (I < 0), the ground state of is an orbital doublet and is followed by six excited states of spin one. A noteworthy feature of the spectral function are the steps typical for an inelastic spin-flip transition (see for example [57, 58] and references therein). The inelastic steps enhanced by Kondo effect have already been explored in the case of singlet–triplet transitions in nanotubes [67]. The spin channel analysis yields that the Stokes peak is now S = 1 which again can be rationalized by noticing that it corresponds to injecting an electron in an excited S = 1 state and the transition is enhanced by an excited Kondo effect. By the same token, the anti-Stokes peak is S = 0 since a hole is created in this channel.

The spectral function, figure 7, shows a small peak on the Fermi level. This peak cannot correspond to a spin-flip Kondo effect. The orbital Kondo resonance is ruled out by its exponentially suppressed energy scale ∝exp(−|epsilonL|/2Γaa) as compared to TK,L. NCA is known to produce spurious peaks at the Fermi level [15, 23, 18]. The temperature scale at which they appear is given by [62] TS = TK,L/5[TK,Laa]5/3 = 0.18 K (strictly valid for I = 0 only). All calculations shown here are performed at 7 K, well above the pathology temperature TS. We conjecture that the zero energy structure is related to the problems of NCA to reproduce spin-split Kondo peaks due to the artificial flow of the marginal potential scattering term as shown in [22, 68].

Figure 7.

Figure 7. LUMO spectral function for I = −25 meV. The spin zero and spin one components are presented. The two excitation steps correspond to inelastic transitions enhanced by the Kondo effect. As in the S = 1 case, I = 25 meV, the steps belong to a channel with well-defined spin. The positive energy step is given by the spin triplet (S = 1) channel and the low-energy peak is due to the S = 0 channel.

Standard image

Experiments with the STM measure the conductance as a function of bias. As a first approximation, the conductance is basically proportional to the substrate spectral function, justifying our using of the spectral function as a first approach to the experiment. However, a more detailed description of the electron transmission process from the STM tip [69] shows that the transmission channels can interfere, giving rise to a complex Fano structure instead of the above peaked spectral function. In [8] the different Fano structures are classified according to the prevalence of one transmission channel above the others. Hence, peaks are expected if the channel involving the adsorbate is preferential in the conductance process, while dips appear in the conductance spectra if the direct contribution of the surface overwhelms the transport process. In the present case [13], peaks are found, suggesting that transport in such a large molecule is molecule-dominated, however recent reports on porphyrins [8] show a dip in the conductance despite the similarity between both systems. Hence, a detailed conductance calculation is necessary to be able to compare with the experimental data.

4. Discussion

The evaluated LDA spectral function, figure 3, seems to suggest a fluctuating valence state while a Kondo peak has been experimentally found [13]. Moreover, if we added a strong Coulomb term to the LUMO, the corresponding Anderson Hamiltonian would correspond to an empty-impurity regime according to scaling theory [70]. For these reasons, we had to rescale the Kondo temperature in our model by shifting the LUMO level and changing the hybridization strength as we showed in section 3.1.2.

These results show that LDA is unreliable to furnish quantitative data. Any ab initio method is bound to failure when trying to estimate the Kondo temperature given the exponential dependence on the main ab initio ingredients: the level position and hybridization. Nevertheless, ab initio calculations can give valuable qualitative input since the correct electronic symmetry addressing both the spin and orbital channels can be directly obtained from DFT. Our analysis of the hybridization function has yielded important information: (i) the existence of two well-defined orbital channels originating in the hybridization of the two LUMO with the substrate has been proved because non-diagonal terms of Γaa'(ω) are vanishingly small, (ii) both LUMO and SOMO are partially charged in the adsorbed system, the LUMO due to the sizable values of Γaa that lead to charge transfer from the substrate, and the SOMO for the vanishing value of ΓS that leaves its spin unperturbed. Hence, LDA reveals the orbital SU(2) symmetry associated with LUMO electron and its substrate channels.

The failure of the present LDA calculations to yield a realistic epsilonLaa ratio is already present in its inability to characterize the electronic structure of the molecule–substrate system. As we just saw, the PDOS onto Kohn–Sham states predicts the system to be in a mixed valence regime while the experimental data show it is a Kondo system. This points out the present failures, namely, the LUMO occupation is poorly accounted for in LDA. Previous works have claimed that LDA is not capable to yield correct hybridization functions, because of the lack of charge discontinuity in the exchange-and-correlation functional [71]. While this is clearly the case in transport calculations, the present hybridization function describes a static situation, where the strength of the coupling to the substrate is evaluated. Furthermore, the hybridization function evaluates the transparency of the barrier between the molecule and the substrate. Hence, we do not think that the hybridization function is affected by the charge discontinuity problem. We think it is rather the electronic configuration that is affected by the charge discontinuity problem yielding wrong occupancies, the alignment of the molecular levels and finally the wrong qualitative picture.

Finally, the cure has already been advanced in the literature by using LDA + U methods [72]. As in [72], U has to be computed for the LUMO orbitals as has been recently realized by an increasing number of groups [54, 72].

It is for these failures of the LDA calculations that the epsilonLaa had to be adapted in order to obtain reasonable spectral functions in the section 3.2. In the physically relevant case of the positive SOMO–LUMO exchange coupling of I = 25 meV we obtained the experimentally observed three-peak structure in the vicinity of the Fermi level [13].

In spite of the shortcomings stemming from LDA's description of the LUMO shown in this work, we emphasize that the Wannier-based approach is not restricted to the use of the LDA functional and can be interfaced to an arbitrary ab initio method which provides a one-particle structure as for example the promising GW + MLWF method [73]. Moreover, Wannier functions allow us to take a step beyond the super-cell approach and simulate a true impurity problem (i.e. a single molecule on a surface) by reconstructing a semi-infinite substrate out of the tight-binding-like Hamiltonian elements obtained in a super-cell.

5. Conclusion

We have implemented a multi-orbital non-crossing approximation approach based on a standard one-electron ab initio electronic structure. The Kohn–Sham orbitals of a DFT calculation are transformed to a maximally localized Wannier function (MLWF) basis set such that a tight-binding-like Hamiltonian is obtained, in view of the locality and orthogonality of MLWF. This procedure is in principle algorithmic and permits us to have a quantum impurity model from a DFT calculation.

We have applied this methodology to the case of a copper phthalocyanine (CuPc) molecule adsorbed on Ag(100) following existing experimental work [31, 13]. From our DFT calculations we conclude that the two-fold degenerate LUMO captures charge and this can give rise to the Kondo effect. The spin of the SOMO is a spectator because its Kondo energy scale is many orders of magnitude smaller than the one of the LUMO. However, intramolecular exchange interaction between the SOMO and LUMO spin gives rise to a rich singlet–triplet phenomenology that our numerical procedure captures. Our calculations show that two well-defined orbital channels emerge in the substrate as dictated by the C4 point group symmetry. We have further investigated the impurity spectral function in terms of spin channels and we have rationalized the inelastic features of the spectral function for both (spin) singlet and triplet ground states.

This physics has been obtained by rescaling the computed Kondo temperature to fit the experimental one. Indeed, our LDA-based calculation fails to yield a Kondo ground state and rather predicts a fluctuating valence system. We attribute this error to the lack of charge discontinuity in the LDA exchange-and-correlation functional and suggest that alternative LDA + U methods for the one-electron calculation will improve the agreement with the experimental electronic structure of CuPc on Ag(100).

Acknowledgments

We are grateful to A Mugarza, P Gambardella for fruitful discussions and providing their experimental data prior to publication. In particular, we thank A Mugarza for showing us that CuPc on Ag(100) was in a triplet state. Discussions with Jean-Pierre Gauyacq, Roberto Robles and Enric Canadell are also gratefully acknowledged. Financial support from the Spanish MICINN (FIS2009-12721-C04-01) is acknowledged. RK is supported by the CSIC-JAE predoctoral fellowship.

Appendix.: Non-crossing approximation

A.1. Spin coefficients for a singlet–triplet impurity

The couplings between α and β (see (9)) are given by the expression

with Vkσ,a from the Hamiltonian (8) and the bracket is a Clebsch–Gordan coefficient. When the substrate is non-magnetic, epsilonkσ and Vkσ,a do not depend on σ and the expression for hybridization intensity (10) factorizes

into orbital

Equation (A.1)

and spin parts

Equation (A.2)

The subsequent identities have proven significant

Equation (A.3)

Equation (A.4)

Equation (A.5)

The first one is due to completeness (A.2), the second and third can be proven using Wigner 3jm symbols [74].

With these identities it is easy to show that the following property holds: let and be some functions diagonal in the total spin representation, so will the self-energies calculated from NCA (11a) and (11b),

Since we solve the equations of NCA iteratively, starting with resolvents having zero self-energies, we conclude that spin off-diagonal terms of vanish.

A.2. Evaluation of the physical Green's function using defect propagators

Equation (16) has to be expressed in terms of defect propagators. For convenience we introduce boldface notation for matrices in the orbital space of LUMO, i.e. for the resolvent of the two-electron configuration, its spectral density, hybridization function (A.1) and the Green's function of LUMO. The starting expression (16) reads

Equation (A.6)

We introduce the operator fraktur P, whose effect on an arbitrary matrix function X(ω) is given by

Applying fraktur P on both sides of NCA equations (14) yields a self-consistent system

for the defect propagators [23], defined by

When these equations are solved, the Green's function (A.6) can be expressed as

Equation (A.7)

Please wait… references are loading.