Skip to main content

Single-Molecule Conductance Theory Using Different Orbitals for Different Spins: Applications to π-Electrons in Graphene Molecules

  • Conference paper
  • First Online:
Nanophotonics, Nanooptics, Nanobiotechnology, and Their Applications (NANO 2018)

Part of the book series: Springer Proceedings in Physics ((SPPHY,volume 222))

Included in the following conference series:

Abstract

In the present paper, we consider many-body π-electron models and computational tools for treating single-molecule conductance in middle-size graphene molecules. Our study highlights the importance of accounting for long-range interactions and electron-correlation effects which are crucial for a correct description of electron transmission in conjugated systems. Here we construct one-electron Green’s function matrix for the half-projected Hartree-Fock method implementing the different orbitals for different spins (DODS) approach. Moreover, the simplified DODS in the form of quasi-correlated tight-binding (QCTB) and related models are invoked. We compare these electron-correlation models with the usual tight-binding (TB) approximation and show that TB is actually incorrect as a single-molecule conductance theory of graphenic and similar structures. In our specific applications, we calculate the conductance spectra for small graphene nanoflakes and find, in particular, that “zigzag” connections can afford significantly higher electron transmission than “armchair” connections.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Abbreviations

AO:

Atomic orbital

DODS:

Different orbitals for different spins

EHF:

Extended Hartree-Fock

FCI:

Full configuration interaction

GF:

Green’s function

GQD:

Graphene quantum dot

HPHF:

Half-projected Hartree-Fock

MO:

Molecular orbital

MSE:

Molecular-scale electronics

QCLRI:

Quasi-correlated long-range interaction

QCTB:

Quasi-correlated tight-binding (model)

RHF:

Restricted Hartree-Fock

TB:

Tight-binding (model)

UHF:

Unrestricted Hartree-Fock

WBL:

Wide-band limit

References

  1. Cuevas JC, Scheer E (2017) Molecular electronics: an introduction to theory and experiment, 2nd edn. World Scientific, Singapore

    Book  Google Scholar 

  2. Solomon GC, Herrmann C, Ratner MA (2012) Molecular electronic junction transport: some pathways and some ideas. Top Curr Chem 313:1–38

    Google Scholar 

  3. Metzger RM (2015) Unimolecular electronics. Chem Rev 115:5056–5115

    Article  Google Scholar 

  4. Moth-Poulsen K (ed) (2016) Handbook of single-molecule electronics. Pan Stanford Publishing Pte Ltd, Singapore

    Google Scholar 

  5. Xiang D, Wang X, Jia C, Lee T, Guo X (2016) Molecular-scale electronics: from concept to function. Chem Rev 116:4318–4440

    Article  Google Scholar 

  6. Tsuji Y, Estrada E, Movassagh R, Hoffmann R (2018) Quantum interference, graphs, walks, and polynomials. Chem Rev 118:4887–4911

    Article  Google Scholar 

  7. Ernzerhof M, Bahmann H, Goyer F, Zhuang M, Rocheleau P (2006) Electron transmission through aromatic molecules. J Chem Theory Comput 2:1291–1297

    Article  Google Scholar 

  8. Fowler PW, Pickup BT, Todorova TZ, Myrvold W (2009) Conduction in graphenes. J Chem Phys 131:244110-1-8

    ADS  Google Scholar 

  9. Markussen T, Stadler R, Thygesen KS (2010) The relation between structure and quantum interference in single molecule junctions. Nano Lett 10:4260–4265

    Article  ADS  Google Scholar 

  10. Pedersen KGL, Borges A, Hedegård P, Solomon GC, Strange M (2015) Illusory connection between cross-conjugation and quantum interference. J Phys Chem C 119:26919–26924

    Article  Google Scholar 

  11. Tsuji Y, Hoffmann R, Movassagh R, Datta S (2014) Quantum interference in polyenes. J Chem Phys 141:224311-1-13

    Article  ADS  Google Scholar 

  12. Markussen T, Stadler R, Thygesen KS (2011) Graphical prediction of quantum interference-induced transmission nodes in functionalized organic molecules. Phys Chem Chem Phys 13:14311–14317

    Article  Google Scholar 

  13. Wang X, Spataru CD, Hybertsen MS, Millis AJ (2008) Electronic correlation in nanoscale junctions: comparison of the GW approximation to a numerically exact solution of the single-impurity Anderson model. Phys Rev B 77:045119-1-10

    ADS  Google Scholar 

  14. Bergfield JP, Stafford CA (2009) Many-body theory of electronic transport in single-molecule heterojunctions. Phys Rev B 79:245125-1-10

    Article  ADS  Google Scholar 

  15. Yeriskin I, McDermott S, Bartlett RJ, Fagas G, Greer JC (2010) Electronegativity and electron currents in molecular tunnel junctions. J Phys Chem C 114:20564–20568

    Article  Google Scholar 

  16. Bergfield JP, Solomon GC, Stafford CA, Ratner MA (2011) Novel quantum interference effects in transport through molecular radicals. Nano Lett 11:2759–2764

    Article  ADS  Google Scholar 

  17. Goyer F, Ernzerhof M (2011) Correlation effects in molecular conductors. J Chem Phys 134:174101-1-10

    Article  ADS  Google Scholar 

  18. Pedersen KGL, Strange M, Leijnse M, Hedegard P, Solomon GC, Paaske J (2014) Quantum interference in off-resonant transport through single molecules. Phys Rev B 90:125413-1-11

    Article  ADS  Google Scholar 

  19. Hoy EP, Mazziotti DA, Seideman T (2017) Development and application of a 2-electron reduced density matrix approach to electron transport via molecular junctions. J Chem Phys 147:184110-1-8

    Article  ADS  Google Scholar 

  20. Luzanov AV (2019) Single-molecule electronic materials: conductance of π-conjugated oligomers within quasi-correlated tight-binding model. Funct. Mater 26:152–163

    Article  Google Scholar 

  21. Luzanov AV (2014) Effectively unpaired electrons in bipartite lattices within the generalized tight-binding approximation: application to graphene nanoflakes. Funct Mater 21:437–447

    Article  Google Scholar 

  22. Luzanov AV (2016) Effectively unpaired electrons for singlet states: from diatomics to graphene nanoclusters. In: Leszczynski J, Shukla MK (eds) Practical aspects of computational chemistry IV. Springer, Boston, pp 151–206

    Google Scholar 

  23. Luzanov AV, Plasser F, Das A, Lischka H (2017) Evaluation of the quasi correlated tight-binding (QCTB) model for describing polyradical character in polycyclic hydrocarbons. J Chem Phys 146:064106-1-12

    Article  ADS  Google Scholar 

  24. Davison SG, Amos AT (1965) Spin polarized orbitals for localized states in crystals. J Chem Phys 43:2223–2233

    Article  ADS  Google Scholar 

  25. Estrada E (2018) The electron density function of the Hückel (tight-binding) model. Proc R Soc A474:20170721-1-18

    ADS  MATH  Google Scholar 

  26. Smeyers YG, Doreste-Suarez L (1973) Half-projected and projected Hartree-Fock calculations for singlet ground states. I. Four-electron atomic systems. Int J Quantum Chem 7:687–698

    Article  Google Scholar 

  27. Cox PA, Wood MN (1976) The half-projected Hartree-Fock method. I. Eigenvalue formulation and simple application. Theor Chim Acta 41:269–278

    Article  Google Scholar 

  28. Luzanov AV (1985) The spin-symmetrized Hartree-Fock method. J Struct Chem 25:837–844

    Article  Google Scholar 

  29. Bone RGA, Pulay P (1992) Half-projected Hartree-Fock natural orbitals for defining CAS–SCF active spaces. Int J Quant Chem 45:133–166

    Article  Google Scholar 

  30. Smeyers YG (2000) The half projected Hartree-Fock model for determining singlet excited states. Adv Quant Chem 36:253–270

    Article  Google Scholar 

  31. Verzijl CJO, Seldenthuis JS, Thijssen JM (2013) Applicability of the wide-band limit in DFT-based molecular transport calculations. J Chem Phys 138:094102-1-10

    Article  ADS  Google Scholar 

  32. Jhan S-M, Jin B-Y (2017) A simple molecular orbital treatment of current distributions in quantum transport through molecular junctions. J Chem Phys 147:194106-1-10

    Article  ADS  Google Scholar 

  33. Langer W, Plischke M, Mattis D (1969) Existence of two phase transitions in Hubbard model. Phys Rev Lett 23:1448–1452

    Article  ADS  Google Scholar 

  34. Langer W, Mattis D (1971) Ground state energy of Hubbard model. Phys Lett A3:139–140

    Article  ADS  Google Scholar 

  35. Tyutyulkov N (1975) A generalized formula for the energies of alternant molecular orbitals. I. Homonuclear molecules. I J Quantum Chem 9:683–68936

    Article  Google Scholar 

  36. Löwdin P-O (1955) Quantum theory of many-particle systems. III. Extension of the Hartree-Fock scheme to include degenerate systems and correlation effects. Phys Rev 97:1505–1520

    ADS  MathSciNet  Google Scholar 

  37. Waller I, Hartree DR (1929) On the intensity of total scattering of X-rays. Proc R Soc London A124:119–142

    Article  ADS  Google Scholar 

  38. Lyakh DI, Musiał M, Lotrich VF, Bartlett RJ (2011) Multireference nature of chemistry: the coupled-cluster view. Chem Rev 112:182–243

    Article  Google Scholar 

  39. Tada T, Yoshizawa K (2002) Quantum transport effects in nanosized graphite sheets. Chem Phys Chem 3:1035–1037

    Article  Google Scholar 

  40. Morikawa T, Narita S, Klein DJ (2005) Molecular electric conductance and long-bond structure counting for conjugated-carbon nano-structures. Chem Phys Lett 402:554–558

    Article  ADS  Google Scholar 

  41. Schomerus H (2007) Effective contact model for transport through weakly-doped graphene. Phys Rev B 76:045433-1-7

    Article  ADS  Google Scholar 

  42. Cuansing E, Wang JS (2009) Quantum transport in honeycomb lattice ribbons with armchair and zigzag edges coupled to semi-infinite linear chain leads. Euro Phys J B69:505–513

    Article  ADS  Google Scholar 

  43. Nelson T, Zhang B, Prezhdo OV (2010) Detection of Nucleic Acids with Graphene Nanopores: Ab Initio Characterization of a Novel Sequencing Device. Nano Lett 10:3237–3242

    Article  ADS  Google Scholar 

  44. Rangel NL, Leon-Plata PA, Seminario JM (2012) Computational Molecular Engineering for Nanodevices and Nanosystems. In: Leszczynski J, Shukla MK (eds) Practical aspects of computational chemistry I. Springer, Heidelberg, pp 347–383

    Google Scholar 

  45. Qiu W, Skafidas E (2013) Quantum conductance of armchair graphene nanopores with edge impurities. J Chem Phys: 114: 073703–073701–8

    Article  ADS  Google Scholar 

  46. Güçlü AD, Potasz P, Korkusinski M, Hawrylak P (2014) Graphene Quantum Dots. Springer, Berlin/Heidelberg/New York

    Book  Google Scholar 

  47. Luzanov A (2018) Graphene Quantum Dots in Various Many-Electron π-Models. In: Fesenko O, Yatsenko L (eds) Nanophysics, nanophotonics, and applications. Springer proceedings in physics, vol 210. Springer, Cham, pp 161–174

    Google Scholar 

  48. Peng J, Gao W, Gupta BK, Liu Z, Romero-Aburto R, Ge L, Song L, Alemany LB, Zhan X, Gao G, Vithayathil SA, Kaipparettu BA, Marti AA, Hayashi T, Zhu JJ, Ajayan PM (2012) Graphene quantum dots derived from carbon fibers. Nano Lett 12:844–849

    Article  ADS  Google Scholar 

  49. McWeeny R (1992) Methods of molecular quantum mechanics. Academic Press, London

    Google Scholar 

  50. Goscinski O, Lindner P (2003) Natural spin-orbitals and generalized overlap amplitudes. J Math Phys 11:1313–1317

    Article  ADS  Google Scholar 

  51. Albert AE (1972) Regression and the Moore-Penrose pseudoinverse. Academic Press, New York

    MATH  Google Scholar 

  52. Amos AT, Woodward M (1969) Configuration-interaction wavefunctions for small pi systems. J Chem Phys 50:119–123

    Article  ADS  Google Scholar 

  53. Luzanov AV, Ivanov VV, Boichenko IV (1996) Semiempirical determination of Dyson’s states in conjugated systems within a full-CI π-electron scheme. J Mol Struct (THEOCHEM) 360:167–174

    Article  Google Scholar 

  54. Pople JA, Hush NS (1955) Ionization potentials and electron affinities of conjugated hydrocarbon molecules and radicals. Trans Faraday Soc 51:600–605

    Article  Google Scholar 

  55. McLachlan AD (1959) The pairing of electronic states in alternant hydrocarbons. Mol Phys 2:271–284

    Article  ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Appendices

Appendices

1.1 Appendix A: Construction of HPHF Green’s Function

Before treating in detail GF for HPHF, it is sensible to consider a standard general expression of GF. Let us first rewrite Eq. (22.2) in an appropriate spectral form:

$$ R=\sum \limits_i\;\frac{\left|{\phi}_i\right\rangle \left\langle {\phi}_i\right|}{E+i{0}^{+}-{\varepsilon}_i}+\sum \limits_a\;\frac{\left|{\phi}_a\right\rangle \left\langle {\phi}_a\right|}{E+i{0}^{+}-{\varepsilon}_a},\vspace*{-3pt} $$
(22.A1)

where ε i and ε a are Koopmans orbital energies, that is eigenvalues of h; |ϕ i〉 and |ϕ a〉 are corresponding occupied and virtual MOs (eigenkets of h), respectively. In fact, the structure of Eq. (22.A1) remains valid in a more general setting (as in Eq. (5.76) from Ref. [1]). In doing so, |ϕ i〉 and |ϕ a〉 should be replaced by the so-called Dyson orbital \( \left|{d}_i^{+}\right\rangle \) for electron detachment, and by Dyson orbital \( \left|{d}_a^{-}\right\rangle \) for electron attachment; they may be nonorthogonal to each other and even be linear dependent [50]. In addition, ε i and ε a are replaced with transition energies \( \Delta {E}_i^{+} \) and \( \Delta {E}_a^{-} \), respectively. Explicitly, \( \Delta {E}_i^{+}={E}^N-{E}_i^{N-1} \) (negative ionization potential), and \( \Delta {E}_a^{-}={E}_a^{N+1}-{E}^N \) (electron affinity). It gives the most general (Lehmann type) spectral representation of GF for N-electron system:

$$ R=\sum \limits_i\;\frac{\left|{d}_i^{+}\right\rangle \left\langle {d}_i^{+}\right|}{E+i{0}^{+}-\Delta {E}_i^{+}}+\sum \limits_a\;\frac{\left|{d}_a^{-}\right\rangle \left\langle {d}_a^{-}\right|}{E+i{0}^{+}-\Delta {E}_a^{-}}.\vspace*{-3pt} $$
(22.A2)

Now we turn to the HPHF model for which the variational Koopmans-like orbitals were constructed in Ref. [28]. We will need the standard (Hermitian) matrix projectors onto the occupied spin-up and spin-down MOs, that is

$$ {\rho}^{\alpha }=\sum \limits_{i=1}^n\left|{\phi}_i^{\alpha}\right\rangle \left\langle {\phi}_i^{\alpha}\right|,\kern1em {\rho}^{\beta }=\sum \limits_{i=1}^n\left|{\phi}_i^{\beta}\right\rangle \left\langle {\phi}_i^{\beta}\right|, $$
(22.A3)

along with a non-Hermitian matrix projector U which is generated by overlapping of ρ α and ρ β:

$$ U={\rho}^{\alpha }{\left({\rho}^{\beta }{\rho}^{\alpha}\right)}^{-1}{\rho}^{\beta }.\vspace*{-3pt} $$
(22.A4)

Matrix inversion here should be understood as the Moore-Penrose pseudoinverse (see, e.g., Ref. [51]). The next are the Fockian matrices, f α and f β, associated with the above projectors:

$$ {f}_{\alpha }=h+J\left({\rho}^{\alpha }+{\rho}^{\beta}\right)-K\left({\rho}^{\alpha}\right),\kern1em {f}_{\beta }=h+J\left({\rho}^{\alpha }+{\rho}^{\beta}\right)-K\left({\rho}^{\beta}\right)\vspace*{-24pt} $$
(22.A5)
$$ {f}_U=h+J\left(U+{U}^{+}\right)-K(U), $$
(22.A6)

with J and K being, respectively, standard Coulomb and exchange (super)operators due to Roothaan. In above, h is a core Hamiltonian which includes not only h 0 but electron-nuclear attraction terms.

Then we can derive the HPHF variational equation for \( \left|{d}_i^{+}\right\rangle \), based on Eqs. (35) and (36) from Ref. [28]. We first define the (nonnormalized) charge density matrix, D, at the HPHF level:

$$ D={\rho}^{\alpha }+{\rho}^{\beta }+\xi\;\left(\;U+{U}^{+}\right), $$
(22.A7)

where ξ is a pseudodeterminant of ρ α ρ β (i.e., the last nonnull (nth) coefficient of its characteristic polynomial). This D serves as an auxiliary matrix in the generalized eigenvalue problem of the form:

$$ \Theta {D}^{-1}\left|{d}_i^{+}\right\rangle =\left({E}_{\mathrm{HPHF}}-\Delta {E}_i^{+}\right)\left|{d}_i^{+}\right\rangle, $$
(22.A8)

where

$$ \Theta ={\rho}^{\alpha}\left({E}_{\rho }-{f}_{\alpha}\right){\rho}^{\alpha }+{\rho}^{\beta}\left({E}_{\rho }-{f}_{\beta}\right){\rho}^{\beta }+\xi\;\left[U\left({E}_U-{f}_U\right)U+\mathrm{h}.\mathrm{c}.\right], $$
(22.A9)

and E ρand E U are usual UHF-like energies for projectors ρ α, ρ β and U, U +, respectively. Moreover, E HPHF (i.e., E N needed for \( \Delta {E}_i^{+} \)) is known beforehand: E HPHF = (E ρ + ξ E U)/(1 + ξ). The eigenvalue problem for \( \left|{d}_a^{-}\right\rangle \)and \( \Delta {E}_a^{-} \) is formulated likewise. Namely, the relevant eigenvalue problem for \( \Delta {E}_a^{-} \) can be obtained from Eqs. (22.A5), (22.A6), (22.A7), (22.A8), and (22.A9) by replacing all projectors by their “vacant” counterparts (ρ α → I − ρ α,U → I − U etc.), but leaving all the Fockians, Eqs. (22.A5) and (22.A6), unchanged. At last, in order to get the resulting R HPHF from the eigensolutions of Eq. (22.A8) and their counterparts for \( \Delta {E}_a^{-} \), we directly apply Eq. (22.A2).

We now shortly discuss the selection rules for matrix R 0, i.e., for GF matrix elements at E = E F, neglecting energy broadening effects. The main rule is that for any correct bipartite-symmetry description we have the same block skew-diagonal structure of R 0 as in the underlying TB Hamiltonian, Eq. (22.5). Thus,

$$ {R}_0=\kern0.36em \left(\begin{array}{l}\;0\kern0.72em {R}_{\ast \circ}\\ {}\;{R}_{\circ \ast}\kern0.36em 0\end{array}\right). $$
(22.A10)

This equation for TB is trivial because \( {R}_0^{\mathrm{TB}}=-{\left({h}^0\right)}^{-1} \). Eq. (22.A10) is indeed the selection rule since it states that there are no nonzero elements of GF for (a,b) connections with a and b belonging simultaneously to the same atomic set, either the starred or unstarred set. Far less trivial is the fact that Eq. (22.A10) holds true for GF at the π-FCI level, as was stated rigorously in the important theorem obtained in Ref. [18]. Therefore, Eq. (22.A10) as originating from the bipartite symmetry, should be valid for any consistent π-approximation not violating a topological symmetry. The same selection is exactly fulfilled for QCTB [20], and it is not so difficult to prove the same rule at the HPHF level as well.

1.2 Appendix B: Approximate Versus “Exact” π-Electron Results for Small Aromatics

In order to estimate reliability of the results obtained by various approximate π-models, we consider briefly the formally exact π-electron theory based on the well-known FCI method (e.g., see [52]). In our computations, we will follow the previously proposed FCI matrix algorithm; for additional references see Ref. [53] where a suitable FCI approach to calculating Dyson orbitals is given. As to the MSE problems, the first important results at the π-FCI level were given only recently in Ref. [18]. In what follows, the FCI results we present here will be taken as the benchmark data against which all the others must be compared.

One special point concerns the actual Fermi energy E F that should be used to ensure Eq. (22.A10) for bipartites. In Ref. [18] the E F value is not given explicitly. At the same time, for bipartites the remarkable Hush and Pople theorem is valid at the π-electron Hartree-Fock level [54], as well as at the FCI level [55]. From this theorem it follows that E F = W C + γ C/2, where W C is the standard effective ionization potential, and γ C is the π-electron one-center Coulomb repulsion integral for the carbon atom. Just this choice of E F ensures the validity of Eq. (22.A10) and other properties of GF for bipartites.

In our specific π-electron computations, we use standard π-electron parameters (in eV): resonance integral of the aromatic π-bond β 0 =  − 2.4 ; W C = 0, γ C = 11.13, and two-center repulsion integrals due to Ohno. For QCTB computations, we adopt δ = 7/24 and E F = 0. The idealized regular geometry was taken for the carbon backbone in all studies of conjugated π-structures (1.4 Å for CC bond lengths, etc.).

Now, let us say few words about the supplementary rescaling of the GF matrix elements for RHF, HPHF, and FCI, following the procedure from Ref. [20]. This was proposed in order to avoid an inevitably large gap between different approaches. When multiplying RHF, HPHF, and FCI matrix elements of GF by the scaling factor β 0/(β 0 − γ 12/2) we make them comparable with their TB and QCTB counterparts. In particular, in the ethylene molecule the respective (1,2) elements, \( {\left({R}_0^{\mathrm{TB}}\right)}_{1,2} \) and \( {\left({R}_0^{\mathrm{RHF}}\right)}_{1,2} \) for the CC π bond, become identical and equal to 1.

Now we describe the results of comparison between π-FCI (the most rigorous π-approach) and main approximations (HPHF, RHF, QCTB, and TB). In addition, we tentatively and preliminary propose an improvement of QCTB in order to include long-range interactions not incorporated in the topological schemes. We simply do the first iteration of an usual self-consistent RHF procedure based on the TB (Hückel) density matrix as a start. It gives us a modified one-electron Hamiltonian of the correct block structure as in Eq. (22.5) for h 0. Then, expressions of the same type as in Eqs. (22.7), (22.8), (22.9), and (22.10) are applied in order to compute an approximated GF. This approach will be termed the quasi-correlated long-range interaction (QCLRI) model, and the respective GF will be denoted by G QCLRI. More detail will be given elsewhere.

Let us examine the numerical results presented in Table 22.3. The specific connections (,∘) are shown in Table 22.3 by stars and cycles. We see that HPHF provides the best (in respect to FCI) results whereas there are marked quantitative deviations of QCTB from FCI. Especially large deviations from FCI occur for TB. It is worth paying attention to a good quality of the RHF results for the considered small aromatic molecules. In fact, RHF provides here better results than TB and even QCTB. However, RHF calls for much more computational efforts, but more essential is that RHF is not appropriate for computing GF in extended π-systems (see Sect. 22.5). It is important for future applications to observe that QCLRI, i.e., the above-proposed simple π-scheme, surprisingly works almost as well as HPHF, at least for the considered molecules. It is noteworthy that, unlike QCTB, the QCLRI method possesses the size-consistency discussed in the last paragraph of Sect. 22.4).

Table 22.3 GF matrix elements R 0 (E = E F) for small aromatic molecules at the various levels of the theory

It is pertinent to understand now how significant in practice can be errors caused by lacking size-consistency in HPHF. A direct way to estimate actual inaccuracy due to the size inconsistency is to compute GF matrix elements in non-covalent intermolecular dimers of the chosen systems. Indeed, GF should be an additive-type size-consistent quantity (as closely related to the one-electron density matrix), and the same follows also from definition (22.A2). It means that the GF matrix of any noncovalent intermolecular dimer or complex, say, complex AB, must take the form of a direct sum when an average intermolecular distance goes to infinity. For example, in a dissociated dimer AB we have at the FCI level, R FCI[AB] = R FCI[A] ⊕ R FCI[B], and likewise for other size-consistent models, such as RHF, QCLRI, QCTB, and TB. Unfortunately, this is not the case of HPHF and related spin-projected Hartree-Fock models.

Let us examine the selected GF elements of the dimerized systems for the molecules studied in Table 22.3. For each dissociated dimer, its constituent monomeric parts A and B were situated at the intermolecular distance equal to 100 Å. Of course, FCI, QCLRI, RHF, QCTB, and TB obey the size-consistent requirement, so that the corresponding GF matrix elements in the initial monomer molecule and in the related parts of the dimer are exactly the same, and we do not repeat these data. At the same time, in the case of HPHF we obtain slightly different results for the monomer and the respective dimer subunits. We find the following HPHF values for GF elements under dissociation of the benzene, butalene, naphthalene, diphenylene, and naphtha[b]cyclobutadiene dimers:

$$ -0.475,-0.407,0.255,-0.142,-0.157 $$

These values should be compared with the respective values in the third column of Table 22.3. We see that in the dissociated dimers the deviation of GF elements from the ones obtained for the monomer are around of order 5%.

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Switzerland AG

About this paper

Check for updates. Verify currency and authenticity via CrossMark

Cite this paper

Luzanov, A.V. (2019). Single-Molecule Conductance Theory Using Different Orbitals for Different Spins: Applications to π-Electrons in Graphene Molecules. In: Fesenko, O., Yatsenko, L. (eds) Nanophotonics, Nanooptics, Nanobiotechnology, and Their Applications. NANO 2018. Springer Proceedings in Physics, vol 222. Springer, Cham. https://doi.org/10.1007/978-3-030-17755-3_22

Download citation

Publish with us

Policies and ethics