Skip to main content
Log in

Molecular orbital concept on spin-flip transport in molecular junctions

wave-packet scattering approach and Green’s function method

  • Regular Article
  • Published:
Theoretical Chemistry Accounts Aims and scope Submit manuscript

Abstract

The spin-dependent electron transport correlated with spin-flip dynamics in a molecular junction was investigated in the wave-packet and Green’s function approaches. The molecular junction adopted in this work is described by a simple one-dimensional tight-binding chain including a localized spin. The spin exchange coupling J between the localized and conduction electron spins was taken into account through the s-d Hamiltonian. The wave-packet simulations showed that the transmission probabilities in both the spin-flip and no-flip processes show large peaks at the eigenvalues of the spin singlet (−3J/4) and triplet (J/4) states, and that, different transmission properties appear at the mid-gap of the two eigenvalues: the spin-flip process shows a moderate decrease, whereas the no-flip process an abrupt drop. Dividing the s-d Hamiltonian into two submatrices and referring to the molecular orbital concept for the coherent electron transport, we found that the moderate decrease in the spin-flip process at the mid-gap is the result of a coherent-and-cooperative contribution from the singlet and triplet states of the conduction and localized electron spins, and that, the abrupt drop in the no-flip process at the mid-gap is mainly caused by the coherent cancellation from the singlet and triplet states. The molecular orbital concept available for the electron transport including spin-flip scattering processes is described in Green’s function method, in analogy to the one derived for the spinless electron transport.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Notes

  1. The accuracy of the calculations based on the many-body Green’s function depends on how the electron-electron interaction is reasonably truncated in self-energy through the Feynman diagrams. It requires technical skills and experience with Feynman diagrams. On the other hand, in the wave-packet simulations, the wave-packet propagates without any approximations, which is much simpler than in the many-body Green’s function, although whether the wave-packet simulation using the s-d Hamiltonian is applicable for the many-body effect is unclear.

  2. The matrix of the left (right) hand side electrode H L (H R) is written as a semi-infinite matrix in Eqs. 2 and 4. When the wave-packet simulation is carried out, the matrix size must be finite to obtain the propagation matrix, but in Green’s function method the matrix size is correctly semi-infinite, which is taken into account through the surface Green’s function technique. The finiteness of the matrices in the wave-packet simulation sometimes causes artificial errors in the calculations of transmission functions, but we simply avoid the problem by using a large matrix. The details are described in Appendix 3.

  3. The wave-packet amplitudes calculated in a spin unpolarized one-dimensional tight-binding chain were assigned to those only for up-spin sites in the left-electrode of the s-d model. This corresponds to the initial condition for the antiferromagnetic coupling between the incoming electron spin and localized spin.

  4. For the \(\uparrow\)-to-\(\downarrow\) case, we picked up the matrix elements in H s relating to the \(\uparrow\)-to-\(\downarrow\) process by neglecting the matrix element \(({\bf H}_{s}) s_{\uparrow, \Downarrow}, {_{s+1}}_{\uparrow} (= -t')\) and for the \(\uparrow\)-to-\(\uparrow\) case by neglecting the matrix element \(({\bf H}_{\bf s})s_{\downarrow, \Uparrow}, {_{s+1}}_{\downarrow} (= -t').\)

  5. When the molecule is in an isolated situation, we can use the spin-orbit interaction for the calculations of J. However, the molecular junction in which the spin-flip is caused by the incoming electrons from an electrode is clearly different from an isolated molecule in which an electron leading to the spin-flip is stationary captured by the molecule. That is, even when we adopted the spin-orbit interaction, the applicability of the strategy is still ambiguous for the present target. Thus we adopted the s-d Hamiltonian as the first step in this study because the Hamiltonian is clearly constructed from the conduction s-electrons and localized spin on the d-level. To investigate the parameter dependence on the spin-flip transport, we simulated the spin-flip dynamics using a strong/weak spin-spin interaction J with a strong/weak electrode-molecule coupling t′, as shown in Fig. 8.

References

  1. Kane BE (1998) A silicon-based nuclear spin quantum computer. Nature 393:133

    Article  CAS  Google Scholar 

  2. Vandersypen LMK, Steffen M, Breyta G, Yannoni CS, Sherwood MH, Chuang IL (2001) Experimental realization of Shor’s quantum factoring algorithm using nuclear magnetic resonance. Nature 414:883

    Article  CAS  Google Scholar 

  3. Jelezko F, Gaebel T, Popa I, Gruber A, Wrachtrup J (2004) Observation of coherent oscillations in a single electron spin. Phys Rev Lett 92:076401

    Article  CAS  Google Scholar 

  4. Xiao M, Martin I, Yablonovitch E, Jiang HW (2004) Electrical detection of the spin resonance of a single electron in a silicon field-effect transistor. Nature 430:435

    Article  CAS  Google Scholar 

  5. Childress L, Dutt MVG, Taylor JM, Zibrov AS, Jelezko F, Wrachtrup J, Hemmer PR, Lukin MD (2006) Coherent dynamics of coupled electron and nuclear spin qubits in diamond. Science 314:281

    Article  CAS  Google Scholar 

  6. Hanson R, Mendoza FM, Epstein RJ, Awschalom DD (2006) Polarization and readout of coupled single spins in diamond. Phys Rev Lett 97:087601

    Article  CAS  Google Scholar 

  7. McCamey DR, Huebl H, Brandt MS, Hutchison WD, McCallum JC, Clark RG, Hamilton AR (2006) Electrically detected magnetic resonance in ion-implanted Si : P nanostructures. Appl Phys Lett 89:182115

    Article  Google Scholar 

  8. Morton JJL, Tyryshkin AM, Brown RM, Shankar S, Lovett BW, Ardavan A, Schenkel T, Haller EE, Ager JW, Lyon SA (2008) Solid-state quantum memory using the 31P nuclear spin. Nature 455:1085

    Article  CAS  Google Scholar 

  9. Schlegel C, van Slageren J, Manoli M, Brechin EK, Dressel M (2008) Direct observation of quantum coherence in single-molecule magnets. Phys Rev Lett 101:147203

    Article  CAS  Google Scholar 

  10. Tada T (2008) Hyperfine switching triggered by resonant tunneling for the detection of a single nuclear spin qubit. Phys Lett A 372:6690

    Article  CAS  Google Scholar 

  11. Smeltzer B, McIntyre J, Childress L (2009) Robust control of individual nuclear spins in diamond. Phys Rev A 80:050302

    Article  Google Scholar 

  12. Loth S, von Bergmann K, Ternes M, Otte AF, Lutz CP, Heinrich AJ (2010) Controlling the state of quantum spins with electric currents. Nat Phys 6:340

    Article  CAS  Google Scholar 

  13. Wang X, Bayat A, Schirmer SG, Bose S (2010) Robust entanglement in antiferromagnetic Heisenberg chains by single-spin optimal control. Phys Rev A 81:032312

    Article  Google Scholar 

  14. Park J, Pasupathy AN, Goldsmith JI, Chang C, Yaish Y, Petta JR, Rinkoski M, Sethna JP, Abruña HD, McEuen PL, Ralph DC (2002) Coulomb blockade and the Kondo effect in single-atom transistors. Nature 417:722

    Article  CAS  Google Scholar 

  15. Liang W, Shores MP, Bockrath M, Long JR, Park H (2002) Kondo resonance in a single-molecule transistor. Nature 417:725

    Article  CAS  Google Scholar 

  16. Heinrich AJ, Gupta JA, Lutz CP, Eigler DM (2004) Single-atom spin-flip spectroscopy. Science 306:466

    Article  CAS  Google Scholar 

  17. Wahl P, Diekhöner L, Wittich G, Vitali L, Schneider MA, Kern K (2005) Kondo effect of molecular complexes at surfaces: ligand control of the local spin coupling. Phys Rev Lett 95:166601

    Article  CAS  Google Scholar 

  18. Zhao A, Li Q, Chen L, Xiang H, Wang W, Pan S, Wang B, Xiao X, Yang J, Hou JG, Zhu Q (2005) Controlling the kondo effect of an adsorbed magnetic ion through its chemical bonding. Science 309:1542

    Article  CAS  Google Scholar 

  19. Iancu V, Deshpande A, Hla S-W (2006) Manipulating kondo temperature via single molecule switching. Nano Lett 6:820

    Article  CAS  Google Scholar 

  20. Parks JJ, Champagne AR, Hutchison GR, Flores-Torres S, Abruña HD, Ralph DC (2007) Tuning the kondo effect with a mechanically controllable break junction. Phys Rev Lett 99:026601

    Article  CAS  Google Scholar 

  21. Gao L, Ji W, Hu YB, Cheng ZH, Deng ZT, Liu Q, Jiang N, Lin X, Guo W, Du SX, Hofer WA, Xie XC, Gao H-J (2007) Site-specific kondo effect at ambient temperatures in Iron-based molecules. Phys Rev Lett 99:106402

    Article  CAS  Google Scholar 

  22. Chen X, Fu Y-S, Ji S-H, Zhang T, Cheng P, Ma X-C, Zou X-L, Duan W-H, Jia J-F, Xue Q-K (2008) Probing superexchange interaction in molecular magnets by spin-flip spectroscopy and microscopy. Phys Rev Lett 101:197208

    Article  Google Scholar 

  23. Sugawara T, Minamoto M, Matsushita MM, Nickels P, Komiyama S (2008) Cotunneling current affected by spin-polarized wire molecules in networked gold nanoparticles. Phys Rev B 77:235316

    Article  Google Scholar 

  24. Tsukahara N, Noto K, Ohara M, Shiraki S, Takagi N, Takata Y, Miyawaki J, Taguchi M, Chainani A, Shin S, Kawai M (2009) Adsorption-induced switching of magnetic anisotropy in a single Iron(II) phthalocyanine molecule on an oxidized Cu(110) surface. Phys Rev Lett 102:167203

    Article  Google Scholar 

  25. Xue Y, Ratner MA (2003) Microscopic study of electrical transport through individual molecules with metallic contacts. II. Effect of the interface structure. Phys Rev B 68:115407

    Article  Google Scholar 

  26. Stokbro K, Taylor J, Brandbyge M, MozosJ-L Ordejón P (2003) Theoretical study of the nonlinear conductance of Di-thiol benzene coupled to Au(111) surfaces via thiol and thiolate bonds. Comput Mater Sci 27:151

    Article  CAS  Google Scholar 

  27. Nara J, Kino H, Kobayashi N, Tsukada M, Ohno T (2003) Theoretical investigation of contact effects in conductance of single organic molecule. Thin Solid Films 438(439):221

    Article  Google Scholar 

  28. Hu Y, Zhu Y, Gao H, Guo H (2005) Conductance of an ensemble of molecular wires: a statistical analysis. Phys Rev Lett 95:156803

    Article  Google Scholar 

  29. Ke S-H, Baranger HU, Yang W (2005) Contact atomic structure and electron transport through molecules. J Chem Phys 122:074704

    Article  Google Scholar 

  30. Tanibayashi S, Tada T, Watanabe S, Sekino H (2006) Effects of energetic stability in transport measurements of single benzene-dithiolate by the STM break junction technique. Chem Phys Lett 428:367

    Article  CAS  Google Scholar 

  31. Andrews DQ, Van Duyne RP, Ratner MA (2008) Stochastic modulation in molecular electronic transport junctions: molecular dynamics coupled with charge transport calculations. Nano Lett 8:1120

    Article  CAS  Google Scholar 

  32. Tawara A, Tada T, Watanabe S (2009) Electrostatic and dynamical effects of an aqueous solution on the zero-bias conductance of a single molecule: a first-principles study. Phys Rev B 80:073409

    Article  Google Scholar 

  33. Monturet S, Lorente N (2008) Inelastic effects in electron transport studied with wave-packet propagation. Phys Rev B 78:035445

    Article  Google Scholar 

  34. Ishii H, Kobayashi N, Hirose K (2010) Order-N electron transport calculations from ballistic to diffusive regimes by a time-dependent wave-packet diffusion method: application to transport properties of carbon nanotubes. Phys Rev B 82:085435

    Article  Google Scholar 

  35. Troisi A, Orlandi G (2006) Charge-transport regime of crystalline organic semiconductors: diffusion limited by thermal off-diagonal electronic disorder. Phys Rev Lett 96:086601

    Article  Google Scholar 

  36. Kondo N, Yamamoto T, Watanabe K (2006) Phonon wavepacket scattering dynamics in defective carbon nanotubes. Jpn J Appl Phys 45:L963

    Article  CAS  Google Scholar 

  37. Yao Y, Zhao H, Moore JE, Wu C-Q (2008) Controllable spin-current blockade in a Hubbard chain. Phys Rev B 78:193105

    Article  Google Scholar 

  38. Bruus H, Flensberg K (2004) Many-body quantum theory in condensed matter physics, chapter 10. Oxford University Press, Oxford

    Google Scholar 

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

    Article  CAS  Google Scholar 

  40. Kondo J (1964) Resistance minimum in dilute magnetic alloys. Prog Theor Phys 32:37

    Article  CAS  Google Scholar 

  41. Press WH, Teukolsky SA, Vetterling WT, Flannery BP (1992) Numerical recipes in fortran 77, 2nd edn. Cambridge University Press, Cambridge

    Google Scholar 

  42. Schiff LI (1968) Quantum mechanics, 3rd edn. McGraw-Hill, New York, pp 62–64

    Google Scholar 

  43. Caroli C, Combescot R, Nozieres P, Saint-James D (1971) Direct calculation of the tunneling current. J Phys C 4:916

    Article  Google Scholar 

  44. Datta S (1995) Electronic transport in mesoscopic systems, chapters 2 and 3. Cambridge University Press, Cambridge

    Google Scholar 

  45. Tada T, Yoshizawa K (2003) Quantum transport effects in nanosized graphite sheets. II. Enhanced quantum transport effects by Heteroatoms. J Phys Chem B 107:8789

    Article  CAS  Google Scholar 

  46. Tada T, Kondo M, Yoshizawa K (2003) Theoretical measurements of conductance in an (AT)12 DNA molecule. ChemPhysChem 4:1256

    Article  CAS  Google Scholar 

  47. Tada T, Yoshizawa K (2004) Reverse exponential decay of electrical transmission in nanosized graphite sheets. J Phys Chem B 108:7565

    Article  CAS  Google Scholar 

  48. Kondo M, Tada T, Yoshizawa K (2004) Wire-length dependence of the conductance of oligo(p-phenylene) dithiolate wires: a consideration from molecular orbitals. J Phys Chem A 108:9143

    Article  CAS  Google Scholar 

  49. Tada T, Nozaki D, Kondo M, Hamayama S, Yoshizawa K (2004) Oscillation of conductance in molecular junctions of carbon ladder compounds. J Am Chem Soc 126:14182

    Article  CAS  Google Scholar 

  50. Tada T, Hamayama S, Kondo M, Yoshizawa K (2005) Quantum transport effects in copper(II) phthalocyanine sandwiched between gold nanoelectrodes. J Phys Chem B 109:12443

    Article  CAS  Google Scholar 

  51. Kondo M, Tada T, Yoshizawa K (2005) A theoretical measurement of the quantum transport through an optical molecular switch. Chem Phys Lett 412:55

    Article  CAS  Google Scholar 

  52. Yoshizawa K, Tada T, Staykov A (2008) Orbital views of the electron transport in molecular devices. J Am Chem Soc 130:9406

    Article  CAS  Google Scholar 

  53. Shiba H (1996) Kotai no denshiron, chapter 4. Maruzen, Tokyo

Download references

Acknowledgments

It is a great pleasure to dedicate this paper to Prof. A. Imamura, who was a supervisor of one of the authors (TT) in Hiroshima University from 1995 to 1998. The lectures on the molecular orbital theory given by A. Imamura have continued to inspire the interests of one of the authors (TT) in the field of the molecular science. This work was partially supported by the Grant-in-Aid for Young Scientists (B), MEXT of Japan. The author would like to thank S. Konabe for technical information on wave-packet dynamics.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tomofumi Tada.

Additional information

Dedicated to Professor Akira Imamura on the occasion of his 77th birthday and published as part of the Imamura Festschrift Issue.

Appendices

Appendix 1: s-d Hamiltonian

The one-dimensional system adopted in this study is composed of electrodes and a localized electron spin. This system, thus, can be regarded as a metal including a single impurity spin. The Anderson Hamiltonian is an appropriate model to investigate the magnetic property of the metal with impurities, and we use the Anderson Hamiltonian to derive the s-d Hamiltonian, which is the story introduced in the book by Shiba [53]. The Anderson Hamiltonian is written as

$$ H = \sum_{k \sigma} \varepsilon_{k} c^{\dag}_{k \sigma} c_{k \sigma} + \sum_{\sigma} \varepsilon_{d} n_{d \sigma} + U n_{d \uparrow} n_{d \downarrow} + \frac{1}{\sqrt {N_A}} \sum_{k \sigma}\left(V_{k} c^{\dag}_{k \sigma} d_{\sigma} + \hbox{h.c.}\right) . $$
(23)

The first term represents the Hamiltonian of conduction electrons in electrodes, and \(\varepsilon_{d}\) in the second term is the impurity d-level. The on-site energy U in the third term represents the coulomb interaction when the impurity d-level is occupied by two electrons. The forth term is the interaction between the orbitals in electrodes and the impurity. \(c^{\dag}_{\sigma}\) and \({d^{\dag}_{\sigma}} (c_{\sigma}\) and d σ) are the creation (annihilation) operators of an electron with spin σ in the electrodes and impurity level, respectively; n dσ is the number operator, \(d^{\dag}_{\sigma} d_{\sigma},\) and N A is the total number of atoms. To investigate physical properties of the system described with the Anderson Hamiltonian, there are two standard approaches: (1) the third term in Eq. 23 is considered as a perturbation, and (2) the fourth term in Eq. 23 is considered as a perturbation. The latter approach is more suitable for the present one-dimensional model including a single localized spin, that is, the unperturbed Hamiltonian H 0 and perturbation term H I are

$$ H_{0} = \sum_{k \sigma} \varepsilon_{k} c^{\dag}_{k \sigma} c_{k \sigma} + \sum_{\sigma} \varepsilon_{d} n_{d \sigma} + U n_{d \uparrow} n_{d \downarrow} , $$
(24)

and

$$ H_{I} = \frac{1}{\sqrt {N_A}} \sum_{k \sigma}\left(V_{k} c^{\dag}_{k \sigma} d_{\sigma} + \hbox{h.c.} \right) , $$
(25)

respectively. The unperturbed Green’s function G 0(E) and perturbed Green’s function G(E) can be represented as (E − H 0)−1 and (E − H)−1, respectively. Using the relation \( (A-B)^{-1} = A^{-1} + A^{-1}B A^{-1} + A^{-1}B A^{-1}B A^{-1} + \cdots , \) we straightforwardly obtain the perturbed Green’s function in terms of G 0 and H I as

$$ G = G_0 + G_0 H_I G_0 + G_0 H_I G_0 H_I G_0 + G_0 H_I G_0 H_I G_0 H_I G_0 + \cdots. $$
(26)

This is useful to define the effective Hamiltonian of the perturbation term.

Let us firstly consider the system in which the d-level is occupied by an electron and investigate the effects from the interactions between the d-level and conduction levels. The unperturbed states are represented as \( d^{\dag}_{\uparrow} | F \rangle \) and \( d^{\dag}_{\downarrow} | F \rangle,\) where \( | F \rangle\) indicates conduction electrons occupying the levels \( \varepsilon_{k}\) up to the Fermi level. When we consider the first-order process H I (i.e., \(V_{k} c^{\dag}_{k \sigma} d_{\sigma})\) with respect to the unperturbed states \( d^{\dag}_{\uparrow} | F \rangle \) and \( d^{\dag}_{\downarrow} | F \rangle , \) we can easily find that the expectation values of the first-order process are vanished because of the following relations: \( d_{\sigma} | F \rangle = 0\) and \( \langle F | d^{\dag}_{\sigma}= 0. \) For example, the first-order term for \( d^{\dag}_{\uparrow} | F \rangle \) has the form of \( \langle F | d_{\uparrow} c^{\dag}_{k \sigma} d_{\sigma} d^{\dag}_{\uparrow} | F \rangle \) or \( \langle F | d_{\uparrow} d^{\dag}_{\sigma} c_{k \sigma} d^{\dag}_{\uparrow} | F \rangle . \) The observation allows us to rewrite the perturbed Green’s function as

$$ G = G_0 + G_0 H_I G_0 H_I G_0 + G_0 H_I G_0 H_I G_0 H_I G_0 H_I G_0 + \cdots. $$
(27)

Defining the second-order term H I G 0 H I as H eff I , we obtain

$$ G = G_0 + G_0 H^{\rm eff}_I G_0 + G_0 H^{\rm eff}_I G_0 H^{\rm eff}_I G_0 + \cdots. $$
(28)

This is the perturbed Green’s function for the Hamiltonian of H 0 + H eff I . Thus, we consider the second-order processes H I G 0 H I to obtain the explicit form of the effective Hamiltonian. It is to be noted that the unperturbed Green’s function G 0 is a function of energy E, and the energy dependence in the effective Hamiltonian H I G 0 H I is not convenient for general use. However, the perturbed states in which the energy is close to the unperturbed energy \(E_0 (= \sum\nolimits_{k'}^{\rm occ} \varepsilon_{k'} + \varepsilon_d \)) are well mixed with the unperturbed state, and the effective Hamiltonian derived from H I G 0(E 0) H I can be recognized as a reasonable one.

Let us next consider the explicit form of the second-order process. There are two types for the intermediate states in which zero/two electrons occupy the d-level. For the zero occupation case with the initial state of \( d^{\dag}_{\uparrow} | F \rangle , \) we have

$$ \frac{1}{E_0 - H_0} V c^{\dag}_{k \uparrow} d_{\uparrow} d^{\dag}_{\uparrow} | F \rangle = \frac{1}{\varepsilon_d - \varepsilon_k } V c^{\dag}_{k \uparrow} | F \rangle, $$
(29)

where we use the anticommutator relation of Fermions (e.g., \([c_{k}, c^{\dag}_{k'}]_{+} = \delta_{k,k'}\)) and the fact \( H_0 c^{\dag}_{k \uparrow} d_{\uparrow} d^{\dag}_{\uparrow} | F \rangle = (\sum\nolimits_{k'}^{\rm occ} \varepsilon_{k'} + \varepsilon_k ) c^{\dag}_{k \uparrow} d_{\uparrow} d^{\dag}_{\uparrow} | F \rangle . \) Operating H I at the left-hand side of Eq. 29, we obtain

$$ \begin{aligned} & V \left(d^{\dag}_{\uparrow} c_{k" \uparrow} + d^{\dag}_{\downarrow} c_{k" \downarrow}\right) \frac{1}{\varepsilon_d - \varepsilon_k } V c^{\dag}_{k \uparrow} | F \rangle \\ &= \frac{V^2}{\varepsilon_d - \varepsilon_k } \left(d^{\dag}_{\uparrow} c_{k" \uparrow}c^{\dag}_{k \uparrow} + d^{\dag}_{\downarrow} c_{k" \downarrow}c^{\dag}_{k \uparrow}\right) | F \rangle \\ &= \frac{V^2}{\varepsilon_d - \varepsilon_k } \left(\delta_{k,k"}d^{\dag}_{\uparrow} -c^{\dag}_{k \uparrow}c_{k" \uparrow} d^{\dag}_{\uparrow} - c^{\dag}_{k \uparrow} c_{k" \downarrow}d^{\dag}_{\downarrow} \right) | F \rangle . \end{aligned} $$
(30)

Since the expectation value of the process is obtained by applying \( \langle F | d_{\sigma} \) at the left-hand side of Eq. 30, we readily confirm the second-order processes have non-zero expectation values.

For the initial state of \( d^{\dag}_{\downarrow} | F \rangle \) in the zero occupation case, we have

$$ \frac{1}{E_0 - H_0} V c^{\dag}_{k \downarrow} d_{\downarrow} d^{\dag}_{\downarrow} | F \rangle = \frac{1}{\varepsilon_d - \varepsilon_k } V c^{\dag}_{k \downarrow} | F \rangle, $$
(31)

and obtain the following expression by operating H I at the left-hand side of Eq. 31 as

$$ \begin{aligned} & V \left(d^{\dag}_{\uparrow} c_{k" \uparrow} + d^{\dag}_{\downarrow} c_{k" \downarrow}\right) \frac{1}{\varepsilon_d - \varepsilon_k } V c^{\dag}_{k \downarrow} | F \rangle \\ &= \frac{V^2}{\varepsilon_d - \varepsilon_k } \left(d^{\dag}_{\uparrow} c_{k" \uparrow}c^{\dag}_{k \downarrow} + d^{\dag}_{\downarrow} c_{k" \downarrow}c^{\dag}_{k \downarrow}\right) | F \rangle \\ &= \frac{V^2}{\varepsilon_d - \varepsilon_k } \left(\delta_{k,k"}d^{\dag}_{\downarrow} -c^{\dag}_{k \downarrow}c_{k" \uparrow}d^{\dag}_{\uparrow} - c^{\dag}_{k \downarrow} c_{k" \downarrow}d^{\dag}_{\downarrow} \right) | F \rangle . \end{aligned} $$
(32)

For the doubly occupation case with the initial states of \( d^{\dag}_{\uparrow} | F \rangle \) and \( d^{\dag}_{\downarrow} | F \rangle,\) we, respectively, have

$$ \begin{gathered} V\left( {c_{{k^{\prime } \uparrow }}^{\dag } d_{ \uparrow } + c_{{k^{\prime } \downarrow }}^{\dag } d_{ \downarrow } } \right)\frac{1}{{\varepsilon _{k} - \varepsilon _{d} - U}}Vd_{ \downarrow }^{\dag } c_{{k \downarrow }} d_{ \uparrow }^{\dag } |F\rangle \hfill \\ = \frac{{V^{2} }}{{\varepsilon _{k} - \varepsilon _{d} - U}}\left( { - c_{{k^{\prime } \uparrow }}^{\dag } c_{{k \downarrow }} d_{ \downarrow }^{\dag } + c_{{k^{\prime } \downarrow }}^{\dag } c_{{k \downarrow }} d_{ \uparrow }^{\dag } } \right)|F\rangle , \hfill \\ \end{gathered} $$
(33)

and

$$ \begin{aligned} & V \left(c^{\dag}_{k" \uparrow}d_{\uparrow} + c^{\dag}_{k" \downarrow}d_{\downarrow} \right) \frac{1}{\varepsilon_k - \varepsilon_d - U} V d^{\dag}_{\uparrow} c_{k \uparrow}d^{\dag}_{\downarrow} | F \rangle \\ &= \frac{V^2}{\varepsilon_k - \varepsilon_d - U } \left( -c^{\dag}_{k" \downarrow}c_{k \uparrow} d^{\dag}_{\uparrow} + c^{\dag}_{k" \uparrow} c_{k \uparrow}d^{\dag}_{\downarrow} \right) | F \rangle. \end{aligned} $$
(34)

From Eqs. 30, 3234, we obtain the expression of the effective Hamiltonian with respect to \(| F \rangle\) as

$$ \begin{aligned} &\frac{V^2}{\varepsilon_d - \varepsilon_k } \left( \delta_{k,k"}d^{\dag}_{\uparrow} -c^{\dag}_{k \uparrow}c_{k" \uparrow} d^{\dag}_{\uparrow} - c^{\dag}_{k \uparrow} c_{k" \downarrow}d^{\dag}_{\downarrow} \right.\\ &\left.+\,\delta_{k,k"}d^{\dag}_{\downarrow} - c^{\dag}_{k \downarrow} c_{k" \downarrow}d^{\dag}_{\downarrow}-c^{\dag}_{k \downarrow}c_{k" \uparrow}d^{\dag}_{\uparrow} \right) \\ &+\,\frac{V^2}{\varepsilon_k - \varepsilon_d - U } \left( -c^{\dag}_{k" \uparrow}c_{k \downarrow} d^{\dag}_{\downarrow} + c^{\dag}_{k" \downarrow} c_{k \downarrow}d^{\dag}_{\uparrow} \right.\\ &\left.-\,c^{\dag}_{k" \downarrow}c_{k \uparrow} d^{\dag}_{\uparrow} + c^{\dag}_{k" \uparrow} c_{k \uparrow}d^{\dag}_{\downarrow} \right) . \end{aligned} $$
(35)

Since this Hamiltonian is derived using the singly occupied d-level (i.e., \( d^{\dag}_{\uparrow} | F \rangle \) and \( d^{\dag}_{\downarrow} | F \rangle \)), the creation operators for the d-level must be replaced with the number operators for general use. The effective Hamiltonian, thus, can be represented as

$$ \begin{aligned} H^{\rm eff}_I &= \frac{1}{N_A}\sum_{k,k"} \left[ \frac{V^2}{\varepsilon_d - \varepsilon_k } \left(-c^{\dag}_{k \uparrow}c_{k" \uparrow} d^{\dag}_{\uparrow}d_{\uparrow} - c^{\dag}_{k \uparrow} c_{k"\downarrow}d^{\dag}_{\downarrow}d_{\uparrow} \right.\right.\\ &\left.\left.-\,c^{\dag}_{k \downarrow} c_{k" \downarrow}d^{\dag}_{\downarrow}d_{\downarrow} -c^{\dag}_{k \downarrow}c_{k" \uparrow}d^{\dag}_{\uparrow}d^{\dag}_{\downarrow} \right)\right.\\ &\left.+\,\frac{V^2}{\varepsilon_k - \varepsilon_d - U } \left( -c^{\dag}_{k" \uparrow}c_{k \downarrow} d^{\dag}_{\downarrow} d_{\uparrow} + c^{\dag}_{k" \downarrow} c_{k \downarrow}d^{\dag}_{\uparrow} d_{\uparrow} \right.\right.\\ &\left.\left.-\,c^{\dag}_{k" \downarrow}c_{k \uparrow}d^{\dag}_{\uparrow}d_{\downarrow} + c^{\dag}_{k" \uparrow} c_{k \uparrow}d^{\dag}_{\downarrow}d_{\downarrow} \right) \right], \end{aligned} $$
(36)

where we omitted several terms that do not include the operators of conduction electrons for simplicity. Using the relations for the number counting and spin-flip operators, \(n_{d \sigma} = d^{\dag}_{\sigma}d_{\sigma}, S_{+} = d^{\dag}_{\uparrow}d_{\downarrow}, \) and \(S_{-} = d^{\dag}_{\downarrow}d_{\uparrow}, \) we have

$$ \begin{aligned} H^{\rm eff}_I &= \frac{1}{N_A}\sum_{k,k"} \left[ \frac{V^2}{\varepsilon_d - \varepsilon_k } \left(-c^{\dag}_{k \uparrow}c_{k" \uparrow} n_{d \uparrow} - c^{\dag}_{k \uparrow} c_{k" \downarrow}S_{-} \right.\right.\\ &\left.\left.-\,c^{\dag}_{k \downarrow} c_{k" \downarrow} n_{d \downarrow} -c^{\dag}_{k \downarrow}c_{k" \uparrow}S_{+} \right)\right. \\ &\left.+\,\frac{V^2}{\varepsilon_k - \varepsilon_d - U } \left( -c^{\dag}_{k" \uparrow}c_{k \downarrow} S_{-} + c^{\dag}_{k" \downarrow} c_{k \downarrow} n_{d \uparrow} \right.\right.\\ &\left.\left.-\,c^{\dag}_{k" \downarrow}c_{k \uparrow} S_{+} + c^{\dag}_{k" \uparrow} c_{k \uparrow} n_{d \downarrow} \right) \right]. \end{aligned} $$
(37)

Since the energy difference between \(\varepsilon_{k}\) and the Fermi level is smaller than \(\varepsilon_{d}\) and \(\varepsilon_{d} + U, \) the effective Hamiltonian can be written as

$$ \begin{aligned} H^{\rm eff}_I &= \frac{1}{N_A}\sum_{k,k"} \left[ V^2 \left(-\frac{1}{\varepsilon_d} - \frac{1}{\varepsilon_d + U}\right) \times \left(c^{\dag}_{k \uparrow}c_{k" \uparrow} + c^{\dag}_{k \downarrow}c_{k" \downarrow} \right)( n_{d \uparrow} + n_{d \downarrow} ) \right.\\ &\left.+ V^2 \left(-\frac{1}{\varepsilon_d} + \frac{1}{\varepsilon_d + U}\right) \left(c^{\dag}_{k \uparrow}c_{k" \downarrow} S_{-} + c^{\dag}_{k \downarrow}c_{k" \uparrow} S_{+} \right) \right]. \end{aligned} $$
(38)

Using the condition n d + n d = 1, we reasonably obtain the effective Hamiltonian as

$$ \begin{aligned} H_{I}^{{{\text{eff}}}} & = \frac{1}{{N_{A} }}\sum\limits_{{k,k^{\prime } ,\sigma }} {\frac{{V^{2} }}{2}} \left( { - \frac{1}{{\varepsilon _{d} }} - \frac{1}{{\varepsilon _{d} + U}}} \right)c_{{k\sigma }}^{\dag } c_{{k^{\prime } \sigma }} \\ & \quad + \frac{1}{{N_{A} }}\sum\limits_{{k,k^{\prime } ,\sigma ,\sigma ^{\prime } }} {V^{2} } {\text{ }}\left( { - \frac{1}{{\varepsilon _{d} }} + \frac{1}{{\varepsilon _{d} + U}}} \right)c_{{k\sigma }}^{\dag } {\text{ }}\sigma _{{\sigma ,\sigma ^{\prime } }} c_{{k^{\prime } \sigma ^{\prime } }} \cdot {\mathbf{S}}, \\ \end{aligned} $$
(39)

where \(\varvec{\sigma}\) is the Pauli matrix,

$$ \varvec{\sigma}_{\bf x}= \bordermatrix{ & \uparrow & \downarrow \cr \uparrow & 0 & 1 \cr \downarrow & 1 & 0 \cr }, \varvec{\sigma}_{\rm y } = \bordermatrix{ & \uparrow & \downarrow \cr \uparrow & 0 & -i \cr \downarrow & i & 0 \cr}, \varvec{\sigma}_{\bf z} = \bordermatrix{ & \uparrow & \downarrow \cr \uparrow & 1 & 0 \cr \downarrow & 0 & -1 \cr}, $$
(40)

and we use the following relations, S + = S x  + i S y and S  = S x  − i S y . The first term represents the potential scattering process that are independent of the spin-direction. The second term represents the spin-dependent scattering processes including spin-flip processes. Regarding the term \(- 2 V^2 \left(-\frac{1}{\varepsilon_d} + \frac{1}{\varepsilon_d + U}\right) \) as the spin-spin interaction J, we finally obtain the s-d Hamiltonian as

$$ H_{s-d} = -\frac{J}{2N_A}\sum_{k,k",\sigma, \sigma'} c^{\dag}_{k \sigma} \varvec{\sigma}_{\sigma,\sigma'} c_{k" \sigma'} \cdot {\bf S}. $$
(41)

From the s-d Hamiltonian, we easily obtain the matrix elements in the spin sub-space as follows:

$$ {\bf H}_{\bf sd} = \bordermatrix{& \uparrow \Uparrow & \downarrow \Downarrow & \uparrow \Downarrow & \downarrow \Uparrow \cr \uparrow \Uparrow & \frac{J}{4}& 0 & 0 & 0 \cr \downarrow \Downarrow & 0 & \frac{J}{4} & 0 & 0 \cr \uparrow \Downarrow & 0& 0 & -\frac{J}{4} & \frac{J}{2} \cr \downarrow \Uparrow & 0 & 0 & \frac{J}{2} & -\frac{J}{4} \cr}. $$
(42)

It is to be noted that the factor N A is taken into account through the normalized amplitudes in the wave packets.

Appendix 2: The Crank-Nicholson scheme

The time-dependent Schrödinger equation is

$$ i \frac{\partial \psi}{\partial t} = H \psi , $$
(43)

and the formal solution of the equation is

$$ \psi (x,t) = e^{-iHt}\psi (x,0). $$
(44)

When we introduce the finite grid for time with the interval of \(\Updelta t, \) the wave function after the time propagation of \(\Updelta t\) can be written as

$$ \psi (x,t + \Updelta t ) = ( 1 - i H \Updelta t) \psi (x,t) . $$
(45)

This is accurate up to the first order in time. However, it is well known that the strategy in Eq. 45 is numerically unstable and the time propagation operator is not unitary. To avoid these difficulties, we consider the time propagation in the reverse direction, t + \(\Updelta t \rightarrow t . \) In this case, we have the following relation:

$$ \psi (x, t) = ( 1 + i H \Updelta t) \psi (x,t + \Updelta t) . $$
(46)

Using Eqs. 45 and 46, the wave function at \(t + \Updelta t / 2 \) is written as

$$ \psi (x,t + \Updelta t / 2 ) = ( 1 - i H \Updelta t / 2) \psi (x,t) $$
(47)

and

$$ \psi (x,t + \Updelta t / 2) = ( 1 + i H \Updelta t / 2) \psi (x, t+ \Updelta t) . $$
(48)

Eliminating the term \( \psi (x,t + \Updelta t / 2) \) using Eqs. 47 and 48, we obtain

$$ (1 + i H \Updelta t / 2) \psi (x, t+ \Updelta t) = ( 1 - i H \Updelta t / 2) \psi (x,t), $$
(49)

and thereby

$$ \psi (x, t+ \Updelta t) = \frac{( 1 - i H \Updelta t / 2)}{ ( 1 + i H \Updelta t / 2)} \psi (x,t). $$
(50)

We can easily confirm that the time propagation operator \( \frac{( 1 - i H \Updelta t / 2)}{ ( 1 + i H \Updelta t / 2)} \) is unitary and that the time propagation is accurate up to the second order in time using the standard mathematical relations: \(e^{A} = 1 + A + \frac{1}{2!} A^2 + \cdots\) and \( (1+A)^{-1} = 1 - A + A^2 -A^3 \cdots. \)

Appendix 3: The computational conditions for wave-packet simulations

There are two important computational conditions for wave-packet simulations: (1) the energy resolution in the wave-packet simulation must be quite fine in order to compare the results with those from Green’s function method and (2) we have to take care that an artificial reflection of the wave-packet at the left-/right-hand edge of the one-dimensional chain does not affect the true dynamics of the wave-packet scattered by the localized spin. As for the first point, the Gaussian type of wave-packet is adopted in this study, and thereby, the wave-packet is spatially well broadened when the energy of the wave-packet is finely focused on a certain value. This is very important for the calculations of the transmission functions from wave-packet dynamics as a function of energy. We tested several Gaussian wave-packets having different broadening width and consequently found that the wave-packet with broadening width of 1,200 sites or more is sufficient for the calculations of transmission functions and for the comparison with Green’s function results. As for the second point, we have to recognize that the firstly reflected wave-packet at the central spin site will arrive at the left-hand edge of the one-dimensional chain and in turn propagates again toward the spin site. If a portion of the wave-packet is trapped on the spin site (this will be probable for the weak interaction between the electrode and molecule), an artificial superposition between the trapped wave-packet and re-reflected wave-packet occurs. Since we calculate the transmission probabilities from the transmitted wave-packet at the right electrodes, the artificial superposition will cause the artificial errors in transmission functions. We of course take care about the reflection of the transmitted wave-packet at the right-hand edge, too. To avoid the artificial errors caused by the re-reflected wave-packets at the left-/right-hand edge, we simply adopted a large number of sites, 20,000 in this study, and confirmed the model size is sufficient for the present purpose.

As for the unit of simulation time, we determined the unit as follows: (1) we first assumed that each site in electrodes corresponds to a carbon atom, and second that the transfer integral t of 1 corresponds to 2.7 eV, which is a typical value in the carbon 2 p π network; (2) we calculated the Fermi velocity from the energy band dispersion in the one-dimensional carbon chain, using the carbon--carbon bond distance of 1.4 Å; and (3) we determined the velocity of the right-moving wave packet at the Fermi level (i.e., the unit of simulation time) so as to be identical to the Fermi velocity calculated in Step 2.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Tada, T., Yamamoto, T. & Watanabe, S. Molecular orbital concept on spin-flip transport in molecular junctions. Theor Chem Acc 130, 775–788 (2011). https://doi.org/10.1007/s00214-011-1028-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00214-011-1028-3

Keywords

Navigation