Connectivity dependence of Fano resonances in single molecules
Received
7th January 2017
, Accepted 10th February 2017
First published on 10th February 2017
Abstract
Using a first principles approach combined with analysis of heuristic tight-binding models, we examine the connectivity dependence of two forms of quantum interference in single molecules. Based on general arguments, Fano resonances are shown to be insensitive to connectivity, while Mach–Zehnder-type interference features are shown to be connectivity dependent. This behaviour is found to occur in molecular wires containing anthraquinone units, in which the pendant carbonyl groups create Fano resonances, which coexist with multiple-path quantum interference features.
I. Introduction
Fano resonances1 occur when a localised state interacts with a continuum.2 In the context of single-molecule electronics, the continuum is provided by electrodes, which feed electrons into the molecule over a continuous range of energies. The appearance of Fano resonances in electron transport through single molecules was recognised in a series of studies of a family of rigid molecules3–5 which contained pendant groups. Subsequently it was shown that such groups could be utilised to control transport properties though single molecules6 and enhance their thermoelectric properties.7 The aim of the present paper is to examine how Fano resonances can be distinguished from other quantum interference (QI) effects in molecules.8–21 This issue is of interest, because recently unequivocal realisations of room temperature QI have been demonstrated in the laboratory22–35 and if QI is to be utilised in the design and optimisation of molecular-scale devices36–39 and thin films, the signatures of distinct forms of QI need to be clarified.
The aim of the present work is to demonstrate that connectivity can be used to distinguish between Fano resonances and multiple-path Mach–Zehnder-like interference effects. To illustrate how connectivity can be utilised, we start with an introductory discussion of the properties of a series of tight-binding lattices, which have also been used to study Fano resonances in quantum dot structures.40–42 Consider first the doubly-infinite one-dimensional tight-binding chain shown in Fig. 1a, described by the Schrodinger equation
| ε0φj − γφj−1 − γφj+1 = Eφj (−∞ < j < ∞), | (1) |
whose solutions are
φj = e
ikj where −π <
k < π. Substituting this into
eqn (1) yields the dispersion relation
E =
ε0 − 2
γ![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
cos
k. We are interested in solving scattering problems for a given choice of
E, for which the dimensionless wave vector
k satisfies
k(
E) = cos
−1[(
ε0 −
E)/2
γ] and the group velocity
v(
E) (in units of the atomic spacing) is given by

, where
ħ is Planck's constant. In what follows, the sign of
k(
E) is chosen such that the retarded Greens function for such a chain is
|  | (2) |
which means that: (a) when
k(
E) is real,
k(
E) is chosen such that
v(
E) > 0 (b) when
k(
E) is complex,
k(
E) is chosen such that Im
k(
E) > 0. These choices ensure that relative to the source
l, when considered as a function of
j,
gjl is either outgoing or decaying.
 |
| Fig. 1 A doubly-infinite one-dimensional crystalline tight-binding (TB) chain. (b) As for (a), but with a pendant site of energy ε1, attached to the chain by a coupling –α. (c) Decimating the pendant site from (b) yields an equivalent lattice with a single impurity site of energy ε0 + η. (d) As for (c), except the impurity is coupled to the lattice by weak links −γ1. | |
Now consider the chain of Fig. 1b, which contains a pendant site of energy ε1, connected to site j = 0 by a matrix element −α. Since this pendant site interacts with the continuum of energies associated with the doubly-infinite chain, we expect a destructive Fano resonance to appear. This is most easily demonstrated by decimating the pendant site to yield the mathematically equivalent lattice shown in Fig. 1c, which contains a single impurity on site j = 0 of energy ε0 + η, where η = α2/(E − ε1). Clearly for such an impurity-containing lattice, if η = ∞, the chain divides into two sections separated by an infinite tunnel barrier and therefore the transmission coefficient T(E) vanishes. This occurs when E = ε1 and therefore we conclude that T(ε1) = 0. More generally, the energy dependence of T(E) is obtained by computing the Greens function Gjl(E) of the chain of Fig. 1c and noting that for any j ≥ 0 and l ≤ 0,
| T(E) = [ħv(E)]2|Gjl(E)|2, | (3) |
Dysons equation yields
|  | (4) |
where the last step makes use of the fact that for
j ≥ 0 and
l ≤ 0,

. Hence, we obtain three equivalent expressions for
T(
E), all of which vanish when
E =
ε1 and
η = ∞:
|  | (5) |
where in the last expression,
Γ =
ħv(
E)/2 =
γ![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
sin
k,
σ = 2
γ![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
cos
k and

.
In the literature,43 it has been suggested that although the above anti-resonance can be classified as a Fano resonance, because the continuum of states within the periodic chain interferes with the state localized on the pendant site, the analogy does not hold for single molecule junctions where the continuum of the electrode states is interrupted by a junction, which acts as a tunnelling barrier. We see no compelling reason to make such a distinction, because the last expression in eqn (4) holds even when tunnelling barriers are present. For example for the lattice shown in Fig. 1d, where the site at j = 0 is connected to the 1-d periodic chains on the left and right by weak coupling elements −γ1, which act as tunnel barriers, eqn (4) still holds, but σ and Γ are replaced by
and
. Therefore as γ1/γ varies smoothly from zero to unity, the same functional form persists.
Nevertheless it is of interest to enquire how the above Fano resonances can be distinguished from other forms of destructive interference. In what follows, we propose that the key distinguishing feature lies in their connectivity dependence.
To clarify the role of connectivity consider the tight-binding ring of N sites shown in Fig. 2a, described by the Schrodinger equation
ε0φj − γφj−1 − γφj+1 = Eφj (1 < j < N) |
The solutions to this are
φjn = e
i2nπj/N where
n = 1, 2,…
N and the eigenvalues are
En =
ε0 − 2
γ![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
cos
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
2
nπ/
N. Similarly, the Green's function of this closed ring
44 is
|  | (6) |
where

. Since the Green's function of a closed system diverges when
E coincides with an eigenvalue, as expected,
A diverges when

;
i.e. when
E =
En.
 |
| Fig. 2 (a) A TB ring of N sites. (b) A TB ring with pendant sites attached to sites l and m. (c) A para-connected 6-membered ring. (d) A meta-connected 6-membered ring. | |
Now consider coupling such a ring via sites i and j to 1-d semi-infinite crystalline leads by weak coupling elements −γ1 and −γ2 respectively. As noted in ref. 44, the transmission coefficient for such a structure is
|  | (7) |
where

. In this expression,
|  | (8) |
Therefore provided E ≠ En, (so that Gring does not diverge) x vanishes in the limit
. Consequently in this weak coupling limit,
|  | (9) |
Eqn (7) and (9) illustrate the connectivity dependence of multiple-path QI in such rings. For example when electrons enter a 6-membered ring with energy in the middle of the HOMO–LUMO gap,
E =
ε0 and
k(
ε0) = π/2, as shown in
Fig. 2c, for
N = 6, the transmission in the case of
para coupling is determined by the Greens function
Gring(4, 1) =
A, which is non-zero and corresponds to constructive QI, whereas in the case of meta coupling (
Fig. 2d), it is determined by
Gring(3, 1) =
A![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
cos
k, which vanishes at the gap centre and corresponds to destructive QI.
The above sensitivity to connectivity (i, j) arises from the presence of Mach–Zehnder-type multiple-path interference in the ring, which for certain values of i, j and E can cause Gring(j, i) to vanish. As illustrated by the difference between the transmission coefficients of meta- and para-connected rings, this type of QI changes from destructive to constructive by only a single site shift of the connecting site (i.e. by changing the connection from site 3 to 4). We now examine the effect of introducing pendant sites into such rings, as illustrated by Fig. 2b, where the pendant sites of energy ε1 are coupled to sites l and m of the ring. After decimation, this structure again maps onto an equivalent lattice with impurity sites located at positions l and m, of energy ε0 + η, where η = α2/(E − ε1) and when E = ε1, the structure divides into two halves. Consequently T(ε1) vanishes, provided the sites i and j, which connect the ring to external crystalline leads, become disconnected by divergent sites at l and m. Provided the latter condition is satisfied, changing the connectivity does not change the energy E = ε1 at which the Fano destructive interference occurs. On the other hand for E ≠ ε1, transmission remains finite and Mach–Zehnder interference features remain, albeit with energies shifted by the presence of the pendant sites.
II. Transport calculations
After the above discussion of the main concept, in this section, we first employ a tight-binding model to probe the connectivity dependence of transport through the tight binding lattices a–f of Fig. 3 and 5 and then use density functional theory (DFT) to probe transport through the anthraquinone-based molecules, which are realisations of the abstract structure of Fig. 2b. As expected, we find that they possess connectivity-independent Fano resonances, which coexist with connectivity-dependent Mach–Zehnder-type interference features.
 |
| Fig. 3 (a) A bipartite lattice with odd to odd connectivity. (b) A bipartite lattice with odd to even connectivity. | |
For the lattice in Fig. 3, each atom within the core is assigned a site energy ε0 = 0 and a nearest neighbour hopping element −1. The atoms i, j are attached to the terminal sites of semi-infinite 1-d crystalline leads by a hopping element −γ1. In what follows, we choose γ1 = −0.2. The pendant site is assigned a site energy ε1 = −0.22 and is coupled to its nearest neighbour by a coupling α = −0.1.
The structures in Fig. 3 are examples of bipartite lattices, in which sites can be numbered such that odd numbered sites connect to even numbered sites only and vice versa. Consequently if the sites i, j attached to the leads are both even or both odd (as in Fig. 3a), multiple-path destructive interference is expected at E = 0. On the other hand, this destructive QI at E = 0 is expected to disappear when one of the connection points is shifted by one site, such that i is odd and j is even or vice versa, as in Fig. 3b. This connectivity dependence near at E = 0 is clearly evident in the transmission curves of Fig. 4. On the other hand, the transmission curves also show destructive QI at E = ε1, arising from Fano resonances associated with the pendant groups and these persist for both connectivities.
 |
| Fig. 4 Transmission curves for lattices (a and b) of Fig. 3 (ε1 = −0.22). | |
As noted in the previous section, pendant groups can be decimated to yield equivalent lattices with impurity site, whose site energies diverge when E = ε1. In the lattices of Fig. 3, this divergence cuts the structures into separate halves and therefore T(E) vanishes at E = ε1, for both connectivities. In contrast for lattices (c and d) of Fig. 5, only the lower section of the central ring is bisected and a current path from i to j persists. Therefore in this case the transmission does not vanish at E = ε1. For comparison, Fig. 6 shows the transmission coefficients of lattices (e and f), which contain no pendant groups. Since these are bipartite, (e) shows a transmission dips at E = 0 due to its odd–odd connectivity, whereas (f) shows no such dip, due to its odd–even connectivity. Comparison with the transmission curves of (c and d) show that these connectivity-dependent features persist in the presence of pendant groups and furthermore the presence of such groups leads to additional Fano-like interference features. However since these do not bisect the lattice, an anti-resonance does not occur at E = ε1. We therefore refer to this as an “incomplete Fano resonance”. Instead of a connectivity-independent anti-resonance at E = ε1, multiple-path interference features occur in the vicinity of E = ε1, whose energetic location is connectivity dependent.
 |
| Fig. 5 (upper two figures) Two lattices (c and d) with pendant sites, but different connectivities to the leads. (lower two figures) Two lattices (e and f) with different connectivities to the leads and no pendant sites. | |
 |
| Fig. 6 Transmission curves for the four lattices (c–f) of Fig. 5. | |
III. DFT calculations
So far the discussion has focused on tight-binding representations of the molecules shown in Fig. 7.
 |
| Fig. 7 (upper two figures) Two lattices (a and b) with two pendant sites, but different connectivities to the leads. (middle two figures) Two lattices (c and d) with one pendant sites, but different connectivities to the leads. (lower two figures) Two lattices (e and f) with different connectivities to the leads and no pendant sites. | |
For molecules (a and b) containing two carbonyl groups, Fig. 8 shows DFT results for the transmission coefficients (see Methods for more details). In common with the odd–odd connectivity of the tight binding lattice (a), the DFT transmission of (a) shows a clear destructive interference dip at E = 0, which disappears when the connectivity is switched to that of (b). On the other hand for both molecules, destructive interference features appear near E = −2.4 eV, which are almost independent of connectivity, signaling the presence of a Fano resonance.
 |
| Fig. 8 DFT transmission curves for molecules (a and b) of Fig. 7. | |
Fig. 9 shows DFT results for molecules (c–f) of Fig. 7. For molecules (e and f), which contain no pendant atoms, connectivity (e) possesses a transmission dip near E = 3 eV due to the odd–odd connectivity, which disappears upon switching to the odd–even connectivity of molecule (f). This transmission dip is also present for the odd–odd connectivity molecule (c) containing a single pendant group. Again this dip disappears upon switching to the odd–even connectivity of molecule (d). Furthermore, in common with the tight-binding results (c and d) of Fig. 6 and in contrast with the transmission curves of molecules (e and f), molecule (c) also possesses a lower energy destructive QI feature at E = 1.7 eV, which moves to lower energies upon switching connectivity to that of molecule (d), signaling the presence of an incomplete Fano resonance.
 |
| Fig. 9 DFT transmission curves for the four lattices (c–f) of Fig. 7. | |
IV. Conclusion
In summary we have shown that transport through anthraquinone-based molecular wires containing one or two pendant atoms exhibit Fano resonances, which can be differentiated from multiple-path Mach–Zehnder type interference through their resilience to changes in their connectivity to electrodes. This resilience is expected on general grounds, whenever a Fano resonance causes the wave-function of electrons traversing a molecule to decouple at a certain energy. If this does not occur, the Fano resonance is incomplete and the associated interference dip becomes a connectivity-dependent multiple path interference feature. Such interference features have the potential to increase the sensitivity of single-molecule sensors and the thermoelectric performance of single molecules6,7 and molecular films. Therefore the ability to preserving their energetic locations while selecting desirable connectivities is a useful ingredient in design strategies aimed at optimizing such properties.
V. Computational methods
Electronic structure calculations were performed using the DFT code SIESTA.45 The optimum geometry of the isolated a–f molecules were obtained by relaxing the molecules until all forces on the atoms were less than 0.05 eV Ang−1. The SIESTA calculations employed a double-zeta plus polarization orbital basis set, norm-conserving pseudopotentials, an energy cutoff of 200 Rydergs defined the real space grid and the exchange correlation functional was LDA. To calculate the conductance through these two molecules, they were attached to gold leads via the pyridyl anchor groups. The leads were constructed of 6 layers of (111) gold each contain 30 gold atoms and the optimum binding distance was calculated to be 2.3 Å between the terminal carbon atoms and a ‘top’ gold atom. A Hamiltonian describing these structures were produced using SIESTA and the zero-bias transmission coefficient T(E) was calculated using the Gollum code.46
Acknowledgements
This work was supported by the Swiss National Science Foundation (no. 200021-147143) as well as by the European Commission (EC) FP7 ITN “MOLESCO” project no. 606728 and UK EPSRC, (grant no. EP/K001507/1, EP/J014753/1, EP/H035818/1), and the Iraqi Ministry of Higher Education, Tikrit University (SL-20). A. K. I. acknowledges financial support from Tikrit University.
References
- U. Fano, Phys. Rev., 1961, 124, 1866–1878 CrossRef CAS
.
- A. E. Miroshnichenko, S. Flach and Y. S. Kivshar, Rev. Mod. Phys., 2010, 82, 2257 CrossRef CAS
.
- C. Wang, A. S. Batsanov, M. R. Bryce and I. Sage, Org. Lett., 2004, 6(13), 2181 CrossRef CAS PubMed
.
- C. Wang, A. S. Batsanov and M. R. Bryce, Faraday Discuss., 2006, 131, 221–234 RSC
.
- C. Wang, A. S. Batsanov and M. R. Bryce, J. Org. Chem., 2006, 71, 108–116 CrossRef CAS PubMed
.
- T. A. Papadopoulos, I. M. Grace and C. J. Lambert, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74(19), 193306 CrossRef
.
- C. M. Finch, V. M. Garcia-Suarez and C. J. Lambert, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79(3), 033405 CrossRef
.
- R. Stadler, M. Forshaw and C. Joachim, Nanotechnology, 2003, 14, 138 CrossRef CAS
.
- R. Baer and D. Neuhauser, J. Am. Chem. Soc., 2002, 124, 4200 CrossRef CAS PubMed
.
- G. C. Solomon, D. Q. Andrews, R. H. Goldsmith, T. Hansen, M. R. Wasielewski, R. P. Van Duyne and M. A. Ratner, J. Am. Chem. Soc., 2008, 130, 17309 CrossRef PubMed
.
- X. H. Zheng, G. R. Zhang, Z. Zeng, V. M. García-Suárez and C. J. Lambert, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80(7), 075413 CrossRef
.
- A. A. Kocherzhenko, L. D. A. Siebbeles and F. C. Grozema, J. Phys. Chem. Lett., 2011, 2, 1753 CrossRef CAS
.
- R. Stadler, S. Ami, M. Forshaw and C. Joachim, Nanotechnology, 2004, 15, S115 CrossRef CAS
.
- E. H. van Dijk, D. J. T. Myles, M. H. van der Veen and J. C. Hummelen, Org. Lett., 2006, 8, 2333 CrossRef CAS PubMed
.
- T. Markussen, J. Schiötz and K. S. Thygesen, J. Chem. Phys., 2010, 132, 224104 CrossRef PubMed
.
- M. Magoga and C. Joachim, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 16011 CrossRef CAS
.
- J. P. Bergfield, M. A. Solis and C. A. Stafford, ACS Nano, 2010, 4, 5314 CrossRef CAS PubMed
.
- A. B. Ricks, G. C. Solomon, M. T. Colvin, A. M. Scott, K. Chen, M. A. Ratner and M. R. Wasielewski, J. Am. Chem. Soc., 2010, 132, 15427 CrossRef CAS PubMed
.
- T. Markussen, J. Schiötz and K. S. Thygesen, J. Chem. Phys., 2010, 132, 224104 CrossRef PubMed
.
- C. M. Finch, S. Sirichantaropass, S. W. Bailey, I. M. Grace, V. M. Garcia-Suarez and C. J. Lambert, J. Phys.: Condens. Matter, 2008, 20, 022203 CrossRef
.
- G. C. Solomon, J. P. Bergfield, C. A. Stafford and M. A. Ratner, Beilstein J. Nanotechnol., 2011, 2, 862 CrossRef CAS PubMed
.
- H. Vazquez, R. Skouta, S. Schneebeli, M. Kamenetska, R. Breslow, L. Venkataraman and M. Hybertsen, Nat. Nanotechnol., 2012, 7, 663 CrossRef CAS PubMed
.
- S. Ballmann, R. Härtle, P. B. Coto, M. Elbing, M. Mayor, M. Bryce, M. Thoss and H. B. Weber, Phys. Rev. Lett., 2012, 109, 056801 CrossRef PubMed
.
- S. Aradhya, J. S. Meisner, M. Krikorian, S. Parameswaran, M. L. Steigerwald, C. Nuckolls and L. Venkataraman, Nano Lett., 2012, 12, 1643 CrossRef CAS PubMed
.
- V. Kaliginedi, P. Moreno-García, H. Valkenier, W. Hong, V. García-Suárez, P. Buiter, J. L. Otten, J. C. Hummelen, C. J. Lambert and T. Wandlowski, J. Am. Chem. Soc., 2012, 134, 5262 CrossRef CAS PubMed
.
- S. V. Aradhya and L. Venkataraman, Nat. Nanotechnol., 2013, 8, 399 CrossRef CAS PubMed
.
- C. R. Arroyo, S. Tarkuc, R. Frisenda, J. S. Seldenthuis, C. H. Woerde, R. Eelkema, F. C. Grozema and H. S. van der Zant, Angew. Chem., Int. Ed., 2013, 125, 3234 CrossRef
.
- C. M. Guédon, H. Valkenier, T. Markussen, K. S. Thygesen, J. C. Hummelen and S. J. van der Molen, Nat. Nanotechnol., 2012, 7, 305 CrossRef PubMed
.
- J. Xia, B. Capozzi, S. Wei, M. Strange, A. Batra, J. R. Moreno, R. J. Amir, G. C. Solomon, L. Venkataraman and L. M. Campos, Nano Lett., 2014, 14, 2941 CrossRef CAS PubMed
.
- L. Rincón-García, A. K. Ismael, C. Evangeli, I. Grace, G. Rubio-Bollinger, K. Porfyrakis, N. Agraït and C. J. Lambert, Nat. Mater., 2016, 15, 289–293 CrossRef PubMed
.
- C. S. Lau, H. Sadeghi, G. Rogers, S. Sangtarash, P. Dallas, K. Porfyrakis, J. H. Warner, C. J. Lambert, G. Andrew, D. Briggs and J. Mol, Nano Lett., 2016, 16, 170–176 CrossRef CAS PubMed
.
- P. Gehring, H. Sadeghi, S. Sangtarash, C. S. Lau, J. Liu, A. Ardavan, J. H. Warner, C. J. Lambert, G. Andrew, D. Briggs and J. A. Mol, Nano Lett., 2016, 16(7), 4210–4216 CrossRef CAS PubMed
.
- H. Sadeghi, J. Mol, C. Lau, G. A. D. Briggs, J. Warner and C. J. Lambert, Proc. Natl. Acad. Sci. U. S. A., 2015, 9, 2658–2663 CrossRef PubMed
.
- S. Sangtarash, C. Huang, H. Sadeghi, G. Sorohhov, J. Hauser, T. Wandlowski, W. Hong, S. Decurtins, S. Liu and C. J. Lambert, J. Am. Chem. Soc., 2015, 137(35), 11425–11431 CrossRef CAS PubMed
.
- Y. Geng, S. Sangtarash, C. Huang, H. Sadeghi, Y. Fu, W. Hong, T. Wandlowski, S. Decurtins, C. J. Lambert and S. Liu, J. Am. Chem. Soc., 2015, 137(13), 4469–4476 CrossRef CAS PubMed
.
- D. Manrique, Q. Al-Galiby, W. Hong and C. J. Lambert, Nano Lett., 2016, 16, 1308–1316 CrossRef CAS PubMed
.
- S. Sangtarash, H. Sadeghi and C. J. Lambert, Nanoscale, 2016, 8, 13199–13205 RSC
.
- Q. Al Galiby, H. Sadeghi, L. Algharagholy, I. Grace and C. J. Lambert, Nanoscale, 2016, 8, 2428–2433 RSC
.
- A. Ismael, I. Grace and C. J. Lambert, Nanoscale, 2015, 7(41), 17338–17342 RSC
.
- L. Haizhou, R. Lü and B.-F. Zhu, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71(23), 235320 CrossRef
.
- X. Jun-Jie, S.-Q. Duan and W. Zhang, Nanoscale Res. Lett., 2015, 10(1), 223 CrossRef PubMed
.
- X. R. Wang, Y. Wang and Z. Z. Sun, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 65(19), 193402 CrossRef
.
- R. Stadler and T. Markussen, J. Chem. Phys., 2011, 135, 154109 CrossRef PubMed
.
- C. J. Lambert, Chem. Soc. Rev., 2015, 44, 875–888 RSC
.
- J. M. Soler, E. Artacho, J. D. Gale, J. García, J. Junquera, P. Ordejón and D. Sánchez-Portal, The SIESTA method for ab initio order-N materials simulation, J. Phys.: Condens. Matter, 2002, 14(11), 2745 CrossRef CAS
.
- J. Ferrer, C. J. Lambert, V. García-Suárez, D. Manrique, D. Visontai, L. Oroszlany, R. Rodríguez-Ferradás, I. Grace, S. Bailey, K. Gillemot and H. Sadeghi, GOLLUM: a next-generation simulation tool for electron, thermal and spin transport, New J. Phys., 2014, 16(9), 093029 CrossRef
.
|
This journal is © the Owner Societies 2017 |
Click here to see how this site uses Cookies. View our privacy policy here.