thorsten hansen et al- interfering pathways in benzene: an analytical treatment

Upload: dfmso0

Post on 06-Apr-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    1/12

    Interfering pathways in benzene: An analytical treatment

    Thorsten Hansen,a Gemma C. Solomon, David Q. Andrews, and Mark A. RatnerDepartment of Chemistry, Northwestern University, 2145 Sheridan Road, Evanston, Illinois 60208, USA

    Received 17 June 2009; accepted 18 October 2009; published online 18 November 2009

    The mechanism for off-resonant electron transport through small organic molecules in metallic

    junctions is predominantly coherent tunneling. Thus, new device functionalities based on quantum

    interference could be developed in the field of molecular electronics. We invoke a partitioning

    technique to give an analytical treatment of quantum interference in a benzene ring. We interpret the

    antiresonances in the transmission as either multipath zeroes resulting from interfering spatial

    pathways or resonance zeroes analogous to zeroes induced by sidechains. 2009 American

    Institute of Physics. doi:10.1063/1.3259548

    I. INTRODUCTION

    Quantum coherences of exciton transfer processes have

    been measured in photosynthetic reaction centers, suggesting

    that quantum mechanical effects may underlie the near-

    perfect quantum efficiency of photosynthesis.1,2

    Inspired by

    nature, one could envision molecular electronic devices with

    functionalities derived from quantum interference.39

    This

    appears feasible, as the dominant mechanism for electron

    transfer in many molecular transport junctions is coherent

    transport, and the current is generally well described by the

    Landauer equation. The coherent transport picture breaks

    down as the junction approaches resonant transport due to

    charging and increased electronic-vibrational coupling; con-

    sequently, coherent molecular devices should function in the

    off-resonant regime. Understanding destructive interference

    features is therefore important for the prospects of molecular

    electronics as these features can potentially be integrated into

    designs and manipulated to obtain molecular devices with alarge dynamic range.1025

    Interference effects in transport through aromatic sys-

    tems in general, and through benzene in particular, have be

    studied both experimentally26,27

    and theoretically.2838

    The

    large variation in transmission through a benzene ring as the

    substitution is changed from ortho to meta and to para

    shown in Fig. 1 has been attributed to quantum interfer-ence. Working with a model Hamiltonian, at the Hckel level

    of electronic structure, we shall demonstrate how to interpret

    the negative interference features of the transmission func-

    tions for benzene in terms of interfering pathways by which

    the tunneling electron travels one way and another through

    the benzene ring. All but one of the antiresonances fit intothis interpretation, and we shall make the distinction between

    multipath zeroes and resonance zeroes of the transmission

    function, a distinction that may prove helpful for the design

    of novel molecular devices.

    The quantitative accuracy of contemporary transport cal-

    culations, with the electronic structure treated at the density

    functional theory level, is far superior to the simplistic model

    calculations used in this work. Yet, we expect the conceptual

    conclusions of our analytical treatment to carry over to larger

    numerical calculations, not unlike the computational work of

    Ke et al.,37

    where interference features were not seen in ben-

    zene due to strong mixing of the molecular -system and the

    orbitals of the electrode, but they were evident in largerconjugated rings.

    The key mathematical entity in molecular conductance

    calculations is the molecular Greens function, which is cal-

    culated from the molecular Hamiltonian and, possibly, addi-

    tional self-energies describing interactions with surroundings

    or additional degrees of freedom. Most computational ap-

    proaches calculate the transmission via the Greens function.

    Working at the Hckel level of theory, we are able to give a

    fully analytical treatment, dissecting the molecular Greens

    functions using two different theoretical techniques. We in-

    voke a Lwdin partitioning scheme to interpret the antireso-

    nances of the molecular Greens function as the result ofinterfering spatial pathways. To describe the coupling be-

    tween the molecule and metallic leads, we use perturbation

    theory to first order in the self-energy. This will allow us to

    treat linewidths and the splitting of degenerate orbitals on the

    imaginary energy axis, which causes dips in the transmis-

    sion.

    aAuthor to whom correspondence should be addressed. Electronic mail:

    [email protected]. FIG. 1. Three different ways to attach electrodes to benzene.

    THE JOURNAL OF CHEMICAL PHYSICS 131, 194704 2009

    0021-9606/2009/13119 /194704/12/$25.00 2009 American Institute of Physics131, 194704-1

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

    http://dx.doi.org/10.1063/1.3259548http://dx.doi.org/10.1063/1.3259548http://dx.doi.org/10.1063/1.3259548http://dx.doi.org/10.1063/1.3259548
  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    2/12

    II. MOLECULAR TRANSMISSION

    The rate of electron transfer in molecular donor-bridge-

    acceptor systems is often described by a generalized Fermis

    golden rule,

    kAD =2

    ad

    tadd2a d. 1

    Here tad describes the transition matrix and the sums runover orbitals, with energies d and a, spanning the donor

    and acceptor components of the molecule, respectively. A

    definition of the transition matrix is given by

    tadE = Vad + ij

    VaiGijretEVjd. 2

    Here V is the coupling, or perturbation, and GretE is theretarded Greens function, described below.

    3943In this work,

    we assume that the electron transfer is bridge mediated; thus

    Vad= 0.

    Most contemporary approaches to molecular transport

    junctions work with the Landauer equation

    I=2e

    h dEfLE fRETE. 3

    The transmission function,

    TE = TrLGretERGadvE , 4

    is the property that is calculated in most computational ap-

    proaches to molecular electronics.4452

    The advanced

    Greens function, GadvE, is the Hermitian conjugate of theretarded Greens function described below, GadvE= GretE, and X is the spectral density of the left or rightlead, X=L and R and is defined as

    ijLE = 2d VidVdjE d. 5

    To elucidate the analogy between the rate expression 1and the Landauer equation, we rewrite the transmission as a

    FisherLee type expression18,53,54

    TE = 42ad

    ijkl

    VaiGijretEVjdVdkGkl

    advEVla

    E dE a

    = 42ad

    tadE2E dE a . 6

    The rate expression, Eq. 1, is then recovered by integratingover the independent energy variable

    kAD =1

    h dETE. 7

    A. The Greens function

    Under quite general conditions we can study the quan-

    tum interference phenomena occurring in either photochemi-

    cal donor-bridge-acceptor systems or molecular transport

    junctions by focusing on the retarded Greens operator of the

    bridging molecule, which is defined by

    GretE =1

    E H+ i. 8

    Here H is the Hamiltonian in question, and is a positive

    infinitesimal. In the case of donor-bridge-acceptor systems,

    we take H to be the Hckel Hamiltonian of the molecular

    bridge. For a molecular transport junction, we will add the

    coupling to the leads as an imaginary potential. All Greens

    functions described below will be retarded, so the superscriptret and the imaginary infinitesimal i will be omitted.

    Mathematically, we shall adopt two complementary

    viewpoints when analyzing the Greens functions. From a

    linear algebra point of view, the retarded Greens operator is

    a restriction of the resolvent, GE = EH1, which is ob-tained by matrix inversion of the operator EH. A closed

    expression for a specific matrix element of the resolvent is

    given as the ratio between the so-called cofactor, CjiEH,and the secular determinant,

    GijE =CjiE H

    detE H

    . 9

    The jith cofactor of a matrix A is given as

    CjiA = 1j+idetAji, 10

    where Aji is the submatrix of A obtained by removing the jth

    row and the ith column.

    From a complex analysis point of view, we consider the

    Greens function, Eq. 9, as a function of the in generalcomplex energy variable E. It is meromorphic since it isgiven as the ratio of two polynomials in E. The poles of the

    Greens function thus coincide with the zeros of the seculardeterminant, i.e., the eigenenergies of the molecular Hamil-

    tonian.

    Likewise, antiresonances in the transmission, where the

    Greens function and hence the transmission function equal

    zero, are due to zeroes in the cofactor. However, not all ze-

    roes of the cofactor result in an antiresonance in the trans-

    mission. The Hamiltonian H is a simple matrix, i.e., it has a

    full set of eigenvectors and is thus diagonalizable. A theorem

    for simple matrices implies that all poles of the Greens func-

    tion are simple poles; thus any degenerate eigenvalues of H

    are roots of the cofactor.55

    As we shall see below, the break-

    ing of a degeneracy will therefore make an additional anti-

    resonance appear.In simple cases, where the Hamiltonian can be diagonal-

    ized exactly, it can be helpful to expand the Greens operator

    in a basis of eigenvectors, ui. We write

    GE =1

    E H=

    i

    uiui

    E i. 11

    Thus the residue of the pole at i is the orthogonal projection

    onto the eigenvector, uiui. The residue of a degenerateeigenvalue is obtained by summation over a full basis for the

    eigenspace, resi = j;j=iujuj.

    194704-2 Hansen et al. J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    3/12

    III. MOLECULAR TOPOLOGY

    Only in the simplest cases can we diagonalize the

    Hckel Hamiltonian exactly and obtain analytical expres-

    sions for the eigenenergies and eigenvalues. However, by

    projecting the full problem onto a small subspace of the full

    Hilbert space, we obtain an effective Hamiltonian for this

    smaller space. The parts of the molecule that have been dis-

    regarded, e.g., some substituents, are therefore only de-

    scribed indirectly, yet their impact on the subspace in ques-

    tion is described exactly.

    The Greens functions of interest for electron transmis-

    sion are then expressed in terms of the effective Hamiltonian.The parts of the molecule that have been removed will ap-

    pear as self-energies in the Greens functions, and the math-

    ematical structure of the self-energies will directly reflect the

    topology of the molecule. In this way we obtain an exact

    analytical expression for the Greens functions.

    A. Lwdin partitioning

    First, we shall briefly review the Lwdin partitioning

    technique.5658

    Let P = iii be a projection operator ontothe subspace of interest, and let Q = 1 P be the projection

    onto the complementary subspace. Orthogonal projection op-erators are Hermitian, P = P, and idempotent, P2 = P. We can

    write the uncoupled Hamiltonian as H0 = iii; thus theprojection operators commute with H0, P ,H0 = Q ,H0 = 0.Beginning with the definition of the Greens function,

    E H0 VGE = 1 , 12

    and multiplying on the right with P and on the left with P or

    Q, we insert P + Q =1 and use the commutation relations to

    obtain

    PE HPPGEP PVQQGEP = P , 13

    QVPPGEP + QE HQQGEP = 0. 14From these equations, the projection, PGEP, of theGreens function can be isolated,

    PGEP =P

    E PH0P PEP. 15

    Here E is a self-energy, or level-shift, operator defined by

    E = V+ VQ

    E QH0Q QVQV. 16

    Note the close analogy to Eq. 2. As an example, considerthe three-site Hamiltonian of the system depicted in Fig. 2,

    H= 0

    0 . 17

    Now we remove the sidechain by defining P = 11 + 22.This implies Q = 33, and using Eq. 16, we obtain a self-energy

    PEP = PVQ 1E H

    QVP =

    2

    E 22 , 18

    which is introduced in the effective Hamiltonian

    Heff= + E

    . 19Even though we have reduced the dimension of the Hamil-

    tonian, the self-energy ensures that we obtain the exact result

    for Greens functions defined in the reduced Hilbert space,

    G21E = 21

    E Heff1

    =

    E EE 2

    =1

    E

    1

    E E 2

    E

    . 20

    The last expression could be obtained via Dysons equation,

    which in this case reads G21E = G220 EG11E. When the

    energy variable E is resonant with the sidechain energy ,

    the self-energy diverges. This implies G11 =0, which of-fers a characterization of the antiresonances induced by

    sidechains. A pole in the self-energy implies an antireso-

    nance in the Greens function.1321,24

    In molecular electronics, the metal leads bonded to a

    molecule are also described with the partitioning technique.

    Often, the so-called wide-band approximation is used, in

    which case only the resonant part of the self-energy remains,

    constituting an imaginary potential in the molecular Hamil-

    tonian. This approach is taken in Sec. V.

    B. Paths through cyclic systems

    The ring topology of any cyclic molecule means that the

    partitioning technique can be used to describe the transmis-

    sion in terms of different spatial paths around the ring. For

    example, in benzene the self-energy terms appearing in the

    effective Hamiltonian HeffE can be assigned to one of twodistinct spatial paths, A or B, between sites a and b, asshown in Fig. 3,

    HeffE = + aaA + aaB abA + abBba

    A + baB + bb

    A + bbB . 21

    The Hamiltonian HretE is symmetric since abX =ba

    X for X

    =A ,B.

    If we consider the diagonal part of this effective Hamil-

    tonian as the unperturbed Hamiltonian and the off-diagonal

    FIG. 2. A three-site model wire.

    194704-3 Interfering pathways in benzene J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    4/12

    elements as the perturbation, we can use Dysons equation to

    write the off-diagonal element of the Greens function

    GbaE as

    GbaE =1

    E bbA bb

    B ba

    A+ ba

    B abA

    + abB

    E aaA aa

    B

    baA

    + baB

    1

    E aaA

    aaB . 22

    This expression takes the form GbbEbaEGaa0 E, where

    the three factors give the contributions from site b, the twopaths, and site a, respectively. The first factor, GbbE, isthe full Greens function for the site b, with self-energiesdescribing virtual excursions into the bridges, bb

    A and bbB ,

    and to the site a and back, baA +ba

    B abA +ab

    B / E aa

    A aaB . Note that the two distinct paths lead to 22 = 4

    similar terms in the self-energy. The last factor is the un-

    coupled Greens function for the site a, with self-energiesdescribing virtual excursions into the bridges A and B.

    There are two conditions that will cause the Greens

    function to have a zero and thus an antiresonance in the

    transmission. Either the sum of the self-energies, A +B, is

    zero or the product GbbEGaa0 E is zero because of a pole in

    one of the self-energies, A or B. We will refer to the

    former as multipath zeroes because they occur when the self-

    energies of the two paths around the ring cancel each other.

    We will refer to the latter as resonance zeroes because they

    occur when the energy variable E is resonant with an

    eigenenergy of the substructure of the molecule described by

    the self-energy. This is directly analogous to the zeroes

    caused by sidechains, as described above.

    In the terminology of the renormalized perturbation

    expansion,

    59,60

    used, e.g., to describe electron transfer inproteins,41

    multipath zeroes are the result of interference of

    two or more skeleton paths. Resonance zeroes, on the other

    hand, result from a decoration that all skeletons paths have in

    common.

    C. Cofactor factorization theorem

    We now have two ways of characterizing the zeroes of

    the Greens function and thus the antiresonances of the trans-

    mission function. On the one hand, zeroes of the Greens

    function are zeroes of the cofactor, as seen from Eq. 9. Onthe other hand, from the discussion of Eq. 22, zeroes of the

    Greens function are either multipath zeroes, meaning zeroes

    of the sum of self-energies A +B, or resonance zeroes,

    where either A or B is divergent. We can elucidate the

    connection with a factorization theorem for the cofactor.

    In the simple case, when the Hamiltonian is reduced to

    two dimensions, say, P = 11 + 22, the off-diagonal co-factors that we are interested in can be factorized as:

    C12E H = detQE HQ2E1 . 23

    To derive this result, we begin with the definition of the

    cofactor from Eq. 9,

    C12E H = detE H2E1. 24

    With the projection operators P and Q, we can write the

    Hamiltonian as a block matrix,

    H= PH P P HQQHP QHQ

    , 25

    and use a standard identity for the determinant of a blockmatrix to obtain

    detE H = detQE HQdetE PH0P PEP.

    26

    Defining Heff= PH0P + PEP, we can now write Eq. 24as

    C12E H = detQE HQdetE Heff2

    1

    E Heff1

    27

    =detQE HQC12E Heff. 28

    When Heff is a two by two matrix, we have C12EHeff

    = 2E1, which completes the proof. Thus multipath ze-roes are zeroes of the self-energy 2E1, whereas reso-nance zeroes are zeroes of the determinant detQEHQ.We shall use this factorization below when discussing the

    ortho, meta, and para cofactors of benzene.

    IV. UNDERSTANDING BENZENE

    We now apply these methods to study the transport

    through the prototypical cyclic system: benzene. A Hckel

    Hamiltonian for the system of benzene can be written as

    H= i=1

    6

    ii + i=1

    6

    ii + 1 + i + 1i , 29

    where i represents the p-orbital on atom i, and we define7 1 to describe the ring topology. The different substi-tution patterns result from coupling different sites to the elec-

    trodes, sites 1 and 2, 1 and 3, and 1 and 4 for ortho,meta, and para, respectively.

    FIG. 3. Two pathways through benzene; in this case, para.

    194704-4 Hansen et al. J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    5/12

    These changes in the substitution pattern result in dra-

    matically different transmissions through the system. Figure

    4 illustrates the different transmissions on both linear and

    logarithmic scales. The differences in the resonances can be

    seen clearly on the linear scale and, in particular, the dra-

    matic differences in the antiresonances on the logarithmic

    scale.

    The relevant energy range for chemical reactivity is be-tween the highest occupied and the lowest unoccupied mo-

    lecular orbitals. In this entire range the transmission in the

    meta case is largely suppressed due to the antiresonance at

    zero. This is fully consistent with the rules for electrophilic

    substitution reactions on benzene; the ortho and para posi-

    tions are electron rich or deficient depending on whether the

    first substituent is electron donating or withdrawing. The

    meta position, on the other hand, has no coupling to the first

    substituent due to the suppression of the transmission

    function.61

    Now we apply the techniques developed previ-

    ously to see how these differences arise from the properties

    of the Greens function.

    A. Constructing the Greens function

    The Hckel Hamiltonian of benzene can be diagonalized

    exactly, and we take advantage of this to calculate explicit

    expressions for the Greens functions by using Eq. 11.Whereas the results from the previous section do not require

    diagonalization and are thus applicable to all molecules,

    knowledge of the eigenenergies and molecular orbitals of

    benzene enables us to give a more explicit and detailed

    analysis of our case study. The eigenvalues are + 2,

    + , , and 2, and the secular determinant factorizes

    as

    detE H = E 2E 2E + 2

    E + 2. 30

    We construct a unitary matrix U by writing a set of normal-

    ized eigenvectors ui as columns,

    0

    1

    -1 0 1

    (E)

    E10

    -6

    10-4

    10-2

    100

    -1 0 1

    T(E)

    E

    0

    1

    -1 0 1

    (E)

    E10-6

    10-4

    10-2

    100

    -1 0 1

    T(E)

    E

    0

    1

    -1 0 1

    (E)

    E

    10-6

    10-4

    10-2

    100

    -1 0 1

    T(E)

    E

    FIG. 4. Transmission function TE for the ortho top, meta middle, and para bottom systems. = 0, =0.5, and =0.2. Linear scale on the left andlogarithmic scale on the right. In all cases both the exact solution red and a simplified perturbative treatment of the electrodes black as outlined in Sec. Vare shown.

    194704-5 Interfering pathways in benzene J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    6/12

    U=1

    61

    1

    23

    2

    1

    2 3

    2 1

    1 2 0 2 0 1

    1 1

    2 3

    2

    1

    23

    2 1

    11

    2 3

    2

    1

    2 3

    21

    1 2 0 2 0 1

    11

    23

    2

    1

    23

    21

    . 31When donor and acceptor entities, or two electrodes, are at-

    tached to a benzene molecule in either ortho, meta, or para

    configurations, different matrix elements of the Greens op-

    erator, Eq. 11, are relevant. In the ortho case the attach-ments are to adjacent atoms, e.g., 1 and 2, and we chooseto calculate G21E = 2GE1 as

    G21E = i

    2uiui1

    E i32

    =1

    6

    1

    E 2+

    1

    6

    1

    E

    1

    6

    1

    E +

    1

    6

    1

    E + 2. 33

    The residues for the ortho, meta, and para Greens function

    elements are tabulated in Table I. Each residue is a matrix

    element of an off-diagonal projection operator, which is whythe sum of residues in each case is zero. Note also that the

    ortho and para Greens functions are even functions of en-

    ergy, G21E = G21E+ , whereas the meta matrix ele-ment is an odd function of energy, G31E = G31E+ ; thus the meta matrix element will have a zero, an anti-resonance, at E= . These symmetries derive from Coulsons

    pairing theorem, which ensures a symmetric Hckel spec-

    trum for alternant hydrocarbons.62

    In graph theory, these are

    the molecules that can be represented by bipartite, or two-

    color, graphs.63

    After calculating the Greens functions ex-

    actly, we now turn to a detailed analysis, in terms of different

    pathways, for each substitution pattern.

    1. Ortho

    First, we identify the paths through the molecule. In

    ortho substituted benzene one pathway is the direct bond

    between the two atoms; thus 21A = and 11

    A =22A =0, as

    there are no intervening sites. The second pathway consists

    of four sites and has the Hckel spectrum of butadiene. The

    eigenenergies are + , + /, /, , where

    = 1 +5 /2 = 1.618. . . is the golden ratio and1 /=0.618.. . = 1. A set of eigenvectors is given

    by c1 , c2 , c2 , c1 , c2 , c1 , c1 , c2 , c2 , c1 , c1 , c2 , c1 ,c2 , c2 , c1, where c1 = 1 / 22 + and c2 = c1. The off-diagonal self-energy is given as

    21B E =

    c122

    E

    c222

    E

    +c2

    22

    E +

    c1

    22

    E + . 34

    The diagonal self-energy 11B

    =22B

    is given by a similar ex-pression, except that all residues are positive.

    The ortho cofactor can now be expressed as the product

    of the sum of the self-energies of the paths, +21B E, and

    the secular determinant of the Hckel Hamiltonian for buta-

    diene,

    C12E H = + 21B detQE HQ

    = E 2 2E 2 22 . 35

    Figure 5 shows the Greens function, the sum of the

    self-energies, and the cofactor plotted as functions of energy.

    TABLE I. Residues in the ortho, meta, and para cases.

    GijE +2 + 2

    Ortho G21E1

    6

    1

    6

    1

    6

    1

    6

    Meta G31E1

    6

    1

    6

    1

    6

    1

    6

    Para G41E 16

    13

    13

    1

    6

    2

    1

    0

    1

    2

    -1 0 1 E

    G31(E)

    A31(E) +

    B31(E)

    C31(E-H)

    FIG. 6. For meta substitution, the Greens function green, 31A E

    +31B E blue, and cofactor red.

    2

    1

    0

    1

    2

    -1 0 1 E

    G21(E)

    + B21(E)

    C21(E-H)

    FIG. 5. For ortho substitution, the Greens function green, +21B E

    blue, and cofactor red.

    194704-6 Hansen et al. J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    7/12

    The cofactor is a fourth order polynomial in E and has four

    real roots. As mentioned above, there will always be roots at

    the degenerate eigenvalues, in this case , to ensure that

    these poles of the Greens function are simple poles. As we

    shall see below, the zeroes at will manifest themselves

    when the degeneracy of the eigenvalues is lifted. The two

    other roots of the cofactor, at 2, cause the Greensfunction to have zeroes.

    The sum of the self-energies for each path, +21B E,

    has the four zeroes in common with the cofactor, and we call

    them multipath zeroes by the distinction made above. Thus

    the idea of interfering pathways can explain all features of

    ortho substituted benzene.

    2. Meta

    In meta substituted benzene the shorter path comprises a

    single site; thus 31A =11

    A =33A = 2 / E . The longer path

    consists of three sites and has the Hckel spectrum of pro-

    pene, with eigenvalues +2, , 2. A set of eigen-vectors is given by 1 /2 , 1 /2 , 1 /2 , 1 /2 , 0 ,1 /2 , 1 /2,1 /2 , 1 /2. The off-diagonal self-energy isgiven as

    -2

    -1

    0

    1

    2

    -1 0 1 E

    G41(E)

    A41(E) +

    B41(E)

    C41(E-H)

    FIG. 7. For para substitution, the Greens function green, 41A E +41

    B Eblue, and cofactor red.

    4

    3

    2

    1

    0

    1

    2

    3

    4

    -1 0 1

    ReG(E)

    E-5

    -4

    -3

    -2

    -1

    0

    1

    2

    3

    4

    5

    -1 0 1

    ImG(E)

    E

    4

    3

    2

    1

    0

    1

    2

    3

    4

    -1 0 1

    ReG(E)

    E-5

    -4

    -3

    -2

    -1

    0

    1

    23

    4

    5

    -1 0 1

    ImG(E)

    E

    4

    3

    2

    1

    0

    1

    2

    3

    4

    -1 0 1

    ReG(E)

    E-5

    -4

    -3

    -2

    -1

    0

    12

    3

    4

    5

    -1 0 1

    ImG(E)

    E

    FIG. 8. The real left and imaginary right parts of the Greens functions for from top to bottom ortho, meta, and para substitution patterns. Left: the realpart Re GE for the uncoupled system green, the coupled system calculated exactly red, and the coupled system calculated by first order perturbation in black. Right: the imaginary part Im GE for the coupled system calculated exactly blue and the first order perturbation in black. For all plots: = 0, =0.5, and =0.2.

    194704-7 Interfering pathways in benzene J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    8/12

    31B E =

    2

    4

    1

    E 2

    2

    2

    1

    E +

    2

    4

    1

    E + 2.

    36

    The diagonal self-energy 11B =33

    B is given by a similar

    expression, except that all residues are positive. The cofactor

    is given as the sum of the self-energies, 31A E +31

    B E,times the secular determinant of the paths,

    C13E H = 31A E + 31

    B EdetQE HQ

    = 2E E 2 2. 37

    Figure 6 shows the Greens function, the sum of the self-

    energies, and the cofactor plotted as functions of energy. The

    cofactor is a third order polynomial in E and has three real

    roots. As mentioned above, there will always be roots at the

    degenerate eigenvalues, in this case , to ensure that

    these poles of the Greens function are simple poles. The

    third root causes the Greens function to have a zero at E

    = .

    The sum of the self-energies, 31A E +31

    B E, has two

    zeroes in common with the cofactor E= . Again, theseare multipath zeroes, which will manifest themselves when

    the eigenvalue degeneracy is lifted. Notably, the third zero, at

    E= , cannot be thought of as resulting from interfering path-

    ways. We call it a resonance zero and emphasize that it is

    analogous to zeroes induced by sidechains in a molecular

    structure, a fact that should be considered when designing

    molecules for interference experiments since the zero is at

    E= and thus will be near the Fermi level in experimental

    realizations.

    3. Para

    For para substitution, the two paths are identical and

    each consists of two sites. The eigenvalues are

    with bonding and nonbonding eigenvectors

    1 /2 , 1 /2 , 1 /2,1 /2. The off-diagonal self-energyis given by

    41A E = 41

    B E =2

    2

    1

    E

    2

    2

    1

    E + . 38

    Each diagonal self-energy is given by a similar expression

    with both residues positive. The cofactor is given as the

    product of the sum of the self-energies and the secular deter-

    minants of the paths,

    C14E H = 41A E + 41

    B EE 2 22

    = 23E 2 2 . 39

    Figure 7 shows the Greens function, the sum of the self-

    energies, and the cofactor, plotted as a function of energy.

    The cofactor is a second order polynomial in E and has two

    real roots. These are located at the degenerate eigenvalues

    as described above. These zeroes are provided by the

    secular determinant and are therefore resonance zeroes. Aswe shall discuss below, they will not, for symmetry reasons,

    manifest themselves in the transmission function when the

    degeneracies are lifted.

    V. LINEWIDTHS

    Until now, we have focused on the Greens function de-

    rived from the Hamiltonian of the isolated molecule. As the

    molecule couples to the two metal electrodes, the molecular

    levels broaden and the resonances in the transmission func-

    tion acquire width. A self-energy in the molecular Hamil-

    tonian captures this. The imaginary part of the self-energy

    describes the broadening, whereas the real part shifts theenergy levels. We work in the wide-band approximation

    where the self-energy is purely imaginary. For each atomic

    site that couples to an electrode, we shift the energy as

    i

    2. 40

    The spectral density is denoted . The sign of the energy

    shift is negative because we work with retarded Greens

    functions.

    We focus on the atomic sites, say, a , b, that arecoupled directly to the electrodes. Using the Lwdin parti-

    tioning to project onto the Hilbert space that they span, we

    obtain an effective Hamiltonian, which for our test case ofbenzene reads

    HeffE = i

    2+ aa

    A+ aa

    Bab

    A + abB

    baA + ba

    B i

    2+ bb

    A+ bb

    B ,41

    where A and B, as above, denote the two pathways through

    benzene. For a 22 Hamiltonian, the resolvent can be ob-

    tained analytically, and the GbaE element reads

    GbaE =ba

    A+ ba

    B

    E + i2

    aaA

    aaB E + i

    2 bb

    A bb

    B abA + abB baA + baB . 42

    This expression is exact and it requires no knowledge of

    eigenvalues or eigenvectors. We use it when plotting the

    Greens functions and transmission functions.

    Since the perturbation is imaginary, the perturbed Hamil-

    194704-8 Hansen et al. J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    9/12

    tonian is no longer Hermitian, and the residue of a given

    pole, at i, is no longer the matrix element of an orthogonal

    projection, uiui, as in Eq. 11. A general expression isgiven by

    GbaE = bi

    uivi

    E ia, 43

    where vi is a left eigenvector of the Hamiltonian viH= vii. The perturbed Hamiltonians that we consider arecomplex symmetric matrices, which implies that any left ei-

    genvector is the complex conjugate of the corresponding

    right eigenvector vi = ui, and we can construct the Greensfunction as

    GbaE = bi

    uiui

    E ia. 44

    The details of calculating the Greens functions for ortho,

    meta, and para substitutions are given in the Appendix. Both

    the perturbed and exact Greens functions are shown in Fig.

    8. Overall, the first order perturbation expression is in good

    agreement with the exact result, the largest discrepancies are

    around the degenerate eigenvalues at . With our choice of

    parameters, the ratio / equals 0.4. For a that repro-

    duces the gap between the highest occupied molecular or-

    bital and lowest unoccupied molecular orbital of benzene

    10 eV, this corresponds to a spectral density of 2 eV.An experimental spectral density of the order of 1 eV is

    very large, and we thus conclude that the first order pertur-

    bation employed here is a sensible approach.The Greens functions calculated to first order in are

    given by

    G43orthoE =

    1

    6

    R1o

    E 2+ i/6

    1

    12

    R2o

    E + i/12

    +1

    4

    R3o

    E + i/4+

    1

    12

    R4o

    E + + i/12

    1

    4

    R5o

    E + + i/4

    1

    6

    R6o

    E + 2+ i/6, 45

    G31metaE =

    1

    6

    R1m

    E 2+ i/6

    +1

    12

    R2m

    E + i/12

    1

    4

    R3m

    E + i/4

    +1

    12

    R4m

    E + + i/12

    1

    4

    R5m

    E + + i/4

    +1

    6

    R6m

    E + 2+ i/6, 46

    G52para

    E =

    1

    6

    R1p

    E 2+ i/6

    1

    3

    R2p

    E + i/3

    +1

    3

    R4p

    E + + i/3

    1

    6

    R6p

    E + 2+ i/6. 47

    The numerical factors Rix x = o , m ,p have the property that

    Rix1 for = / 120.

    The transmission functions, TE =2GE2, are de-picted in Fig. 4. In all three cases the widths of the lines at

    2, say, , are given by the shift in the poles, /2

    = /6= /3. No simple notion of linewidth applies to

    the split peaks at in the ortho and meta cases. They are

    dominated by the no longer degenerate levels, whose per-

    turbed energies are split on the imaginary energy axis, whichis why the antiresonances at appear. The slight asym-

    metry of the lineshapes that is exaggerated by the perturba-

    tive treatment stems from the imaginary parts of the eigen-

    vectors. In the para case the width of the resonances at

    equals 2 /3.

    VI. CONCLUSION

    In an attempt to open the black box of quantum transport

    calculations, we have given a fully analytical model treat-

    ment of molecular transmission. We have endeavored to

    probe the relationship between transmission features, elec-

    tronic structure, and chemical understanding. Quite gener-ally, the Greens functions derived from a Hckel Hamil-

    tonian can be expressed as the ratio of the so-called cofactor

    over the secular determinant. The presence of the secular

    determinant in the denominator reiterates the known result

    that the poles in the Greens function coincide with the mo-

    lecular orbital energies. Using the Lwdin partitioning tech-

    nique, we have derived a Dysons equation for the Greens

    functions of cyclic molecules. This in turn leads to a factor-

    ization theorem for the cofactor, which we have used to dis-

    tinguish between two types of zeroes of the Greens func-

    tions: multipath zeroes, which are a result of the interference

    of the different spatial pathways through benzene, and reso-

    TABLE II. Ortho case. Perturbed eigenvalues and eigenvectors and the resi-

    dues of the poles of the Greens function.

    i i

    1Eigenvectors, u

    i

    1 Residue

    1 + 2 i

    6Nou1 + i6u3 + i2

    3u4 16R1o

    2 + i

    12Nou2 i

    23

    u6 + i32

    u5 112R2o

    3 + i

    4Nou3 i6u1 i32 u4

    1

    4R3

    o

    4 i

    12Nou4 i

    23

    u1 + i32

    u3 112R4o

    5 i

    4Nou5 + i6u6 i

    32

    u2 14

    R5o

    6 2 i

    6Nou6 i6u5 + i2

    3u2 16R6o

    194704-9 Interfering pathways in benzene J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    10/12

    nance zeroes, which are a result of the tunneling electron

    being resonant with a substructure of the molecule. Antireso-

    nances induced by sidechains fit into the second category.

    For a simple yet important example molecule, benzene, we

    have given a detailed analysis of the interference patterns

    seen in the transmission functions for the different substitu-

    tion patterns: ortho, meta, and para.

    Some would argue, and rightfully so, that any distinction

    between different flavors of zeroes is somewhat arbitrary. If

    the benzene molecule of our analysis is embedded into alarger molecule, the multipath zeroes of the benzene ring will

    appear as resonance zeroes of this larger molecule. However,

    we believe that the distinction provides valuable insights into

    the interference patterns of benzene and could be a useful

    tool when developing a better chemical understanding of

    quantum interference in molecules. In fact, the distinction

    emphasizes the strong analogy to the different resonator de-

    signs used in meso- and macroscopic systems: double-slit or

    stub resonators.

    To account for the coupling of the metal electrodes we

    introduced using perturbation theory an imaginary potential.

    This caused the transmission resonances to broaden. We cal-culated the eigenenergies and eigenvectors to first order in

    the perturbation. The linewidths of the resonances are given

    by the imaginary shift in energy. In the ortho and meta cases,

    the degenerate eigenstates are split on the imaginary axis,

    which allows the hidden multipath zeroes to emerge.

    In conclusion, we emphasize the usefulness of analytical

    techniques, such as Lwdin partitioning and perturbation

    theory, in analyzing interference in molecular wires. We rec-

    ommend their use for larger model systems than benzene.

    Furthermore, we speculate that the application of Lwdin

    partitioning schemes within the current approach to quantum

    transport calculations could provide valuable insights intothe chemical properties of molecular transport junctions.

    ACKNOWLEDGMENTS

    T.H. thanks Dr. Paul Sherratt for helpful discussions.

    T.H. thanks the Carlsberg Foundation Carlsbergfondet forsupport. We thank the MURI program of the DoD, the NCN

    and MRSEC programs of the NSF, and the ONR and NSF

    Chemistry divisions for support.

    APPENDIX: PERTURBATIVE EIGENVALUES ANDEIGENVECTORS

    Here we consider the coupling to the electrodes as a

    perturbation and calculate the poles and residues of the

    Greens functions to first order in . Remember that the first

    order correction to an eigenenergy i is simply the expecta-

    tion value of the perturbation V. We write the perturbed ei-

    genvalue i

    1as

    i1 = i + uiVui. A1

    The corresponding perturbed eigenvector is given by

    ui1 ui +

    ji

    ujuj

    i jVui , A2

    which can then be normalized. From the perturbed eigenval-

    ues and eigenvectors, we construct the Greens function.

    In the ortho case the two electrodes bind to neighboring

    sites. One could choose sites 1 and 2, but to take advan-tage of the eigenvector symmetry, we choose 3 and 4. Theperturbation operator in the ortho case now reads

    Vortho = i

    233 + 44 . A3

    When the perturbation is rotated to the molecular orbital

    MO basis, Vo = U1VorthoU, it becomes block-diagonal dueto our choice of sites. The three molecular orbitals that are

    even with respect to the symmetry plane of Vortho, namely,

    u1 , u3 , u4, mix under the even submatrix,

    Veveno = i

    2

    1

    62 6 2

    6 3 3 2 3 1

    , A4whereas the three molecular orbitals that are odd with respect

    to the symmetry plane of Vortho

    , namely, u2 , u5 , u6, mixunder the odd submatrix,

    TABLE III. The normalization factors.

    Ortho Meta Para

    No = 1 + 569

    21/2 Nm = 1 + 8936

    21/2 Np = 1 + 89

    21/2

    No = 1 + 274 21/2

    Nm = 1 + 9421/2

    No

    =

    1 +

    35

    362

    1/2

    TABLE IV. The numerical factors Rix, where x = o , m ,p, appearing in the residues.

    i Rio Ri

    m Rip

    1 No21 100 /9 2 i20 /3 Nm

    2 1 121 /362 i 11 /3 Np21 16 /92 i 8 /3

    2 No

    2 1 169 /362 i13 /3 Nm2 1 25 /362 + i5 /3 Np

    21 4 /9 2 i 4 /3

    3 No

    2 1 9 /4 2 + i3 Nm

    2 1 9 /4 2 i3

    4 No

    2 1 169 /362 + i13 /3 Nm2 1 25 /362 i5 /3 Np

    21 4 /9 2 + i 4 /3

    5 No

    2 1 9 /4 2 i3 Nm2 1 9 /4 2 + i3

    6 No21 100 /9 2 + i20 /3 Nm

    2 1 121 /362 + i 11 /3 Np21 16 /92 + i 8 /3

    194704-10 Hansen et al. J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    11/12

    Voddo = i

    216

    1 3 2 3 3 62 6 2

    . A5

    The perturbed eigenvalues and eigenvectors, to first order in

    = / 12, are given in Table II. All normalization con-stants appear in Table III. Also the residues, calculated using

    Eq. 44, are in the table, and the ortho Greens function,appearing in Eq. 45, can be written out. The numericalfactors Ri

    x appear in Table IV.In the meta case we also choose the form of the pertur-

    bation operator to take advantage of eigenvector symmetry.

    We have

    Vmeta = i

    211 + 33 . A6

    This is rotated to the MO basis, Vm = U1VmetaU. Four eigen-

    vectors are even with respect to the symmetry plane of Vmeta.

    In the basis u1 , u2 , u4 , u6 the even submatrix of Vm is

    given by

    Vevenm = i

    2

    1

    6

    2 2 2 2 2 1 1 2 2 1 1 2

    2 2 2 2

    . A7

    Likewise, in the basis u3 , u5 the odd submatrix of Vm is

    given by

    Voddm = i

    2

    1

    6 3 3

    3 3 . A8

    On display in Table V are the perturbed eigenvalues and

    eigenvectors as well as the residues of the poles of the

    Greens function.

    The simplest case is the para case. The perturbation is

    chosen to be

    Vpara = i

    222 + 55 , A9

    and it is rotated to the MO basis, Vp = U1VparaU. The eigen-

    vectors u3 and u5 have nodes at the sites 2 and 5, sothey are not mixed by the perturbation and remain eigenvec-

    tors of the full Hamiltonian. Thus the broadened para

    Greens functions have two poles on the real axis at .

    Two eigenvectors, u1 and u4, are even with respect to the

    symmetry plane of V

    para

    . In the basis u1 , u4 the evensubmatrix of Vp is given by

    Vevenp = i

    2

    1

    6 2 22

    22 4 . A10

    In the basis u2 , u6 the odd submatrix of Vp is given by

    Voddp = i

    2

    1

    6 4 22

    22 2 . A11The perturbed eigenvalues and eigenvectors and the residues

    of the poles of the para Greens function are given in Table

    VI.

    1G. Engel, T. Calhoun, E. Read, T.-K. Ahn, T. Mancal, Y.-C. Cheng, R.

    Blankenship, and G. Fleming, Nature London 446, 782 2007.2

    H. Lee, Y.-C. Cheng, and G. R. Fleming, Science 316, 1462 2007.3

    R. Baer and D. Neuhauser, J. Am. Chem. Soc. 124, 4200 2002.4

    R. Baer and D. Neuhauser, Chem. Phys. 281, 353 2002.5

    D. M. Cardamone, C. A. Stafford, and S. Mazumdar, Nano Lett. 6, 2422

    2006.6

    R. Stadler, M. Forshaw, and C. Joachim, Nanotechnology 14, 138

    2003.7

    R. Stadler, S. Ami, C. Joachim, and M. Forshaw, Nanotechnology 15,

    S115 2004.8

    D. Q. Andrews, G. C. Solomon, R. P. Van Duyne, and M. A. Ratner, J.

    Am. Chem. Soc. 130, 17309 2008.9

    D. Xiao, S. S. Skourtis, I. V. Rubtsov, and D. N. Beratan, Nano Lett. 9,

    1818 2009.10 P. R. Levstein, H. M. Pastawski, and J. L. DAmato, J. Phys.: Condens.

    Matter 2, 1781 1990.11

    E. G. Emberly and G. Kirczenow, J. Phys.: Condens. Matter 11, 6911

    1999.12

    H. W. Lee, Phys. Rev. Lett. 82, 2358 1999.13

    C. Kalyanaraman and D. Evans, Nano Lett. 2, 437 2002.14

    R. Collepardo-Guevara, D. Walter, D. Neuhauser, and R. Baer, Chem.

    Phys. Lett. 393, 367 2004.15

    M. Ernzerhof, M. Zuang, and P. Rocheleau, J. Chem. Phys. 123, 134704

    2005.16

    D. Brisker, I. Cherkes, C. Gnodtke, D. Jarukanont, S. Klaiman, W. Koch,

    S. Weissman, R. Volkovich, M. Toroker, and U. Peskin, Mol. Phys. 106,

    281 2008.17

    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.

    130, 17301 2008.

    TABLE V. Meta case. Perturbed eigenvalues and eigenvectors and the resi-

    dues of the poles of the Greens function.

    i i

    1Eigenvectors, u

    i

    1 Residue

    1 + 2 i

    6Nmu1 + i2u2 + i2

    3u4 + i

    1

    2u6 16R1m

    2 + i

    12Nmu2 i2u1 i2

    3u6 i

    1

    2u4 112R2m

    3 + i

    4Nmu3 + i32 u5

    1

    4R3

    m

    4 i

    12Nmu4 i2u6 i2

    3u1 + i

    1

    2u2 112R4m

    5 i

    4Nmu3 i32 u5

    1

    4R5

    m

    6 2 i

    6Nmu6 + i2u4 + i2

    3u2 i

    1

    2u1 16R6m

    TABLE VI. Para case. Perturbed eigenvalues and eigenvectors and the resi-

    dues of the poles of the Greens function.

    i i

    1Eigenvectors, u

    i

    1 Residue

    1 +2 i /6 Npu1 i22 /3 u4 1 /6R1p

    2 + i /3 Npu2 + i22 /3 u6 1 /3R2p

    3 + u3 0

    4 i /3 Npu4 + i22 /3 u1 1 /3R4p

    5 u5 0

    6 2 i /6 Npu6 i22 /3 u2 1 /6R6p

    194704-11 Interfering pathways in benzene J. Chem. Phys. 131, 194704 2009

    Downloaded 09 Feb 2010 to 129.105.55.216. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

    http://dx.doi.org/10.1038/nature05678http://dx.doi.org/10.1038/nature05678http://dx.doi.org/10.1038/nature05678http://dx.doi.org/10.1038/nature05678http://dx.doi.org/10.1126/science.1142188http://dx.doi.org/10.1021/ja016605shttp://dx.doi.org/10.1016/S0301-0104(02)00570-0http://dx.doi.org/10.1021/nl0608442http://dx.doi.org/10.1088/0957-4484/14/2/307http://dx.doi.org/10.1088/0957-4484/15/4/001http://dx.doi.org/10.1021/ja804399qhttp://dx.doi.org/10.1021/ja804399qhttp://dx.doi.org/10.1021/nl8037695http://dx.doi.org/10.1088/0953-8984/2/7/009http://dx.doi.org/10.1088/0953-8984/2/7/009http://dx.doi.org/10.1088/0953-8984/11/36/308http://dx.doi.org/10.1103/PhysRevLett.82.2358http://dx.doi.org/10.1021/nl010074ohttp://dx.doi.org/10.1016/j.cplett.2004.06.042http://dx.doi.org/10.1016/j.cplett.2004.06.042http://dx.doi.org/10.1063/1.2049249http://dx.doi.org/10.1080/00268970701793904http://dx.doi.org/10.1021/ja8044053http://dx.doi.org/10.1021/ja8044053http://dx.doi.org/10.1080/00268970701793904http://dx.doi.org/10.1063/1.2049249http://dx.doi.org/10.1016/j.cplett.2004.06.042http://dx.doi.org/10.1016/j.cplett.2004.06.042http://dx.doi.org/10.1021/nl010074ohttp://dx.doi.org/10.1103/PhysRevLett.82.2358http://dx.doi.org/10.1088/0953-8984/11/36/308http://dx.doi.org/10.1088/0953-8984/2/7/009http://dx.doi.org/10.1088/0953-8984/2/7/009http://dx.doi.org/10.1021/nl8037695http://dx.doi.org/10.1021/ja804399qhttp://dx.doi.org/10.1021/ja804399qhttp://dx.doi.org/10.1088/0957-4484/15/4/001http://dx.doi.org/10.1088/0957-4484/14/2/307http://dx.doi.org/10.1021/nl0608442http://dx.doi.org/10.1016/S0301-0104(02)00570-0http://dx.doi.org/10.1021/ja016605shttp://dx.doi.org/10.1126/science.1142188http://dx.doi.org/10.1038/nature05678
  • 8/3/2019 Thorsten Hansen et al- Interfering pathways in benzene: An analytical treatment

    12/12

    18G. C. Solomon, D. Q. Andrews, T. Hansen, R. H. Goldsmith, M. R.

    Wasielewski, R. P. Van Duyne, and M. A. Ratner, J. Chem. Phys. 129,

    054701 2008.19

    D. Q. Andrews, G. C. Solomon, R. H. Goldsmith, T. Hansen, M. R.

    Wasielewski, R. P. V. Duyne, and M. A. Ratner, J. Phys. Chem. C 112,

    16991 2008.20

    G. C. Solomon, D. Q. Andrews, R. P. Van Duyne, and M. A. Ratner, J.

    Am. Chem. Soc. 130, 7788 2008.21

    G. C. Solomon, D. Q. Andrews, R. P. V. Duyne, and M. A. Ratner,

    ChemPhysChem 10, 257 2009.22

    B. Pickup and P. Fowler, Chem. Phys. Lett. 459, 198 2008.23 P. Fowler, B. Pickup, and T. Todorova, Chem. Phys. Lett. 465, 142

    2008.24

    P. W. Fowler, B. T. Pickup, T. Z. Todorova, and T. Pisanski, J. Chem.

    Phys. 130, 174708 2009.25

    L.-Y. Hsu and B.-Y. Jin, Chem. Phys. 355, 177 2009.26

    C. Patoux, C. Coudret, J. P. Launay, C. Joachim, and A. Gourdon, Inorg.

    Chem. 36, 5037 1997.27

    M. Mayor, H. Weber, J. Reichert, M. Elbing, C. von Hnisch, D. Beck-

    mann, and M. Fischer, Angew. Chem., Int. Ed. 42, 5834 2003.28

    P. Sautet and C. Joachim, Chem. Phys. Lett. 153, 511 1988.29

    P. Sautet and C. Joachim, Chem. Phys. 135, 99 1989.30

    C. Joachim, J. Gimzewski, and H. Tang, Phys. Rev. B 58, 16407 1998.31

    S. Yaliraki and M. Ratner, Ann. N.Y. Acad. Sci. 960, 153 2002.32

    D. Walter, D. Neuhauser, and R. Baer, Chem. Phys. 299, 139 2004.33

    M. Ernzerhof, H. Bahmann, F. Goyer, M. Zuang, and P. Rocheleau, J.

    Chem. Theory Comput. 2, 1291 2006.34 T. A. Papadopoulos, I. M. Grace, and C. J. Lambert, Phys. Rev. B 74,

    193306 2006.35

    S. K. Maiti, Chem. Phys. 331, 254 2007.36

    K. Yoshizawa, T. Tada, and A. Staykov, J. Am. Chem. Soc. 130, 9406

    2008.37

    S.-H. Ke, W. Yang, and H. U. Baranger, Nano Lett. 8, 3257 2008.38

    D. Darau, G. Begemann, A. Donarini, and M. Grifoni, Phys. Rev. B 79,

    235404 2009.39

    M. A. Ratner, J. Phys. Chem. 94, 4877 1990.40

    S. S. Skourtis and D. N. Beratan, Theories of Structure-Function Rela-

    tionship for Bridge-Mediated Electron Transfer, Advances in Chemical

    Physics, Vol. 106 Wiley, New York, 1999, Chap. 8, pp. 377452.41

    C. Goldman, Phys. Rev. A 43, 4500 1991.42

    Y. Magarshak, J. Malinsky, and A. D. Joran, J. Chem. Phys. 95, 418

    1991.43

    O. L. de Santana and A. A. S. da Gama, Chem. Phys. Lett. 314, 508

    1999.44

    S. Datta, Quantum Transport: Atom to Transistor Cambridge UniversityPress, Cambridge, 2005.

    45S. Datta, Phys. Rev. B 40, 5830 1989.

    46Y. Meir and N. S. Wingreen, Phys. Rev. Lett. 68, 2512 1992.

    47 W. Tian, S. Datta, S. Hong, R. Reifenberger, J. I. Henderson, and C. P.

    Kubiak, J. Chem. Phys. 109, 2874 1998.48

    M. P. Samanta, W. Tian, S. Datta, J. I. Henderson, and C. P. Kubiak,

    Phys. Rev. B 53, R7626 1996.49

    L. E. Hall, J. R. Reimers, N. S. Hush, and K. Silverbrook, J. Chem. Phys.

    112, 1510 2000.50

    J. Taylor, H. Guo, and J. Wang, Phys. Rev. B 63, 245407 2001.51

    M. Brandbyge, J.-L. Mozos, P. Ordejon, J. Taylor, and K. Stokbro, Phys.

    Rev. B 65, 165401 2002.52

    Y. Xue, S. Datta, and M. A. Ratner, Chem. Phys. 281, 151 2002.53

    D. S. Fisher and P. A. Lee, Phys. Rev. B 23, 6851 1981.54

    T. N. Todorov, J. Phys.: Condens. Matter 14, 3049 2002.55

    P. Lancaster and M. Tismenetsky, The Theory of Matrices, 2nd ed. Aca-demic, San Diego, 1985.

    56P.-O. Lwdin, J. Math. Phys. 3, 969 1962.

    57C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom-Photon In-

    teractions Wiley, New York, 1992.58

    A. Messiah, Quantum Mechanics Dover, Mineola, 1999.59

    P. W. Anderson, Phys. Rev. 109, 1492 1958.60

    E. Economou, Greens Functions in Quantum Physics, 3rd ed. Springer,New York, 2006.

    61E. V. Anslyn and D. A. Dougherty, Modern Physical Organic Chemistry

    University Science Books, Sausalito, 2005.62

    C. Coulson and G. Rushbrooke, Proc. Cambridge Philos. Soc. 36, 193

    1940.63

    N. Trinajstic, Chemical Graph Theory, 2nd ed. CRC, Boca Raton, 1992.

    194704-12 Hansen et al. J. Chem. Phys. 131, 194704 2009

    http://dx.doi.org/10.1063/1.2958275http://dx.doi.org/10.1021/jp805588mhttp://dx.doi.org/10.1021/ja801379bhttp://dx.doi.org/10.1021/ja801379bhttp://dx.doi.org/10.1002/cphc.200800591http://dx.doi.org/10.1016/j.cplett.2008.05.062http://dx.doi.org/10.1016/j.cplett.2008.09.048http://dx.doi.org/10.1063/1.3124828http://dx.doi.org/10.1063/1.3124828http://dx.doi.org/10.1016/j.chemphys.2008.12.015http://dx.doi.org/10.1021/ic970013mhttp://dx.doi.org/10.1021/ic970013mhttp://dx.doi.org/10.1002/anie.200352179http://dx.doi.org/10.1016/0009-2614(88)85252-7http://dx.doi.org/10.1016/0301-0104(89)87009-0http://dx.doi.org/10.1103/PhysRevB.58.16407http://dx.doi.org/10.1016/j.chemphys.2003.12.015http://dx.doi.org/10.1021/ct600087chttp://dx.doi.org/10.1021/ct600087chttp://dx.doi.org/10.1103/PhysRevB.74.193306http://dx.doi.org/10.1016/j.chemphys.2006.10.013http://dx.doi.org/10.1021/ja800638thttp://dx.doi.org/10.1021/nl8016175http://dx.doi.org/10.1103/PhysRevB.79.235404http://dx.doi.org/10.1021/j100375a024http://dx.doi.org/10.1103/PhysRevA.43.4500http://dx.doi.org/10.1063/1.461443http://dx.doi.org/10.1016/S0009-2614(99)01183-5http://dx.doi.org/10.1103/PhysRevB.40.5830http://dx.doi.org/10.1103/PhysRevLett.68.2512http://dx.doi.org/10.1063/1.476841http://dx.doi.org/10.1103/PhysRevB.53.R7626http://dx.doi.org/10.1063/1.480696http://dx.doi.org/10.1103/PhysRevB.63.245407http://dx.doi.org/10.1103/PhysRevB.65.165401http://dx.doi.org/10.1103/PhysRevB.65.165401http://dx.doi.org/10.1016/S0301-0104(02)00446-9http://dx.doi.org/10.1103/PhysRevB.23.6851http://dx.doi.org/10.1088/0953-8984/14/11/314http://dx.doi.org/10.1063/1.1724312http://dx.doi.org/10.1103/PhysRev.109.1492http://dx.doi.org/10.1017/S0305004100017163http://dx.doi.org/10.1017/S0305004100017163http://dx.doi.org/10.1103/PhysRev.109.1492http://dx.doi.org/10.1063/1.1724312http://dx.doi.org/10.1088/0953-8984/14/11/314http://dx.doi.org/10.1103/PhysRevB.23.6851http://dx.doi.org/10.1016/S0301-0104(02)00446-9http://dx.doi.org/10.1103/PhysRevB.65.165401http://dx.doi.org/10.1103/PhysRevB.65.165401http://dx.doi.org/10.1103/PhysRevB.63.245407http://dx.doi.org/10.1063/1.480696http://dx.doi.org/10.1103/PhysRevB.53.R7626http://dx.doi.org/10.1063/1.476841http://dx.doi.org/10.1103/PhysRevLett.68.2512http://dx.doi.org/10.1103/PhysRevB.40.5830http://dx.doi.org/10.1016/S0009-2614(99)01183-5http://dx.doi.org/10.1063/1.461443http://dx.doi.org/10.1103/PhysRevA.43.4500http://dx.doi.org/10.1021/j100375a024http://dx.doi.org/10.1103/PhysRevB.79.235404http://dx.doi.org/10.1021/nl8016175http://dx.doi.org/10.1021/ja800638thttp://dx.doi.org/10.1016/j.chemphys.2006.10.013http://dx.doi.org/10.1103/PhysRevB.74.193306http://dx.doi.org/10.1021/ct600087chttp://dx.doi.org/10.1021/ct600087chttp://dx.doi.org/10.1016/j.chemphys.2003.12.015http://dx.doi.org/10.1103/PhysRevB.58.16407http://dx.doi.org/10.1016/0301-0104(89)87009-0http://dx.doi.org/10.1016/0009-2614(88)85252-7http://dx.doi.org/10.1002/anie.200352179http://dx.doi.org/10.1021/ic970013mhttp://dx.doi.org/10.1021/ic970013mhttp://dx.doi.org/10.1016/j.chemphys.2008.12.015http://dx.doi.org/10.1063/1.3124828http://dx.doi.org/10.1063/1.3124828http://dx.doi.org/10.1016/j.cplett.2008.09.048http://dx.doi.org/10.1016/j.cplett.2008.05.062http://dx.doi.org/10.1002/cphc.200800591http://dx.doi.org/10.1021/ja801379bhttp://dx.doi.org/10.1021/ja801379bhttp://dx.doi.org/10.1021/jp805588mhttp://dx.doi.org/10.1063/1.2958275