• Aucun résultat trouvé

Dynamic current suppression and gate voltage response in metal-molecule-metal junctions

N/A
N/A
Protected

Academic year: 2021

Partager "Dynamic current suppression and gate voltage response in metal-molecule-metal junctions"

Copied!
13
0
0

Texte intégral

(1)

Dynamic current suppression and gate voltage

response in metal-molecule-metal junctions

The MIT Faculty has made this article openly available. Please share how this access benefits you. Your story matters.

Citation Evans, Jeremy S., and Troy Van Voorhis. “Dynamic Current

Suppression and Gate Voltage Response in Metal−Molecule−Metal Junctions.” Nano Letters 9.7 (2009): 2671–2675.

As Published http://dx.doi.org/10.1021/nl9011134

Publisher American Chemical Society

Version Author's final manuscript

Citable link http://hdl.handle.net/1721.1/69836

Terms of Use Article is made available in accordance with the publisher's

policy and may be subject to US copyright law. Please refer to the publisher's site for terms of use.

(2)

Dynamic current suppression and gate voltage response in

metal-molecule-metal junctions

Jeremy S. Evans and Troy Van Voorhis

Department of Chemistry Massachusetts Institute of Technology

77 Massachusetts Ave. Cambridge, MA 02139 USA

Abstract

We critically re-examine conductance in benzenedithiol(BDT)/gold junctions using real-time DFT simulations. Our results indicate a powerful influence of the BDT molecular charge on current, with negative charge suppressing electron transport. This effect occurs dynamically as the BDT charge and current oscillate on the femtosecond timescale, indicating that a steady state picture may not be appropriate for this single molecule conducting device. Further, we exploit this effect to show that a gate voltage can be used to indirectly control the device current by adjusting the molecular charge. Thus, it appears that transport in even this simple molecular junction involves a level of sophistication not heretofore recognized.

(3)

The conductance behavior of molecular devices subjected to an external potential has been the source of significant theoretical interest since the first experiments were performed some time ago1,2. As one of the earliest and simplest single-molecule conductors to be studied, benzene dithiol (BDT) has become the de facto standard model system for both experimental1,3 and theoretical4–8 studies. In this work, we use a recently proposed real-time density functional theory (DFT) propagation technique9,10 to study several properties of electron transport through BDT at low bias.

When examining conductance through systems, such as organic molecules, in which elec-tronic motion is hindered relative to metals, one must consider the Coulomb blockade (CB). First used to describe transport through quantum dots11, CB refers to the resistance caused by the energetic cost of adding or removing an electron from an already charged molecule12,13. As a result of this energy cost, conductance is especially sensitive to the charge on the molec-ular device in the CB regime4–6,14.

Closely related to molecular charge and CB is the application of a gate potential. While the current is driven by a voltage applied within the metallic leads, a gate voltage is ap-plied to the conducting device to alter its conductance properties. By shifting the energies of the molecule states, the gate voltage can induce charging of the molecule making gat-ing a natural method to study CB effects. Sgat-ingle-molecule gate voltage experiments have been performed3,13,15, however the difficulty in constructing three terminal devices using a molecule as small as BDT makes obtaining reliable results difficult. Theoretical studies of gating effects are not thus hindered16,17 , so that simulations have significant flexibility in studying CB physics.

The most popular theoretical method to examine molecular conductance is the nonequi-librium Green’s function (NEGF)7,18,19method in which the Hamiltonian is used to generate Landauer-type20conductance channels. Now, the NEGF method assumes a steady state con-duction picture. Aside from being potentially inaccurate in the presence of molecular bound states21, the steady state picture ignores the interior life of the junction: How and (how quickly) does the current build up? How many electrons are moving through the molecule at one time? In order to address these questions, one can instead use time-dependent den-sity functional theory (TDDFT) to model dynamic transport in molecular junctions22. Here, one applies a time-dependent potential to drive current in a fully nonlinear real-time prop-agation of a microcanonical system23. We have previously applied this method to simple

(4)

1D conjugated carbon wires and have explored the conductance behaviors of both closed-shell9 and open-shell10 systems. Here we apply our real-time DFT propagation method to the gold-BDT-gold system (Fig. 1) in order to better understand the dynamics of trans-port in a real junction. We calculate the dynamic current-voltage response using several exchange-correlation functionals and observe a strong negative correlation between the in-stantaneous charge on BDT and the current through the junction - the dynamic equivalent of CB. Further, we find that our currents oscillate on the femtosecond timescale, calling in to question the coherent steady-state picture invoked in the NEGF approach. Finally, we calculate currents under a gate voltage and observe that Vg provides indirect control over the current by modifying the BDT charge. Taken together, our results suggest a much more in-tricate picture of transport through this junction - involving dynamic fluctuations between nonequilibrium states and strong electron-electron interactions - that has not previously been appreciated.

The system geometry under study is a BDT molecule covalently bonded to the 111 face between two Au114 clusters in the fcc position. We construct the geometry by reflection symmetry on portions of a previously optimized geometry24 for phenylthiol chemisorbed to a gold surface. We replicate the gold layers in the 111 reciprocal direction to produce an elongated wire structure. The system geometry is included as supplementary data. We choose a structure elongated in the direction of electron transport (Z) to allow sufficient density of momentum states in the Z-direction and sufficient time before reflection at the wire ends manifests in the current. The electrons are described by the Hay-Wadt pseudopotential and minimal basis25 on the gold atoms and the Aldrichs VDZ basis26augmented with heavy atom d functions on all other atoms. Calculations were performed using a development version of Q-Chem27. We divide our system into an electron source, molecular device, and electron drain region as shown in Fig. 1a. We will use this system division to apply a voltage potential and to define the resulting time-dependent electron densities.

The real-time DFT propagation method used to calculate conductance properties have been described previously9,10. Here we summarize the closed-shell method and its appli-cation in this work. We choose an unperturbed Hamiltonian and determine our initial state as the ground state of the entire system under that Hamiltonian. At time t = 0, we change the Hamiltonian and propagate the electron density, ρ(t), according to the time-dependent Kohn-Sham (KS) method under the perturbed Hamiltonian28. We construct the

(5)

-1.2 -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0 0.5 1 1.5 2 NS -N D time (fs) b 0.544 V 1.09 V 1.63 V 2.18 V 2.72 V Fits 0 5 10 15 20 25 30 35 40 45 0 0.5 1 1.5 2 2.5 3 Current ( µ A) Voltage (V) c

FIG. 1: (a) Au114-BDT-Au114 system with source and lead regions labeled along with the effect of lead potential, V , and gate potential, Vg, on electronic energies in each region. Atoms depicted

include H (gray), C (light blue), S (green), and Au (gold) (b) Time dependent density difference (Ntot) with linear fitting in the steady state region and (c) resulting current-voltage curve.

KS Hamiltonians using readily available functionals, and use the same numerical propagation techniques as our previous work.

Using a voltage potential definition, we choose our initial Hamiltonian according to un-biased ground state DFT. We add an external potential at t = 0 to determine the adjusted Hamiltonian. The potential is defined using the population operators on the source (ˆnS), molecular device (ˆnM), and drain (ˆnD). We follow our previous work and choose our popu-lation operators according to the L¨owdin population definition29. For a lead bias of V , the adjusted KS Hamiltonian is given by ˆHKS[ρ] +V2nˆS−V2ˆnD where ˆHKS[ρ] is the unperturbed KS Hamiltonian. We include a gate voltage with the additional term VgnˆM. The effect of

(6)

0 5 10 15 20 25 30 35 40 45 0 0.5 1 1.5 2 2.5 3 Current ( µ A) Voltage (V) LDA B3LYP HF

FIG. 2: TDDFT calculated current-voltage curve using B3LYP, LDA, and Hartree-Fock exchange-correlation functionals.

the lead and gate voltages on the electronic state energies in each region is shown in Fig 1a. This method allows us to calculate the time dependent electron populations in the source, molecular device, and drain using the population operators defined above:

NR(t) = 

nR(r)ρ(t)dr R = S, M, D. (1)

From these values, we define the current (I = 12d(ND(t)−NS(t))

dt ) as the least-squares linear regression slope in 12(ND − NS) over the time period of constant charge transport that begins shortly after application of the potential. In this work, the current is calculated over the time period of 0.24 to 1.57 fs. As an example, the regression fitting and resulting IV plot for the LDA functional are shown in Fig. 1b and c.

Using these methods, we have generated small bias IV curves with the LDA, B3LYP, and Hartree-Fock functionals. These results are shown in Fig. 2. We note that the calculated IV curves are mostly linear over the low-bias regime shown. For Hartree-Fock, the average conductance over the calculated range is 0, indicating that this voltage range lies within the HF predicted conductance gap for the system. On the other hand, both B3LYP and LDA predict no conductance gap at the plot resolution of Fig. 2. A decrease in the conductance gap with methods that do not include 100% exact exchange has been previously reported30,31. The average conductances, calculated from the largest plotted voltage point are 16 µS with LDA, and 10 µS with B3LYP. For comparison, the conductance quantum is 77 µS and the contemporary experimental estimates of the conductance of BDT are approximately 1 µS.3 This order of magnitude overestimation of current over experiment is a common shortcoming

(7)

of conduction calculations19,32.

Although the difference in conductance between these methods has been associated with a decrease in the device HOMO-LUMO gap with the introduction of self-interaction error30,33,34, we also notice a correlation with the density distribution in the equilibrium state. Using the L¨owdin population definition, the total charge on the molecule is -1.38 with Hartree-Fock, -0.955 with B3LYP, and -0.851 with LDA. Thus, a more negative electronic charge on the molecular device correlates with a decrease in its conductance. This behavior resembles the Coulomb blockade described earlier in that the additional Coulombic energy cost from increased electronic density correlates with decreased conductance.

We can estimate the Coulomb energy associated with the charge differences if we simplify the molecular device to be a sphere of uniform charge density. In that case, the Coulombic repulsion energy is EC = 20π3q2

0R where q is the total charge, and R is the radius. We choose the radius to be half the distance between the sulfur atoms in the BDT device, so R = 3.2 ˚A. The resulting energy difference between the LDA and HF calculated charges is 3.2eV. This energy is certainly sufficient to push the conduction states outside of the voltage window. Furthermore, it is likely that charging the device to -1 or below saturates the conduction bands. Thus it is reasonable that the difference in equilibrium molecular charge predicted by the various functionals accounts for some of the differences in conduction behavior.

We generate the IV curves from relatively short time dynamics (t < 1.75 fs) in which the current is approximately steady, but we find significantly different behavior at longer times. In Fig. 3, we display real-time results for the junction over a larger time window under several external potentials. The dynamic data indicate that the electron density transport from the source to the drain does not occur at a constant rate. All three of the propagations include regions in which the time derivative of ND−NS, and therefore the transient current, is 0 or even negative. The complex dynamics are a direct result of the non-equilibrium initial condition we have set up, and as such the precise oscillations will reflect the details of how the system is initialized23. For example, if the Bias is turned on over the first 1 fs (as opposed to instantaneously at t=0) the oscillations persist, although their magnitude and phase are somewhat altered. Further, we see similar current oscillations with B3LYP, but because these long-time dynamics are computationally intensive, we choose to focus on the LDA results with the expectation that the B3LYP results will be qualitatively similar. Qualitative similarities in the IV curve outside of the conductance gap with the various

(8)

0 0.5 1 1.5 0 2 4 6 8 10 ND -N S time (fs) a 0.544 V 1.09 V 2.18 V -0.9 -0.88 -0.86 -0.84 -0.82 -0.8 0 2 4 6 8 10 Charge (q e ) time (fs) b 0.544 V 1.09 V 2.18 V

FIG. 3: Time dependent density difference (Ntot) (a) and device charge (b) as calculated by LDA.

The vertical lines are guides to the eye that indicate approximate times at which the first zero current regions begin for each potential.

functionals30,31 suggests we may see similar results with HF as well.

In addition to the current oscillations, the long time simulations also demonstrate oscil-lations in the density on the molecule. These current and density osciloscil-lations demonstrate that the system does not evolve towards a steady state, but instead show oscillatory behavior consistent with the presence of a bound molecular state in the voltage gap21. Although this suggests that the oscillations represent resonance with molecular states, energies associated with the large period oscillations are smaller than the HOMO-LUMO gap of the bare molec-ular device. Additionally, their periods are sensitive to the applied potential (cf. Fig. 3) , so the states corresponding to these oscillations must have a significant component on the metallic leads.

(9)

effect and that the periods of zero current may be caused by either: 1) saturation of the finite leads or 2) reflection off the ends of the wire. We discount the first possibility, because the positive current always reappears and transports more electrons to the drain after the plateau. We examined the second possibility by performing test calculations using shorter gold wires and found that the oscillation frequencies do not show a systematic change with wire length. Thus, the oscillations would not seem to be a finite lead effect.

The dynamics in Fig. 3 indicate a correlation of a period of zero current with increased negative molecular device charge. The first period of zero current begins at 8.8 fs for the 0.544 V propagation, 4.0 fs for 1.09V , and 1.9 fs for 2.18V. At all of these points, the molecular charge profile indicates a region of increased negative charge. Although the 0.03 electron variation in charge is small, the uniform spherical charge approximation discussed above leads to a Coulombic energy variation of approximately 0.15 eV, which is significant relative to the voltages considered. Thus, we see a dynamic as well as static association between device charge and current, indicating a dynamic equivalent of Coulomb blockade. It is somewhat surprising that something similar to CB physics can be obtained from semilocal TDDFT, which is usually only able to describe mean-field-like evolution accurately. The key in this case is that while semilocal functionals neglect explicit electron correlation in

space - so that static CB effects are absent - they recover correlation in time - so that the

dynamic oscillations remain.

We note that the correlation observed between molecular charge and current is opposite to the previously reported5,6 tendency of a more negative charge to increase current. This calculation, unlike the others reported results, does not involve changes in lead-device sep-aration which may impact current. Our results may implicate a LUMO-based transport pathway, which would cause current to decrease with a more negatively charged device.

The oscillatory current behavior shown in Fig. 3 is relevant to the NEGF formalism, because the NEGF formalism is based upon a steady state assumption. The long time calculations indicate that the system reaches an oscillatory, not steady, state. Our results indicate that some of the overestimation common to NEGF simulations19 may be due to this steady state assumption. To this point, we have computed net currents from the early (e.g. 1 fs) steady state period during which the density fits well into the NEGF picture. However, the average current over longer times (e.g. 10 fs) is clearly lower than the initial current because of the oscillations and one might argue that this is the more relevant timescale for

(10)

26 28 30 32 34 36 38 40 42 44 -1 -0.5 0 0.5 1 Current ( µ A) Gate Voltage (V)

FIG. 4: TDDFT calculated current as a function of gate voltage with left-right voltage set to 1.09 V and calculated with LDA

the experiment. If the development of non-equilibrium oscillations is a general feature of molecular junctions, using the long-time current would thus systematically reduce the cal-culated conductance and improve the agreement of theory and experiment. The oscillations make precise calculations impossible, but the results in Fig. 3 suggest that the new definition would lead to a two- to three-fold improvement in calculated conductance for the BDT-gold junction. We believe this observation is significant and may play a role in bridging the gap between theory and experiment.

The association between charge and current suggests that we may be able to alter the current by inducing a change in the device charge . To that end, we introduce a gate voltage to the system under an external bias35. Based upon the definition of gate voltage in Fig. 1a, we expect a positive potential to produce a less negative charge on the molecular device.

We show in Fig. 4 the early time current with a fixed bias of 1.09 V as a function of gate voltage. As we did for external potentials, we focus on small gate voltages to avoid overwhelming the chosen basis. Examination of the charge dynamics (not shown) verifies that the linear fit time period for current calculation occurs after the molecular charge has equilibrated in response to the gate voltage. The results indicate that by increasing the gate voltage, and thereby making the device charge less negative, we increase the current. We calculate a low voltage response of current to gate voltage of 5.8 µA/V, or 36% of the response to external voltage.

This simulation suggests that we can control the current through BDT with the applica-tion of a gate voltage. As a positive gate voltage creates a less negative device charge, the

(11)

current increases. The device charge changes at a rate of 7.5 qe/V leading to a current to charge response of 0.77 µA/qe. This is clearly much less than the response demonstrated in the dynamic results. However the impact of charge on current is significantly counteracted by the gate voltage required to generate the charge response.

In conclusion, our real-time simulations of the Au114-BDT-Au114 cluster suggest a picture where oscillations in charge drive transport, in contrast to the standard steady-state picture of electron transport. At a given instant, a reduction in current is associated with a more negative charge on BDT. This current suppression a dynamic effect similar to the Coulomb blockade, arising out of the temporal (rather than spatial) correlation between electrons. We further demonstrate that a gate voltage can be used to induce molecular charging, and thus impact the current by modulating the dynamic charge-current suppression. Moving forward, it should be noted that our results neglect relaxation of the nuclear framework. As these motions occur on the same timescale as the slowest oscillations shown in Fig. 3, the potential for vibronic interactions cannot be neglected and future work will examine the interplay between these slow oscillations and vibrational motion. Thus it would seem that time offers a new dimension for unravelling the rich mechanisms at work in molecular transport.

Acknowledgments

T. V. gratefully acknowledges support from an NSF CAREER award (CHE-0547877) and a Packard Fellowship.

1 M. A. Reed, C. Zhou, C. J. Muller, T. P. Burgin, and J. M. Tour, Science 278, 252 (1997). 2 M. Dorogi, J. Gomez, R. Osifchin, R. P. Andres, and R. Reifenberger, Physical Review B 52,

9071 (1995).

3 X. Li, B. Xu, X. Xiao, X. Yang, L. Zang, and N. Tao, Faraday Discussions 131, 111 (2006). 4 Y. Xue, S. Datta, and M. A. Ratner, The Journal of Chemical Physics 115, 4292 (2001). 5 S. Ke, H. U. Baranger, and W. Yang, The Journal of Chemical Physics 122, 074704 (2005). 6 D. Q. Andrews, R. Cohen, R. P. V. Duyne, and M. A. Ratner, The Journal of Chemical Physics

(12)

7 P. Damle, A. W. Ghosh, and S. Datta, Chemical Physics 281, 171 (2002).

8 L. E. Hall, J. R. Reimers, N. S. Hush, and K. Silverbrook, The Journal of Chemical Physics

112, 1510 (2000).

9 C. Cheng, J. S. Evans, and T. V. Voorhis, Physical Review B (Condensed Matter and Materials

Physics) 74, 155112 (2006).

10 J. S. Evans, C. Cheng, and T. V. Voorhis, Physical Review B (Condensed Matter and Materials

Physics) 78, 165108 (2008).

11 C. Livermore, C. H. Crouch, R. M. Westervelt, K. L. Campman, and A. C. Gossard, Science

274, 1332 (1996).

12 R. P. Andres, T. Bein, M. Dorogi, S. Feng, J. I. Henderson, C. P. Kubiak, W. Mahoney, R. G.

Osifchin, and R. Reifenberger, Science 272, 1323 (1996).

13 H. Park, J. Park, A. K. L. Lim, E. H. Anderson, A. P. Alivisatos, and P. L. McEuen, Nature

407, 57 (2000), ISSN 0028-0836.

14 Y. Xue and M. A. Ratner, Physical Review B 68, 115406 (2003).

15 J. O. Lee, G. Lientschnig, F. Wiertz, M. Struijk, R. A. J. Janssen, R. Egberink, D. N. Reinhoudt,

P. Hadley, and C. Dekker, Nano Letters 3, 113 (2003).

16 W. D. Wheeler and Y. Dahnovsky, The Journal of Chemical Physics 129, 154112 (2008). 17 A. W. Ghosh, T. Rakshit, and S. Datta, Nano Letters 4, 565 (2004).

18 S. Datta, Superlattices and Microstructures 28, 253 (2000).

19 M. Koentopp, C. Chang, K. Burke, and R. Car, Journal of Physics: Condensed Matter 20,

083203 (2008), ISSN 0953-8984.

20 M. Buttiker, Y. Imry, R. Landauer, and S. Pinhas, Physical Review B 31, 6207 (1985).

21 G. Stefanucci, Physical Review B (Condensed Matter and Materials Physics) 75, 195115 (2007). 22 G. Stefanucci and C. Almbladh, EPL (Europhysics Letters) 67, 14 (2004), ISSN 0295-5075. 23 M. D. Ventra and T. N. Todorov, Journal of Physics: Condensed Matter 16, 8025 (2004), ISSN

0953-8984.

24 A. Bilic, J. R. Reimers, and N. S. Hush, The Journal of Chemical Physics 122, 094708 (2005). 25 P. J. Hay and W. R. Wadt, The Journal of Chemical Physics 82, 299 (1985).

26 A. Schafer, H. Horn, and R. Ahlrichs, The Journal of Chemical Physics 97, 2571 (1992). 27 Y. Shao, L. F. Molnar, Y. Jung, J. Kussmann, C. Ochsenfeld, S. T. Brown, A. T. Gilbert, L. V.

(13)

(2006).

28 E. Runge and E. K. U. Gross, Physical Review Letters 52, 997 (1984). 29 P. Lowdin, The Journal of Chemical Physics 18, 365 (1950).

30 S. Ke, H. U. Baranger, and W. Yang, The Journal of Chemical Physics 126, 201102 (2007). 31 J. S. Evans, O. Vydrov, and T. V. Voorhis, manuscript in preparation.

32 M. Di Ventra, S. T. Pantelides, and N. D. Lang, Phys. Rev. Lett. 84, 979 (2000).

33 M. Koentopp, K. Burke, and F. Evers, Physical Review B (Condensed Matter and Materials

Physics) 73, 121403 (2006).

34 N. Sai, M. Zwolak, G. Vignale, and M. D. Ventra, Physical Review Letters 94, 186810 (2005). 35 M. D. Ventra, S. T. Pantelides, and N. D. Lang, Applied Physics Letters 76, 3448 (2000).

Figure

FIG. 1: (a) Au 114 -BDT-Au 114 system with source and lead regions labeled along with the effect of lead potential, V , and gate potential, V g , on electronic energies in each region
FIG. 2: TDDFT calculated current-voltage curve using B3LYP, LDA, and Hartree-Fock exchange- exchange-correlation functionals.
FIG. 3: Time dependent density difference ( N tot ) (a) and device charge (b) as calculated by LDA.
FIG. 4: TDDFT calculated current as a function of gate voltage with left-right voltage set to 1.09 V and calculated with LDA

Références

Documents relatifs

Metal-oxide polymer diodes exhibit non-volatile resistive switching. The current- voltage characteristics have been studied as function of temperature. The low- conductance

Abstract.- Structure has been observed in the background signal of d2J/dv2 characteristics of metal point contacts, implying double-phonon backflow-processes.. An explanation of

Hence, groundwater quality in near-river aquifers of restored river reaches could vary markedly because of the high spatial and temporal variability of both infiltration travel

Keywords : metal powders, Mechanical Alloying, Magnetic Properties, Eddy Current, NDT, nondestructive testing, process monitoring, nanocrystalline materials, planetary ball

Analysis of the data showed that the sensor output was in relation with different parameter of powder (diameter, density and structural) that the sensitivity of the sensor

The Tropical Plant Biological Resource Center of the French West Indies is eager up to provide plant material to end users throughout the Caribbean region, in order to help

Secondly, in a high oxidative stress condition the highest concentration of tea extracts, i.e., a sufficient amount of antioxidant compounds (about 100 µg of total polyphenols per

The influence of gold precursor concentration, light intensity and nature of the solvents on the formation and properties of gold nanoparticles (size, dispersion, stability) on