You are on page 1of 6

LETTERS

PUBLISHED ONLINE: 6 JANUARY 2013 | DOI: 10.1038/NNANO.2012.238

The role of van der Waals forces in the


performance of molecular diodes
Nisachol Nerngchamnong1†, Li Yuan1†, Dong-Chen Qi2,3, Jiang Li1, Damien Thompson4 *
and Christian A. Nijhuis1,5 *

One of the main goals of organic and molecular electronics is to transport, because it yields junctions with high yield in working
relate the performance and electronic function of devices to the devices (up to 100%), statistically large amounts of data, and is com-
chemical structure and intermolecular interactions of the patible with a wide variety of SAMs18–21. Figure 1a presents an
organic component inside them, which can take the form of optical micrograph of a cone-shaped tip of Ga2O3/EGaIn in
an organic thin film, a self-assembled monolayer or a single contact with a ferrocene-alkanethiolate SAM of S(CH2)nFc or
molecule1–7. This goal is difficult to achieve because organic SCnFc (n ¼ 6–15; Fc ¼ ferrocene) on a AgTS bottom electrode,
and molecular electronic devices are complex physical– together with schematic illustrations of ‘ideal’ junctions with
organic systems that consist of at least two electrodes, an SAMs of SC10Fc (right) and SC11Fc (left). The SAMs of SCnFc are
organic component and two (different) organic/inorganic inter- covalently bound to the AgTS bottom electrodes by Ag–S bonds,
faces. Singling out the contribution of each of these com- and form van der Waals contact with the Ga2O3/EGaIn top electro-
ponents remains challenging. So far, strong p–p interactions des. This figure also defines the tilt angles a of the Fc units with
have mainly been considered for the rational design and optim- respect to the plane perpendicular to the surface.
ization of the performances of organic electronic devices8–10, The junctions with SAMs of SC11Fc rectify currents with a
and weaker intermolecular interactions have largely been log-average rectification ratio (R ; |J(21.0 V)|/|J(þ1.0 V)|) of
ignored. Here, we show experimentally that subtle changes in 1.5 × 102 with a log standard deviation of 2.1. Figure 1b shows
the intermolecular van der Waals interactions in the active the energy level diagrams when a bias of 21.0 or 1.0 V is applied
component of a molecular diode dramatically impact the per- across these molecular diodes (defined here as forward and
formance of the device. In particular, we observe an odd–even reverse bias, respectively, in analogy to conventional diodes).
effect as the number of alkyl units is varied in a ferrocene– Temperature-dependent measurements indicated that the mechan-
alkanethiolate self-assembled monolayer. As a result of a ism of charge transport changes from tunnelling in one direction of
more favourable van der Waals interaction, junctions made bias to tunnelling followed by hopping in the other21. With positive
from an odd number of alkyl units have a lower packing bias, the highest occupied molecular orbital (HOMO) of the mol-
energy (by ∼0.4–0.6 kcal mol–1), rectify currents 10 times ecules does not fall within the energy window of the Fermi levels
more efficiently, give a 10% higher yield in working devices, of the electrodes, so charges cannot be injected into the HOMO;
and can be made two to three times more reproducibly than both the Fc units, dFc , and the alkyl spacer, d Cn , form a barrier
junctions made from an even number of alkyl units. against tunnelling with a width of dtot,reverse ¼ dFc þ d Cn . At negative
In general, small changes in the chemical structures of molecules bias, the HOMO of the molecules falls between the Fermi levels of
can result in large changes in the overall interaction energy between the electrodes, so charge can hop to the Fc units and, consequently,
them and, consequently, in large changes in macroscopic structure reduce the width of the tunnelling barrier to dtot,forward ¼ d Cn . This
and orientation11. This effect is crucial in many applications, includ- reduction in the width of the tunnelling barrier only at forward bias
ing crystal engineering12, drug design13 and organic electronics14. results in the large observed current rectification.
The importance of weak intermolecular interactions is evident We formed junctions with SAMs of SCnFc and measured statisti-
from studies in which the physical properties of materials change cally large numbers of J(V) data (N ¼ 500–700; see Supplementary
as a function of the number n of a small repetitive unit such as Table S2 for all statistics) following previously described pro-
2(CH2)2; materials with neven can have very different properties cedures21,22. Figure 2a shows the log-average J(V) curves of junctions
from those with nodd. These so-called odd–even effects determine with SAMs of SC10Fc and SC11Fc, and Fig. 2b presents histograms
the properties of both bulk and nanoscopic materials, for example, of the values of R, as well as Gaussian fits to these histograms
the stability and packing densities of self-assembled monolayers (Supplementary Figs S3–S5 show all data). Figure 2c–f summarizes
(SAMs)15, the melting points of n-alkanes15 and tunnelling rates16. four characteristics of the junctions with SAMs of n ¼ 6–15 as a
To study the role of van der Waals interactions in the perform- function of n. Junctions with SAMs with n ¼ 9, 11 and 13 had (i)
ance of molecular diodes, we used the ‘EGaIn’ technique17. This smaller leakage currents, by a factor of 10 (Fig. 2c), (ii) larger
technique uses template-stripped silver (AgTS) bottom electrodes values of R, by a factor of 10 (Fig. 2d), (iii) 10% higher
supporting SAMs and a soft liquid metal (a eutectic alloy of yields in working devices (Fig. 2e), and (iv) smaller log-standard
gallium and indium with a 0.7-nm-thick surface layer of predomi- deviations of the R values, by a factor of 2–3 (Fig. 2f ), than junctions
nantly Ga2O3) as a non-invasive top electrode18. The EGaIn tech- with n ¼ 8, 10 and 12. Junctions with SAMs of n ¼ 6 and 7 have low
nique is suitable for conducting physical–organic studies of charge yields in working devices (10% and 43%, respectively); these SAMs

1
Department of Chemistry, National University of Singapore, 3 Science Drive 3, Singapore 117543, 2 Department of Physics, National University of Singapore,
2 Science Drive 3, Singapore, 117551, Singapore, 3 Institute of Materials Research and Engineering, 3 Research Link, Singapore 117602, Singapore, 4 Tyndall
National Institute, University College Cork, Lee Maltings, Dyke Parade, Cork, Ireland, 5 Graphene Research Centre, National University of Singapore, 2 Science
Drive 3, Singapore 117542, Singapore, † These authors contributed equally to this work. *e-mail: christian.nijhuis@nus.edu.sg; damien.thompson@tyndall.ie

NATURE NANOTECHNOLOGY | VOL 8 | FEBRUARY 2013 | www.nature.com/naturenanotechnology 113


LETTERS NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2012.238

a
Bulk Ga-In Bulk Ga-In

Ga2O3 Ga2O3

αodd αeven

αodd < αeven


vdW

200 μm
AgTS AgTS

b Hopping
Reverse bias Forward bias
leakage current: current: −3.3 eV
Tunnelling barrier Tunnelling barrier
dtot = dFc + dC dtot = dC
n n
HOMO
−4.3 eV
−4.5 eV −4.5 eV
−5.7 eV −5.3 eV
HOMO

dC dFc dC dFc
n n

S (CH2)n Fe S (CH2)n Fe
Ga2O3− Ga2O3−
Ag EGaIn Ag EGaIn
Ground n = 6−15 6.7 Å Bias Ground n = 6−15 6.7 Å Bias

Figure 1 | Schematic illustration of junctions of the type AgTS–SCnFc//Ga2O3/EGaIn and the mechanism of charge transport across them. a, Left: optical
micrograph image of the tunnelling junction, showing a cone-shaped tip of Ga2O3/EGaIn suspended from a microneedle in contact with a SAM on a AgTS
surface. Middle and right: schematic illustrations of ‘ideal’ tunnelling junctions with SAMs of SCnFc with n ¼ 10 (top right) or 11 (top middle). The difference
in the tilt angle a of the Fc units with respect to the surface normal is also indicated. b, Energy level diagrams of the junctions at a bias of 21.0 V and
þ1.0 V. At negative bias (or forward bias), when the molecular diode allows current to pass through, the HOMO level centred at the Fc units falls between
the energy window of both Fermi levels and can participate in the mechanism of charge transport. The two-step mechanism of charge transport involves
tunnelling of an electron from the HOMO of the Fc across the alkyl chain to the bottom electrode, followed by hopping of an electron to the Fc unit in a
second step. The width of the tunnelling barrier, dtot , is determined by the length of the alkyl chain, dalkyl , or dtot ¼ dalkyl. At positive bias (or reverse bias),
when the diode blocks the current, this HOMO level falls below both Fermi levels and cannot participate in the mechanism of charge transport.
Consequently, the whole length of the molecule forms a barrier against tunnelling of width dtot ¼ dalkyl þ dFc. For ideal diodes, this current would be
infinitesimally small and is called ‘leakage current’ in analogy to conventional diodes.

are disordered, probably because of weak alkyl–alkyl chain inter- molecular dynamics simulations—to the electrical characteristics
actions. Junctions with SAMs of n ¼ 14 and 15 gave good yields of the junctions. Our molecular diodes perform better on silver sur-
in working devices (79% and 71%, respectively), but low values of faces than on gold surfaces, but electrochemical characterization of
R; these SAMs are disordered and suffer from back-bending Fc moi- the SAMs is only possible on gold surfaces (Fig. 3). To be able to
eties (that is, Fc moieties that fold back, pointing towards the bottom relate the data obtained by these two techniques, we characterized
electrode; see below)23. the SAMs on both silver and gold surfaces using NEXAFS and mol-
Figure 2c shows that the odd–even effect of the leakage current ecular dynamics (Fig. 4).
measured at positive bias, or when J depends on dtot,reverse , is Figures 3 and 4 highlight four important observations. (i) The
more pronounced than that of the current measured at negative values of a as a function of n, determined theoretically by molecular
bias, or when J depends on dtot,forward. The measured values of J dynamics (Fig. 3a,b) and experimentally by NEXAFS spectroscopy
depend exponentially on the values of d, as described by a simplified (Fig. 3c,d), are on average 58 smaller (that is, the Fc units are
form of the Simmons equation J ¼ J0e2bd, where b is the tunnelling standing up more) for SAMs with nodd on AgTS substrates than
decay coefficient (Å21) and J0 (A cm22) is the pre-exponential for SAMs with neven , while the reverse is true for SAMs on gold
factor (which depends on several parameters including contact (Supplementary Figs S7–S9). (ii) SAMs on AuTS with nodd are
resistance)24. Thus, only at positive bias can a small variation in more easily oxidized than SAMs with neven , by 27+12 meV (or
the orientation of the Fc units as a function of n cause large vari- 0.6+0.3 kcal mol21; Fig. 4a,b). (iii) The Fc–Fc and Fc–alkyl chain
ations in the measured values of J and, consequently, change the interactions are constant as a function of n (Fig. 4c). (iv) The
performance of the molecular diodes. alkyl–alkyl chain interactions increase with increasing n and are
We have verified this hypothesis by relating the supramolecular more favourable (by 0.4+0.6 kcal mol21) for SAMs on gold
structure of the SAMs—determined by near-edge X-ray absorption with neven than with nodd , while the opposite is true for SAMs on
fine structure (NEXAFS) spectroscopy, electrochemistry and silver (Fig. 4d).

114 NATURE NANOTECHNOLOGY | VOL 8 | FEBRUARY 2013 | www.nature.com/naturenanotechnology


NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2012.238 LETTERS
a d
Low Ordered SAMs Back-bending
SC10Fc
10−2 yields SAMs
SC11Fc
100
10−3
Large leakage

Rectification ratio, R
current:
Current (A cm−2)

10−4 R ≈ 17
10
10−5

10−6
1
Small leakage
10−7 current:
R ≈ 151
10−8 0.1
−1 0 1 6 7 8 9 10 11 12 13 14 15
Voltage (V) Chain length, n

b 60 e 100
SAM: SC10Fc
R = 17
50
σR = 5.11
80

Yield in working devices (%)


SAM: SC11Fc
40 R = 151
σlog = 2.14 60
Counts

30
σR σR
40
20

20
10 σR σR

0 0
10−2 10−1 100 101 102 103 104 105 6 7 8 9 10 11 12 13 14 15
Rectification ratio, R Chain length, n

c 10−1 f 7
|J| at −1.0 V (forward bias)
|J| at +1.0 V (reverse bias)
10−2 6
Log-standard deviation of R, σR

10−3
Current (A cm−2)

10−4 4

10−5 3

Small leakage
10−6 2
current
Large leakage current
10−7 1
6 7 8 9 10 11 12 13 14 15 6 7 8 9 10 11 12 13 14 15
Chain length, n Chain length, n

Figure 2 | Electrical characteristics of the tunnelling junctions of AgTS–SCn Fc//Ga2O3/EGaIn with n 5 6–15. a, The J(V) curves of junctions with SAMs of
SCnFc with n ¼ 11 or 10. Error bars are omitted for clarity (see Supplementary Fig. S3 for graphs with error bars). b, Histograms of R, with Gaussian fits (grey
lines), are for junctions with SAMs of SCnFc with n ¼ 11 or 10 (see Supplementary Fig. S4 for all histograms of R). c, Values of J measured at þ1.0 and
21.0 V, as a function of n (see Supplementary Fig. S5 for all histograms of J). Error bars represent one log standard deviation (black) and 95% confidence
level intervals (blue). d, Values of R as a function of n. Error bars represent one log standard (black) deviation and 95% confidence level intervals (blue).
e, Yields of working junctions as a function of n. f, Log standard deviations of the value of R as a function of n. In d–f, vertical dashed lines show how the
junction properties depend on n, with boundaries marking the transition between low yields from devices made from SAMs with n , 8, excellent device
properties for SAMs with nodd in the range n ¼ 8–13, and finally less ordered, back-bending SAMs for n . 13.

It has been reported that oxidation of the Fc units results in com- During this process, the molecules of the SAMs are forced to stand
plexation of the ClO42 anions to the Fcþ cations, accompanied by up more25, which can only be achieved by rearranging the van der
structural changes of the SAMs25. During oxidation of the SAMs, Waals interactions between the alkyl chains. For SAMs on gold,
the ClO4 2 anions bind to the ferrocenium cations, forming these alkyl chain interactions are weaker for nodd than for neven
strong ion pairs, and thus break the Fc–Fc and Fc–alkyl interactions. and hence require less energy to rearrange during oxidation,

NATURE NANOTECHNOLOGY | VOL 8 | FEBRUARY 2013 | www.nature.com/naturenanotechnology 115


LETTERS NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2012.238

a b 90
Alkyl chain
Ag H
80
Fc unit C Fe
70

Tilt angle α (°)


S
60
Sulphur

50

40

30
6 7 8 9 10 11 12 13 14 15
Chain length, n

c d 60
Cp π*
(3e2u)
SC7Fc 58
Normalized intensity (a.u.)

Cp π*
(4e1g)
56

Tilt angle α (°)


54
SC6Fc
52

θ = 20°
50
θ = 90°

280 290 300 310 6 7 8 9 10 11 12 13 14 15


Photon energy (eV) Chain length, n

Figure 3 | Structural characterization of the SAMs by molecular dynamics and NEXAFS. a, Zoom-in of chain packing in the centre of a SAM with n ¼ 9,
showing the prevalence of upright chain orientations (hydrogen atoms are omitted for clarity). Inset: structure of the molecule. The full SAM model is shown
in Supplementary Fig. S10, as well as a full description of SAM formation and dynamics on Supplementary Information pages S34–S44. b, Calculated Fc tilt
angles a. An odd–even effect of 58 in a is computed from the average of the calculated tilt angles. c, Near C K-edge NEXAFS spectra for SAMs of SC6Fc and
SC7Fc on AgTS recorded at an incident angle of u ¼ 908 and a grazing angle of u ¼ 208. The signals at 285.4 eV and at 286.9 eV are assigned to C 1s  4e1g
and C 1s  3e2u transitions, respectively, based on previous assignments27. d, Value of a (with an estimated error of approximately +58) as a function of n,
derived from NEXAFS spectra (Supplementary Figs S7 and S8).

lowering the oxidation potential. The electrochemical data, sup- more flat-lying Fc units in SAMs with neven prevent optimal
ported by molecular dynamics simulations and NEXAFS measure- packing of the molecules, resulting in slightly thinner SAMs that
ments, indicate that these SAMs on AuTS have lower packing energy, pack less well and are less stiff. These SAMs are more prone to
DEpack ¼ 0.6+0.3 kcal mol21, because the Fc units are standing defects during fabrication, resulting in working devices with lower
up more, resulting in more favourable alkyl–alkyl chain interactions yields (Fig. 2e) that block the current less efficiently at reverse
for neven than for nodd. bias, resulting in large leakage currents (Fig. 2c) and small rectifica-
In general, the odd–even effects for SAMs on gold and silver are tion ratios of R ≈ 10 (Fig. 2d). These defects result in variation in the
reversed due to the different metal–S–C bond angles (Ag–S–C is average values of dtot,reverse from device to device, which, in turn,
close to 1808 and Au–S–C is close to 1098)15. Thus, we expect causes large error in the rectification ratios (Fig. 2f ).
that the SAMs on silver with nodd have better packing than SAMs The calculated SAM structures (Fig. 3a and Supplementary
with neven , by a similar amount of energy. Indeed, molecular Fig. S11) confirm the electrochemical data (Supplementary
dynamics simulations and NEXAFS (Supplementary Figs S7–S9) Fig. S2), which indicate that most of the Fc groups are confined at
show that for SAMs on gold the Fc units are standing up the top of the SAMs for n ¼ 6–13, but for n ¼ 14 or 15 the SAMs
more, with neven allowing better packing (DEpack) of the SAM by are defective due to back-bending of the Fc units23. Hence, control
0.4+0.6 kcal mol21 (or 1.2+1.8 kcal nm22), whereas for SAMs over the supramolecular structure of the SAMs is essential for the per-
on silver this is true for nodd (Fig. 4d and Supplementary formance of the molecular diodes. Only those devices with well-
Fig. S12). These calculations also show that the more stable SAMs defined supramolecular structures—with most of the Fc units
with smaller tilt angles a are slightly thicker and stiffer located at the top of the SAM—perform well and rectify current
(Supplementary Fig. S13); that is, they have smaller root-mean- with R . 100 (ref. 18). In this study, transmission rates26 across the
square fluctuations in the atom positions. tunnelling junctions do not change significantly, probably because
The more upright Fc units in molecular diodes consisting of of the small change in the average value of a (see Supplementary
SAMs on AgTS with nodd pack better and are stiffer because of Information page S2 for further discussion), which implies that maxi-
more favourable molecule–molecule interactions. These SAMs are mizing the rectification ratio by blocking leakage currents (Fig. 2c) is
stable during fabrication, resulting in working devices with high perhaps the most important issue for optimal device performance.
yields (Fig. 2e) that block the current efficiently at reverse bias, In summary, we have shown that control over the supramolecu-
resulting in large rectification ratios of R ≈ 100 (Fig. 2d). The lar structure of SAMs is required to ensure optimal performance of

116 NATURE NANOTECHNOLOGY | VOL 8 | FEBRUARY 2013 | www.nature.com/naturenanotechnology


NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2012.238 LETTERS
a b
0.10
Epa
450
Epc

Epeak (meV) vs Ag/AgCl


0.05
400
Current (mA)

0.00
350

−0.05 SC11Fc:
Epa = 392 ± 10 mV 300
SC12Fc:
−0.10 Epa = 419 ± 7 mV Back-bending
250
0.0 0.2 0.4 0.6 0.8 6 7 8 9 10 11 12 13 14 15
Voltage (V) vs Ag/AgCl Chain length, n

c d 0.9
SAMs on AgTS
−5
SAMs on AuTS
0.6
Interactions (kcal mol−1)

−10

ΔEpack (kcal mol−1)


0.3
−15
0.0
−20
Fc–Fc −0.3
Alkyl–alkyl Fctot
−25
Fc–alkyl Molecule
−0.6
6 7 8 9 10 11 12 13 14 15 7 8 9 10 11 12 13 14
Chain length, n Chain length, n

Figure 4 | Packing energies as a function of n derived from electrochemistry and molecular dynamics. a, Cyclic voltammograms of SAMs of SC11Fc and
SC12Fc on AuTS with aqueous 1.0 M HClO4 as electrolyte solution, recorded at a scan rate of 1.0 V s21 (see Supplementary Fig. S2 for all cyclic voltammograms).
b, Peak oxidation (Epa) and reduction potentials (Epc) for SAMs on AuTS as a function of n, with aqueous 1.0 M HClO4 as electrolyte solution, recorded at a scan
rate of 1.0 V s21. c, Computed SAM packing interactions on Ag for alkyl–alkyl chain, Fc–alkyl chain and Fc–Fc head group interactions as a function of n. Also
plotted are the full Fc packing (Fctot , composed of Fc–Fc packing plus Fc–alkyl packing) and the full molecule packing (Molecule, composed of alkyl–alkyl and
Fctot packing). The results for SAMs on Au are given in Supplementary Fig. S14 and plots with error bars (0.220.4 kcal mol21) are given in Supplementary
Fig. S12. d, Odd–even differences DEpack in calculated SAM packing energies on Ag and Au substrates; the DEpack value at each chain length n is the deviation of
the Epack value at n from the average of the Epack values at n 2 1 and n þ 1. An odd–even effect of 0.4 kcal mol21 in molecule packing energies is estimated from
the average of the second derivatives around each Epack data point (Supplementary Information page S39).

SAM-based devices. A seemingly small change of 58, on average, in SAMs, we used cone-shaped tips of Ga2O3/EGaIn as top electrodes using a home-
the tilt angle a of the active component of the SAMs (the Fc units built system (Supplementary Information page S21), because the electrical
characteristics are dominated by the chemical and supramolecular structure of the
in our case) as a function of the number of 2CH22 units in the SAMs inside the junctions, not because of any other asymmetries, nor the Ga2O3
backbone of the SAMs improves the packing of the molecules by layer, present in these junctions. We recorded statistically large numbers of J(V) data,
0.4–0.6 kcal mol21 because of more favourable intermolecular which we analysed using previously reported methods (Supplementary Information
van der Waals interactions. This more stable and stiffer SAM struc- page S23)22. We calculated the structures of the SAMs with molecular dynamics
ture results in a remarkable improvement in the performance of the simulations (440 ns in total) of statistically large numbers of molecules (1,216)
adsorbed on Ag(111) and Ag(111) electrodes with areas of 33 × 13 nm2. We
devices. In the future, to ensure optimal performances, the rational modelled monolayer assembly29,30 followed by metal–S bond formation for SAMs
design of SAM-based junctions (and studies involving charge trans- with n ¼ 6–15. To minimize edge effects, we measured the properties of the SAMs
port across (bio)molecules in general) should take into account van within 250-molecule central disks with a radius of 5 nm (or 79 nm2). Full details of
der Waals interactions, as demonstrated in this work. the molecular simulations are given on Supplementary Information page S34 onwards.

Methods Received 10 August 2012; accepted 20 November 2012;


28 17
The ferrocene derivatives , the SAMs , Au (ref. 17) and the Ag surfaces were TS TS 17 published online 6 January 2013
prepared following reported procedures (Supplementary Information page S3). We
characterized the SAMs on AuTS with cyclic voltammetry, using a home-built References
electrochemical cell equipped with Pt-disk counter electrode and a Ag/AgCl 1. Mujica, V., Ratner, M. A. & Nitzan, A. Molecular rectification: why is it so rare?
reference electrode, at a scan rate of 1.0 V s21 in aqueous 1.0 M HClO4 as the Chem. Phys. 281, 147–150 (2002).
electrolyte solution (Supplementary Information page S18). The NEXAFS spectra at 2. Lindsay, S. M. & Ratner, M. A. Molecular transport junctions: clearing mists.
the C K-edge were recorded using the Surface, Interface and Nanostructure Science Adv. Mater. 19, 23–31 (2007).
(SINS) beamline at the Singapore Synchrotron Light Source (SSLS) in an ultrahigh- 3. Moth-Poulson, K. & Bjornholm, T. Molecular electronics with single molecules
vacuum (UHV) chamber with a base pressure of 1 × 10210 mbar. We recorded the in solid-state devices. Nature Nanotech. 4, 551–556 (2009).
angular-dependent NEXAFS spectra in the Auger electron yield (AEY) mode by 4. Diez-Perez, I. et al. Controlling single-molecule conductance through lateral
collecting the Auger electrons resulting from carbon KVV transitions with a Scienta coupling of p orbitals. Nature Nanotech. 6, 226–231 (2011).
R4000 electron energy analyser. To estimate the tilt angle a of the Fc units, we used 5. McCreery, R. L. & Bergren, A. Progress with molecular electronic junctions:
the angular dependence of the intensity of the C 1s  4e1g transitions at u ¼ 208 and meeting experimental challenges in design and fabrication. Adv. Mater. 21,
908 (Supplementary Information page S29). To form good electrical contacts to the 4303–4322 (2009).

NATURE NANOTECHNOLOGY | VOL 8 | FEBRUARY 2013 | www.nature.com/naturenanotechnology 117


LETTERS NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2012.238

6. Choi, S. H., Kim, B. & Frisbie, C. D. Electrical resistance of long conjugated 23. Auletta, T., van Veggel, F. C. J. M. & Reinhoudt, D. N. Self-assembled
molecular wires. Science 320, 1482–1486 (2008). monolayers on gold of ferrocene-terminated thiols and hydroxyalkanethiols.
7. Park, S. et al. Flexible molecular-scale electronic devices. Nature Nanotech. 7, Langmuir 18, 1288–1293 (2002).
438–442 (2012). 24. Joachim, C. & Ratner, M. A. Molecular electronics: some views on transport
8. Cornil, J., Beljonne, D., Calbert, J-P. & Brédas, J-L. Interchain interactions in junctions and beyond Proc. Natl Acad. Sci. USA 102, 8801–8808 (2005).
organic p-conjugated materials: impact on electronic structure, optical response, 25. Ye, S., Sato, Y. & Uosaki, K. Redox-induced orientation change of a self-
and charge transport. Adv. Mater. 13, 1053–1067 (2001). assembled monolayer of 11-ferrocenyl-1-undecanethiol on a gold electrode
9. Henson, Z. B., Müllen, K. & Bazan, G. C. Design strategies for organic studied by in situ FT-IRRAS. Langmuir 13, 3157–3161 (1997).
semiconductors beyond the molecular formula. Nature Chem. 4, 26. Venkataraman, L., Klare, J. E., Nuckolls, C., Hybertson, M. A. &
699–704 (2012). Steigerwald, M. L. Dependence of single-molecule junction conductance on
10. Wu, S. M. et al. Molecular junctions based on aromatic coupling. Nature molecular conformation. Nature 442, 904–907 (2006).
Nanotech. 3, 569–574 (2008). 27. Rühl, E. & Hitchcock, A. P. Carbon K-shell excitation of metallocenes J. Am.
11. Hoeben, F. J. M., Jonkheijm, P., Meijer, E. W. & Schenning, A.P. H. J. Chem. Soc. 111, 5069–5075 (1989).
Supramolecular assemblies of p-conjugated systems. Chem. Rev. 105, 28. Creager, S. E. & Rowe, G. K. Competitive self-assembly and electrochemistry of
1491–1546 (2005). some ferrocenyl-n-alkanethiol derivatives on gold. J. Electroanal. Chem. 370,
12. Zaworotko, M. J. & Moulton, B. From molecules to crystal engineering: 203–211 (1994).
supramolecular isomerism and polymorphism in network solids. Chem. Rev. 29. Perl, A. et al. Gradient-driven motion of multivalent ligand molecules along a
101, 1629–1658 (2001). surface functionalized with multiple receptors. Nature Chem. 3, 317–322 (2011).
13. Krishnamurthy, V. M. et al. Carbonic anhydrase as a model for biophysical and 30. Gannon, G., Greer, J. C., Larsson, J. A. & Thompson, D. Molecular dynamics
physical–organic studies of proteins and protein–ligand binding. Chem. Rev. study of naturally occurring defects in self-assembled monolayer formation. ACS
108, 946–1051 (2008). Nano 4, 921–932 (2010).
14. Ho, P. K. H. et al. Molecular-scale interface engineering for polymer light-
emitting diodes. Nature 404, 481–484 (2000).
15. Tao, F. & Bernasek, S. L. Understanding odd–even effects in organic self-
Acknowledgements
The Singapore National Research Foundation (NRF award no. NRF-RF2010-03 to C.A.N.)
assembled monolayers. Chem. Rev. 107, 1408–1453 (2007).
is acknowledged for supporting this research. D.T. acknowledges financial support from the
16. Thuo, M. N. et al. Odd–even effects in tunneling across self-assembled
Science Foundation Ireland (SFI; grant no. 11/SIRG/B2111) and the use of computing
monolayers. J. Am. Chem. Soc. 133, 2962–2975 (2011).
resources at Tyndall and the SFI/Higher Education Authority Irish Centre for High-End
17. Chiechi, R. C., Weiss, E. A. Dickey, M. D. & Whitesides, G. M. Eutectic gallium–
Computing (ICHEC). The authors thank Su Ying Quek for useful discussions and the
indium (EGaIn): a moldable liquid metal for electrical characterization of self-
technical support from the Singapore Synchrotron Light Source.
assembled monolayers. Angew. Chem. Int. Ed. 47, 142–146. (2008).
18. Reus, W. F., Thuo, M. N., Shapiro, N. D., Nijhuis, C. A. & Whitesides, G. M. The
SAM, not the electrodes, dominates charge transport in metal-monolayer// Author contributions
Ga2O3/gallium-indium eutectic junctions. ACS Nano 6, 4806–4822 (2012). N.N. synthesized the compounds and characterized the SAMs. L.Y. performed the J(V)
19. Fracasso, D., Valkenier, H., Hummelen, J. C., Solomon, G. C. & Chiechi, R. C. measurements. J.L. prepared the template-stripped substrates. Q.D.C. and L.Y. recorded
Evidence for quantum interference in SAMs of arylethynylene thiolates in and analysed the NEXAFS spectra. D.T. performed the molecular dynamics simulations.
tunneling junctions with eutectic Ga-In (EGaIn) top-contacts J. Am. Chem. Soc. C.A.N. supervised the project. All authors contributed to writing the manuscript.
133, 9556–9563 (2011).
20. Masillamani, A. M. et al. Multiscale charge injection and transport properties in
self-assembled monolayers of biphenyl thiols with varying torsion angles Chem.
Additional information
Supplementary information is available in the online version of the paper. Reprints and
Eur. J. 18, 10335–10347 (2012).
permission information is available online at http://www.nature.com/reprints. Correspondence
21. Nijhuis, C. A., Reus, W. F., Barber, J., Dickey, M. D. & Whitesides, G. M. Charge
and requests for materials should be addressed to D.T. and C.A.N.
transport and rectification in arrays of SAM-based tunneling junctions. Nano
Lett. 10, 3611–3619 (2010).
22. Reus, W. F. et al. Statistical tools for analyzing measurements of charge transport. Competing financial interests
J. Phys. Chem. C 116, 6714–6733 (2012). The authors declare no competing financial interests.

118 NATURE NANOTECHNOLOGY | VOL 8 | FEBRUARY 2013 | www.nature.com/naturenanotechnology

You might also like