• Tidak ada hasil yang ditemukan

Tuning the Tunneling Rate and Dielectric

N/A
N/A
Protected

Academic year: 2018

Membagikan "Tuning the Tunneling Rate and Dielectric"

Copied!
7
0
0

Teks penuh

(1)

COMMUNICA

TION

Tuning the Tunneling Rate and Dielectric Response of

SAM-Based Junctions via a Single Polarizable Atom

Dandan Wang, Davide Fracasso, Argo Nurbawono, Harshini V. Annadata,

C. S. Suchand Sangeeth, Li Yuan, and Christian A. Nijhuis*

D. Wang, Dr. D. Fracasso, H. V. Annadata, Dr. C. S. S. Sangeeth, L. Yuan, Prof. C. A. Nijhuis Department of Chemistry

National University of Singapore

3 Science Drive 3, Singapore 117543, Singapore E-mail: chmnca@nus.edu.sg

D. Wang, Dr. D. Fracasso, H. V. Annadata, Dr. C. S. S. Sangeeth, L. Yuan, Prof. C. A. Nijhuis

Centre for Advanced 2D Materials and Graphene Research Centre National University of Singapore

6 Science Drive 2, Singapore 117546, Singapore Dr. A. Nurbawono

Department of Physics National University of Singapore

2 Science Drive 3, Singapore 117551, Singapore

DOI: 10.1002/adma.201502968

0

J=J e−βd (1)

Cahen and co-workers[7,8,25] were able to tune the effective

barrier height by changing the dipole moment of the molecules

(µ in D) using semiconducting bottom electrodes. In contrast,

Whitesides and co-workers showed that in EGaIn-based

junc-tions (see Figure 1) the tunneling rates are insensitive to a

large number of terminal groups including simple aromatic,[9]

polar,[10] and acidic groups.[11] McCreery and co-workers[6] also

found that the tunneling rates are remarkably insensitive to a broad range of chemical groups. On the other hand, subtle changes in the monolayer structure did cause a significant change in the electrical properties of junctions. For instance,

odd–even effects in the tunneling rates as a function of n

have been observed in junctions with SAMs of SCnX (where

X = CH3,[12] phenyl,[13] or ferrocene[14] or replacing a C–H unit

by a N in SAMs of oligo(phenylene)s result in clear change of

the transition voltage.[15] The tunneling rates in EGaIn

junc-tions are sensitive to some terminal groups including

ferro-cene,[16] bipyridyl,[17] and naphthoquinone[18] (all these groups

induced rectification). In addition, examples of junctions with conjugated systems showed quantum interference effects,

which could be theoretically explained,[19,20] or tunneling rates

that could be related to experimentally determined tunneling

barrier heights and widths.[21] These examples illustrate the

dif-ficulty in predicting how the SAM structure will influence the electrical properties of SAM-based junctions.

A major challenge in the organic semiconductor industry is

the development of high-capacitance gate dielectrics.[26] In order

to obtain the required drive currents at relatively low operating voltages for field-effect devices, thin gate dielectrics with high capacitances and low leakage currents are necessary. SAMs have been proposed as attractive candidates since they are thin,

but their εr values (<4) are in general too low for most

applica-tions. The capacitance of a SAM (CSAM in µF cm−2) sandwiched

between two electrodes—a SAM-based tunneling junction—is

proportional to εr and inversely proportional to dSAM (in nm,

the thickness of the SAM) as given by Equation 2 where ε0 is

the permittivity of free space and Ageo (in cm2) is the

geomet-rical area of the electrodes.[27] We note that due to the small tilt

angle of the SAMs on Ag of 11°,[28] the value of d

The leakage current that flows across such junctions is given

by Equation 1 and hence a trade-off exists as increasing CSAM

by lowering dSAM results in large leakage currents.[29] For this

One of the original promises of molecular electronics was that it would make it possible to tune the electronic func-tion of devices at the molecular scale, with a seemingly end-less number of possibilities given the great synthetic freedom

with which the properties of the molecules can be tuned.[1–5]

To achieve this goal, one needs to understand which kinds of chemical groups do affect the electrical characteristics of the junctions and how they do that. Recent studies showed that in practical test beds of the form electrode–SAM–electrode (where SAM refers to a self-assembled monolayer), it is not always straightforward to predict which kind of chemical

functionali-ties will affect the electrical response of the junctions.[6–21] In

addition, systematic experimental investigations concerning the dielectric response of molecular junctions are rare despite their promising potential in applications requiring high-capacitance

gate dielectrics.[22–24] Here, we show that in tunneling

junc-tions with SAMs of S(CH2)11X (SC11X with X = H, F, Cl, Br,

or I), the tunneling rate increases by three orders of

magni-tude and the relative dielectric constant εr by a factor of 4 with

increasing polarizability of X, which also results in a decrease of the highest occupied molecular orbital–lowest unoccupied molecular orbital (HOMO)–(LUMO) gap (and the tunneling barrier height). Hence, introducing polarizable groups seems to be an attractive method to control both the tunneling rate and the dielectric response of two-terminal molecular junctions.

By far most studies have focused on structure–property

rela-tions by measuring tunneling currents (J in A cm−2) across

SAM-based junctions and interpret the data in terms of a

gen-eral tunnel equation (Equation 1) where β (in Å−1) is the

tun-neling decay coefficient, d (in nm) is the tunneling barrier

width (the length of the molecule), and J0 is the current that

(2)

COMMUNICA

TION

reason, it is important to increase εr. Recently, Ratner and

co-workers[24] calculated that the SAM properties such as the

surface coverage, molecular tilt angle, or conjugation, have

a large effect on the value of εr. They showed that densely

packed SAMs, with small tilt angles, and with highly

polariz-able groups, may have large values of εr. For example, they

cal-culated that εr increases from 5.1 to 8.7 for monosubstituted

trans-polyacetylene SAMs when the terminal hydrogen was

replaced by groups with increasing polarizability (the highest εr

was found for SAMs with iodine atoms). How the polarizability

of a single atom changes the values of J and εr in a SAM-based

device has not been experimentally determined yet.

In this work, we aim to address the following question: how

does a change in the polarizability (α in Å3) of a single atom in

a SAM-based junction changes the value of εr and the

meas-ured value J and how significant would this change be? We

used eutectic gallium–indium (GaOx/EGaIn, where GaOx is

0.7 nm thick native oxide of predominantly β-Ga2O3) as a

non-invasive soft top-electrode[30,31] in junctions with the SAMs

sup-ported by template-stripped (TS) Ag that were annealed before

TS (AgA-TS).[32] These bottom electrodes have large grains and

are ultra-smooth (Figure S3, Supporting Information). Figure 1

shows these junctions of the form AgA-TS–SC

11X//GaOx/EGaIn

schematically and the relative sizes of the substituents X = H,

F, Cl, Br, or I. In these junctions, we changed α of a single atom

while keeping the changes to the permanent dipole, the surface coverage, tilt angle of the SAM, and the SAM thickness, to a

minimum. Below we describe how this change in α changes

the tunneling rates and the dielectric constant, but first we dis-cuss the SAM structure in more detail.

The SAMs were derived from the corresponding thiols (the synthetic details are described on page S2–S9 in the Supporting Information) and characterized by angle-resolved X-ray photo-electron spectroscopy (AR-XPS), ultraviolet photophoto-electron spec-troscopy (UPS), and contact angle (CA) measurements. The

results are given in Table1 and Figure2 (see page S15–S22 in

the Supporting Information for more details). Figure 2a shows

the C 1s spectra recorded at incident angle (90°). The spectra

are dominated by the two signals associated with the aliphatic carbon and the carbon next to the halogen, which shifts to lower

binding energy with decreasing electronegativity χ of X. From

the AR-XPS data, we determined dSAM and the surface

cov-erage (ΓSAM in mol cm−2). All SAMs are ≈1.5 nm thick (except

for X = H which is 1.3 nm thick), which is consistent with the

molecular lengths estimated from CPK models. The values of

ΓSAM of 1.1 × 10−9 mol cm−2 are close to theoretical values for

SAMs on Ag with a ( 7× 7) 10.9R ° packing.[33] From these

data, we conclude that the SC11X thiolates form densely packed

SAMs, which are standing up with small tilt angles (likely close

to ≈11° as has been reported before for n-alkanethiolate SAMs

on Ag).[33]

Figure 2b shows the work function (Φ in eV) and the HOMO

(in eV) plotted against χ, the former decreases by 0.40 eV with

increasing atom number Z of the halogen. We ascribe this

decrease in Φ to the decrease in dipole moment as χ decreases

with increasing Z.[34–37] The energy of the highest occupied

molecular orbital (EHOMO in eV, which is also called the

ioniza-tion potential), calculated as the sum of the HOMO onset and

Φ, increases as χ increases (Table 1) while the HOMO-onset

values remain constant. This observation can be explained

as the value Φ can change without changing the energy level

alignment (i.e., the value of EHOMO) because the electrostatic

change as a result of the change in dipole by varying χ is

con-fined within in SAM (i.e., here the electronic structure of the

www.advmat.de

www.MaterialsViews.com

Figure 1. Schematic illustration of the AgA-TS–SC

11X//GaOx/EGaIn

junc-tion with X = H, F, Cl, Br, or I. The equivalent circuit, van der Waals (vdW) radius of X, and how α depends on X are also indicated. RC is the total contact resistance, RSAM is the SAM resistance, CSAM is the capacitance of the SAM (determined from the constant phase element, CPE).

Table 1. The properties of the SAMs.

SAMs dSAMa)

a)Determined by AR-XPS; the error represents the error of the fit, for d

SAM the error was 0.01 nm (see page S16–S18 in the Supporting Information); b)Determined by

UPS; the error in the energy resolution was 0.05 eV; c)Determined from TD-DFT;[60] see page S28–S29 in the Supporting Information; d)Determined using the experimental

HOMO values and theoretical Eg values; e)The polarizability of the respective molecule in the gas phase calculated by DFT (see page S24–S25 in the Supporting

Informa-tion); f)Dispersive γ

sD and polar γsP components of the surface energy estimated from CA measurements (see page S23 in the Supporting Information for the errors); g)We

(3)

COMMUNICA

TION

bottom electrode–SAM interface is not affected by the terminal

group, only the work function).[36–38]

We determined the dispersive (γsD in mN m−1) and polar (γsP

in mN m−1) components of the surface energy of the SAMs by

CA measurements with two liquids (H2O and CH2I2) using

Fowkes’ method.[39] Table 1 shows that the values of γ

sD are

about a factor ten larger than the values of γsP. We determined

the value of α by density functional theory (DFT) calculations

(using B3LYP functionals and the 6-311G** basis set) of gas phase molecules and an empirical method based on an addi-tive approach of the atomic polarizabilities (see page S24–S25 in the Supporting Information for details). The DFT calculations

resulted in values of α that were consistently ≈10% lower than

those obtained by the additive method (Table S2, Supporting Information) from which we conclude that both approaches give reasonable values. Here we used the values obtained by the DFT calculations. To understand the dielectric distribution at the molecular scale, we performed DFT calculations similar to those reported by Ratner and co-workers (see page S26 in the

Supporting Information).[24] Figure 2c shows the planar

aver-aged local dielectric constant (η) profiles for C11X monolayers,

which also include a simple hydrocarbon chain (CnH2n+2 with

n = 11) to underline the effect of X moieties. Similar to the

results described by Ratner and co-workers, the hydrocarbon

alkyl chain has a symmetric distribution of η along the length of

the molecule concentrating towards the center. The halogenated hydrocarbons differ from each other only at the halogen position

and a clear trend is visible for η, which decreases in the order

I > Br > Cl > F. From these data, and the CA measurements, we

conclude that we changed primarily the value of polarizability α

of the SAMs by increasing the size of the halogen substituent.

We formed electrical top-contacts to the SAMs using

cone-shaped tips of GaOx/EGaIn, recorded 400–800 J(V) curves

(Table S4, Supporting Information) for each type of junc-tion, and analyzed these data, using previously reported

methods (see page S30 in the Supporting Information).[14,32,40]

Figure 3a shows the log-average values of J (<log|J|>) vs the

applied bias V (the error bars are the log-standard deviations).

This Figure shows that the J(V) curves for junctions with X = H

or F are indistinguishable. In contrast, when X changes from F

(or H) to I, the values of <log|J|> increase by a factor of 3 (i.e., J

increases by three orders of magnitude). Below we discuss this observation in more detail.

To examine the mechanism of charge transport across the

junctions in more detail, we performed the J(V) measurements

as a function of temperature over the range of temperatures of 240 to 330 K using junctions where the EGaIn is mechanically stabilized in a through-hole in poly(dimethylsiloxane) (PDMS); the fabrication and characterization of these junctions are

described elsewhere.[40] Figure 3b shows that the J(V) curves

are independent of the temperature. From this observation, we conclude that in all cases the mechanism of charge transport across the junctions is coherent tunneling.

Figure 3c,d shows plots of <log|J|> determined at +0.50 V

(<log|J|>+0.50 V) as a function of α, γsD and γsP. We found that

the values of <log|J|>+0.50 V follow a surprisingly good linear

correlation with α and γsD, but no obvious correlation is

vis-ible with γsP or the permanent dipole moment of the molecules

normal to surface µ(estimated from DFT; see page S25 and

Figure S17, Supporting Information). These results suggest that

the increase in the tunneling rates is primarily related to γsD

and α (which are related to each other).[39,41,42] To confirm this

(4)

COMMUNICA

TION

hypothesis, we also determined the values of <log|J|> for

junc-tions with other terminal groups (X = OH, COOH, NO2, CN,

or NH2; Figure S17, Supporting Information). In agreement

with findings by Whitesides and co-workers[9–11] these groups

changed the tunneling rates only within 1 order of magnitude though they span a large range of dipole moments (Table S2, Supporting Information). We note that these dipole moments were estimated from gas phase single molecules and not from

the SAMs. Zojer et al.[38] showed theoretically that the dipole

moment of SAMs depend on various factors, e.g., depolariza-tion effects (which are expected to be small in our study as we used aliphatic molecules), orientation of the molecules, or sur-face coverages, but here we solely wish to point out that for the

series of junctions with X = OH, COOH, NO2, CN, and NH2, J

did not change much as opposed to junctions with halogens. To investigate the mechanism of charge transport in more detail, we characterized the junctions by impedance spectros-copy (IS) because IS measures the dielectric properties of a medium as a function of frequency and makes it possible to

measure the contact resistance (RC in mΩ cm2), resistance of

the SAM (RSAM in Ω cm2), and C

SAM (in µF cm−2), independent

from each other.[43] The junctions were probed in the frequency

range of 100 Hz to 1 MHz using a sinusoidal signal with an amplitude of 30 mV following previously reported procedures (Figure S18, Supporting Information shows the Nyquist and

Bode plots).[44] The linearity of the system was confirmed by

Kramers–Kronig (KK) tests (Figure S19 and S20, Supporting Information). The data were fitted to the equivalent circuit

shown in Figure 1. The equivalent circuit consists of a constant

phase element (CPE) in parallel with RSAM both in series with

RC. We used a CPE instead of a pure capacitor to model CSAM

because the former represents an imperfect capacitor caused by defects in the electrode materials (e.g., grain boundaries,

step-edges, dislocations, etc.).[45,46] Although the fitting results are

very similar, the fitting errors (χ2fit) are reduced by a factor of

2–3 and are similar to the error of the KK-test χ2KK (see Table2

and Table S5, Supporting Information).

As we have shown before, in our junctions the RC is

domi-nated by the SAM//GaOx interface (the 0.7 nm thick GaOx layer

is about two orders of magnitude less resistive than the contact

resistance).[43] Therefore, we approximated the two interface

resistances as a single resistor RC.

www.advmat.de

www.MaterialsViews.com

Figure 3. a) Plots of <log|J|> against V for junctions with X = H, F, Cl, Br, and I. b) The J(V) measurements as a function of temperature ranging from 240 to 330 K for junctions with X = H, F, Cl, Br, and I. c) The values of <log|J|>+0.50 V as a function of α. d) The values of <log|J|>+0.50 V as a function of

γsD (black squares) and γsP (blue triangles). The dashed lines are guides to the eye and the error bars represent one log-standard deviation.

Table 2. Summary of the results of the IS measurements.

SAMsa) R C

[mΩ cm2]

RSAM

[Ω cm2]

CSAM

[µF cm−2]

εrb) dvdWc)

[Å] SC11H 6.8 ± 0.6 (3.4 ± 0.6) × 102 1.54 ± 0.07 2.93 ± 0.09 2.72

SC11F 7.0 ± 0.1 (4.9 ± 0.6) × 102 1.16 ± 0.02 2.00 ± 0.03 2.99

SC11Cl 4.7 ± 0.8 (4.6 ± 1.0) × 101 1.67 ± 0.02 2.96 ± 0.03 3.27

SC11Br 3.7 ± 0.2 (1.5 ± 0.1) × 100 2.6 ± 0.1 4.6 ± 0.2 3.37

SC11I 2.1 ± 0.3 (1.6 ± 0.3) × 10−1 4.4 ± 0.4 7.9 ± 0.7 3.50 a)The error bars represent the standard deviation of three independent

meas-urements; b)ε

r was determined from the CSAM values and using the Equation 2; c)d

(5)

COMMUNICA

TION

We note that, in principle, any junction—regardless whether it is classical or non-classical—can be fitted with an equivalent

circuit (as is common practice to model artificial atoms[47] for

instance). The equivalent circuit given in Figure 1 represents a SAM-based tunneling junction and we explain the physical meaning of the elements using a single tunneling barrier model, which includes contact resistance, in the framework

of the Landauer model given by Equation 3.[48] This analysis

makes it possible to define the contact and SAM resistance clearly (see for details ref. [44]). Here, the contact resistance is associated with the coupling of the molecules to the electrodes which is, in an ideal case, the inverse of the quantum

conduct-ance G0= 2e2/h (for M= 1) where h= Planck’s constant, e =

the charge of an electron, T= the transmission probability, and

M = the number of conduction channels.[48] The value of T

depends on both the molecule–electrode binding energy and the tunneling barrier height.

2 2

1

junction 2 2 C SAM

R h

e M h e M

T

T R R

= + − = + (3)

Figure 4a shows that the value of RC (7.0–2.1 mΩ cm2)

decreases with increasing α. One would expect the value

of resistance of the SAM top contact (RC,t) to increase with

increasing dvdW (estimated from the sums of the

respec-tive vdW radii, Table 2),[49] however, Figure 4b shows that R

C

decreases with increasing dvdW. It is well known that the Debye

(dipole-induced dipole) and London (induced dipole-induced

dipole) forces increase with increasing α (page S39 in the

Supporting Information). The changes of µ of the molecules

are small and do not follow an obvious correlation with RC,t

(Figure S21, Supporting Information). CA measurements

indi-cate that value of γsP (γsP= 32.4 mN m−1 and is similar in value

to γsD= 42.5 mN m−1) of the top electrode is large and

there-fore the permanent dipole associated with the GaOx layer may be important (Table S1, Supporting Information). Therefore we believe that with the increase of the polarizability of X, the vdW force between the top-electrode and SAM increases, which lowers the contact resistance.

Figure 4a shows that RSAM decreases from 490 to 0.16 Ω cm2

when X changes changed from F to I. This result shows that

the three orders of magnitude increase of J, observed in the

DC measurements (Figure 3), is mainly caused by a decrease

of the RSAM (and not by the change in RC because RC<< RSAM).

The substantial change in RSAM cannot be attributed to a

change in dSAM or ΓSAM as these parameters did not vary

sig-nificantly as a function of X. However, DFT (to obtain the projected density of state PDOS; page S27 in the Supporting Information) and TD-DFT (to obtain an approximation of the

HOMO–LUMO gap Eg; page S28–S29 in the Supporting

Infor-mation) calculations qualitatively justify this trend in RSAM as

Eg for the single molecule in gas phase decreases consistently

from C11F (Eg= 6.31 eV) to C11I (Eg= 4.70 eV). As described

earlier, the UPS data indicated that the energy level align-ment did not change with changing X (Table 1) and UV–vis measurements (page S29–S30 in the Supporting Information)

allowed us to estimate the optical Eg experimentally (we note

that UV–vis likely underestimates Eg as is explained in detail

in for example ref. [50]).[50] The DFT, TD-DFT, and UV–vis data

qualitatively agree and Table 1 shows that the LUMO levels

decrease with increasing Z. Since the LUMO levels are closer

to the Fermi levels than the HOMO levels, the mechanism of

(6)

COMMUNICA

TION

charge transport involves electron tunneling[51,52] and hence the

increase in tunneling rates is primarily caused by a decrease in the LUMO levels. We note that the shifts in the energy levels may be considerably larger once the SAMs are contacted by

the top-electrode than those estimated by UPS.[53] As explained

above, a decrease in the values of RC implies an increase in the

molecule–electrode interaction, which would also result in an

increase of T. Thus our data indicate that the large increase in

the values of J is the result of a combination of an increase in

the molecule–top electrode coupling and lowering of the tun-neling barrier height as the molecule–top electrode interaction increases with increasing polarizability of the terminal atom of the SAM.

Figure 4c shows the values of CSAM and εr plotted as a

func-tion of α. Here εr was determined using Equation 2 from the

experimentally obtained CSAM. We found that for junctions with

X = I, CSAM= 4.4 ± 0.4 µF cm−2 and εr= 7.9 ± 0.7, which are

both about four times higher than for junctions with X = F.[43]

The solid line in Figure 4c represents a fit to the Clausius–

Mosotti relation, which correlates εr with α, but ignores bulk

effects.[54,55]

As expected, the Clausius–Mosotti relation does not fit our

data well, but the εDFT values calculated by the van der Waals

DFT-D2 method are in close agreement with our experimental results, as shown in Figure 4d (the off-set between theory and experiment may be attributed to the simplicity of the model as the effects of the electrode and the thiol group, or

depolariza-tion effects[22,38,56,57] were ignored). The details of the model

and calculation are described on page S26 in the Supporting Information. Our calculations qualitatively agree well with the

calculations reported by Ratner and co-workers.[24] They showed

theoretically an increase by a factor of 1.7 in εr from 5.1 to 8.7

for monolayers of trans-polyacetylene by changing the halogen

termini of these SAMs from F to I (who also ignored the elec-trodes and the thiol anchoring group). These authors also

showed that εr of conjugated SAMs is less sensitive to the

pres-ence of highly polarizable atoms than non-conjugated SAMs,

which explains why we observed a factor of 4 increase of εr. This

notion also explains why Yoon et al. only observed marginal effects in the tunneling rates in EGaIn junctions of the form

AgTS–S(CH

2)n(p-C6H5X)//GaOx/EGaIn with n = 0 or 1 and

X = H, F, Cl, Br, and I (these authors did not measure εr).[58] In

addition, these short and conjugated SAMs hybridize strongly

with the bottom-electrode masking molecular effects.[59]

In conclusion, our experiments show that in tunneling

junc-tions of the form AgA-TS–SC

11X//GaOx/EGaIn the tunneling

rate increases by three orders of magnitude and the dielec-tric constant increases from 2 to 8 when the polarizability of

X increases (where X = H, F, Br, Cl, or I). These results show

that it is possible to tune the dielectric constant by controlling

the polarizability via high Z substituents, which in turn also

make it possible to control the HOMO–LUMO gaps and con-sequently the tunneling rates across the SAM-based junctions.

The halogenated SAMs have high capacitances (CSAM =

4.4 ± 0.4 µF cm−2 when X = I), but the leakage currents (i.e.,

the tunneling current across the junctions) increased with increasing capacitance. Although in this study we kept the SAM thickness the same, our results imply that it is possible to engineer the dielectric response of SAMs at the atomic level,

which are needed in optimizing the balance between the dielec-tric response and the leakage current.

Supporting Information

Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements

D.W. and D.F. contributed equally to this work. The authors acknowledge the National Research Foundation (NRF) for supporting this research under the Competitive Research Programme (CRP) Award No. NRF-CRP8-2011-07. This research was supported by the National Research Foundation, Prime Minister’s Office, Singapore under its Medium sized centre programme. C.A.N. acknowledges the Ministry of Education for financial support (MOE R-144-000-325-112). The authors thank Dr Xiao-Jiang Yu for help at the Singapore Synchrotron Light Source (SSLS). R. M. Metzger, M. E. Michel-Beyerle, J. R. Miller, M. D. Newton, D. R. Rolison, O. Sankey, K. S. Schanze, J. Yardley, X. Zhu, J. Phys. M. Baghbanzadeh, F. C. Simeone, G. M. Whitesides, Nano Lett. 2014, 14, 3521.

[12] M. M. Thuo, W. F. Reus, C. A. Nijhuis, J. R. Barber, C. Kim, M. D. Schulz, G. M. Whitesides, J. Am. Chem. Soc.2011, 133, 2962. [13] T. Toledano, H. Sazan, S. Mukhopadhyay, H. Alon, K. Lerman,

T. Bendikov, D. T. Major, C. N. Sukenik, A. Vilan, D. Cahen, Lang-muir2014, 30, 13596.

[14] N. Nerngchamnong, L. Yuan, D.-C. Qi, J. Li, D. Thompson, C. A. Nijhuis, Nat. Nanotechnol.2013, 8, 113.

[15] D. Fracasso, M. I. Muglali, M. Rohwerder, A. Terfort, R. C. Chiechi,

J. Phys. Chem. C2013, 117, 11367.

[19] D. Fracasso, H. Valkenier, J. C. Hummelen, G. C. Solomon, R. C. Chiechi, J. Am. Chem. Soc.2011, 133, 9556.

www.advmat.de

(7)

COMMUNICA

TION

[20] C. M. Guedon, H. Valkenier, T. Markussen, K. S. Thygesen, J. C. Hummelen, S. J. van der Molen, Nat. Nano2012, 7, 305. [21] B. Kim, S. H. Choi, X. Y. Zhu, C. D. Frisbie, J. Am. Chem. Soc.2011,

133, 19864.

[22] L. Romaner, G. Heimel, C. Ambrosch-Draxl, E. Zojer, Adv. Funct. Mater.2008, 18, 3999.

[23] V. Obersteiner, D. A. Egger, G. Heimel, E. Zojer, J. Phys. Chem. C 2014, 118, 22395.

[24] H. M. Heitzer, T. J. Marks, M. A. Ratner, ACS Nano2014, 8, 12587. [25] T. Toledano, A. Biller, T. Bendikov, H. Cohen, A. Vilan, D. Cahen, J.

Phys. Chem. C2012, 116, 11434. [26] C. Groves, Nat. Mater.2013, 12, 597.

[27] R. P. Ortiz, A. Facchetti, T. J. Marks, Chem. Rev.2010, 110, 205. [28] J. C. Love, L. A. Estroff, J. K. Kriebel, R. G. Nuzzo, G. M. Whitesides,

Chem. Rev.2005, 105, 1103.

[29] M. Halik, H. Klauk, U. Zschieschang, G. Schmid, C. Dehm, M. Schutz, S. Maisch, F. Effenberger, M. Brunnbauer, F. Stellacci,

Nature2004, 431, 963.

[30] R. C. Chiechi, E. A. Weiss, M. D. Dickey, G. M. Whitesides, Angew. Chem., Int. Ed.2008, 47, 142.

[31] W. F. Reus, C. A. Nijhuis, J. R. Barber, M. M. Thuo, S. Tricard, G. M. Whitesides, J. Phys. Chem. C2012, 116, 6714.

[32] L. Yuan, L. Jiang, B. Zhang, C. A. Nijhuis, Angew. Chem., Int. Ed. 2014, 53, 3377.

[33] J. B. Schlenoff, M. Li, H. Ly, J. Am. Chem. Soc.1995, 117, 12528. [34] B. de Boer, A. Hadipour, M. M. Mandoc, T. van Woudenbergh,

P. W. M. Blom, Adv. Mater.2005, 17, 621.

[35] A.-E. Haj-Yahia, O. Yaffe, T. Bendikov, H. Cohen, Y. Feldman, A. Vilan, D. Cahen, Adv. Mater.2013, 25, 702.

[36] A. Natan, L. Kronik, H. Haick, R. T. Tung, Adv. Mater. 2007, 19, 4103.

[37] G. Heimel, L. Romaner, J.-L. Brédas, E. Zojer, Phys. Rev. Lett.2006,

96, 196806.

[38] L. Wang, G. M. Rangger, L. Romaner, G. Heimel, T. Bucˇko, Z. Ma, Q. Li, Z. Shuai, E. Zojer, Adv. Funct. Mater.2009, 19, 3766.

[39] F. M. Fowkes, Ind. Eng. Chem. Res.1964, 56, 40.

[40] A. Wan, L. Jiang, C. S. S. Sangeeth, C. A. Nijhuis, Adv. Funct. Mater. 2014, 24, 4442.

[41] H. Margenau, Rev. Mod. Phys.1939, 11, 1.

[42] H. Y. Erbil, Surface Chemistry: Of Solid and Liquid Interfaces, Black-well Publishing Ltd., Oxford, UK 2006.

[43] C. S. S. Sangeeth, A. Wan, C. A. Nijhuis, J. Am. Chem. Soc.2014,

136, 11134.

[44] C. S. Suchand Sangeeth, A. Wan, C. A. Nijhuis, Nanoscale2015, 7, 12061.

[45] J.-B. Jorcin, M. E. Orazem, N. Pébère, B. Tribollet, Electrochim. Acta 2006, 51, 1473.

[46] M. E. Orazem, B. Tribollet, Electrochemical Impedance Spectroscopy, John Wiley & Sons, Inc., Hoboken, NJ, USA 2008.

[47] W. G. van der Wiel, S. De Franceschi, J. M. Elzerman, T. Fujisawa, S. Tarucha, L. P. Kouwenhoven, Rev. Mod. Phys.2002, 75, 1. [48] S. Datta, Electronic Transport in Mesoscopic Systems, Cambridge

University Press, Cambridge, UK 1995.

[49] K. E. Riley, K. M. Merz, J. Phys. Chem. A2007, 111, 1688. [50] J.-L. Bredas, Mater. Horiz.2014, 1, 17.

[51] K. Smaali, N. Clément, G. Patriarche, D. Vuillaume, ACS Nano 2012, 6, 4639.

[52] Y. Qi, O. Yaffe, E. Tirosh, A. Vilan, D. Cahen, A. Kahn, Chem. Phys. Lett.2011, 511, 344.

[53] K. Moth-Poulsen, T. Bjornholm, Nat. Nanotechnol.2009, 4, 551. [54] A. Natan, N. Kuritz, L. Kronik, Adv. Funct. Mater.2010, 20, 2077. [55] A. R. Blythe, D. Bloor, Electrical Properties of Polymers, 2nd ed.,

Cam-bridge University Press, CamCam-bridge, UK 2005.

[56] D. Cornil, Y. Olivier, V. Geskin, J. Cornil, Adv. Funct. Mater.2007, 17, 1143.

[57] M. L. Sushko, A. L. Shluger, Adv. Funct. Mater.2008, 18, 2228. [58] G. D. Kong, M. Kim, H.-J. Jang, K.-C. Liao, H. J. Yoon, Phys. Chem.

Chem. Phys.2015, 17, 13804.

[59] L. Yuan, N. Nerngchamnong, L. Cao, H. Hamoudi, E. del Barco, M. Roemer, R. K. Sriramula, D. Thompson, C. A. Nijhuis, Nat. Commun.2015, 6, 6324, DOI:10.1038/ncomms7324.

Gambar

Figure 1. Schematic illustration of the Agradius of X, and how tion with X A-TS–SC11X//GaOx/EGaIn junc-= H, F, Cl, Br, or I
Figure 2. a) The C1s spectra recorded on SAMs on AgA-TS surfaces. b) The work function Φ (black squares) and HOMO (red diamonds) for SC11X SAMs as a function of Pauling electronegativity χ of X
Table 2. Summary of the results of the IS measurements.
Figure 4. a) Plot of RC (black dots) and RSAM (red triangles) as a function of α. b) RC as a function of the dvdW

Referensi

Dokumen terkait

Panitia Pengadaan Barang/Jasa Dinas Kelautan dan Perikanan Kabupaten Aceh Tamiang akan melaksanakan Pelelangan Sederhana dengan prakualifikasi untuk paket pekerjaan Jasa

Management of Provincial Office of Drug and Food Control. Proof has been furnished that the requirements according

Surat Pernyataan melakukan kegiatan pembelajaran/pembimbingan dan tugas tambahan dan/atau tugas lain yang relevan dengan fungsi sekolah/madrasah.. Surat Pernyataan melakukan

Hasil analisis ragam menunjukkan bahwa suplementasi rumput mutiara kedalam ransum sampai taraf 0,8% tidak memberikan pengaruh yang nyata (P&gt;0,05) terhadap bobot

The mediating influence of market orientation on the relationship between organizational culture and firm performance was tested through the four steps of path analysis proposed by

Laporan kerja praktek yang berjudul “ IMPLEMENTASI SISTEM INFORMASI MANAJEMEN DAERAH BARANG MILIK DAERAH (SIMDA- BMD) PADA DINAS PENGELOLAAN KEUANGAN DAN ASET DAERAH

[r]

Abdul dan Nasuhi (2000), dalam penelitian yang menguji pengaruh volume perdagangan dan return terhadap bid-ask spread saham dari 3 perusahaan rokok yang aktif dan liquid