You are on page 1of 7

Applied Surface Science Advances 5 (2021) 100094

Contents lists available at ScienceDirect

Applied Surface Science Advances


journal homepage: www.elsevier.com/locate/apsadv

Self-assembled monolayers of oligophenylenes stiffer than steel and silicon,


possibly even stiffer than Si3 N4
Ioan Bâldea
Theoretical Chemistry, Heidelberg University, Im Neuenheimer Feld 229, Heidelberg D-69120, Germany

a r t i c l e i n f o a b s t r a c t

Keywords: To quantify charge transport through molecular junctions fabricated using the conducting probe atomic force
Interface phenomena microscopy (CP-AFM) platform, information on the number of molecules 𝑁 per junction is absolutely necessary.
Self-assembled monlayers 𝑁 can be currently obtained only via contact mechanics, and the Young’s modulus 𝐸 of the self-assembled
AFM
monolayer (SAM) utilized in a key quantity for this approach. The experimental determination of 𝐸 for SAMs of
Molecular junctions
CP-AFM junctions fabricated using oligophenylene dithiols (OPDn, 1 ≤ 𝑛 ≤ 4) and gold electrodes turned out to
Nanoelectronics
Charge transport be too challenging. Recent measurements (Z. Xie et al, J. Am. Chem. Soc. 139 (2017) 5696) merely succeeded
to provide a low bound estimate (𝐸 ≈ 58 GPa). Supplying this missing experimental information is the aim of
the present theoretical investigation. Our microscopic calculations yield values 𝐸 ≈ 240 ± 6 GPa for the OPDn
SAMs of the aforementioned experimental study, which are larger than those of steel (𝐸 ≈ 180 − 200 GPa) and
silicon (𝐸 ≈ 130 − 185 GPa). The fact that the presently computed 𝐸 is much larger than the aforementioned
experimental lower bound explain why experimentally measuring 𝐸 of OPDn SAM’s is so challenging. Having
𝐸 ≈ 337 ± 8 GPa, OPDn SAMs with herringbone arrangement adsorbed on fcc (111)Au are even stiffer than Si3 N4
(𝐸 ≈ 160 − 290 GPa).

1. Introduction Given the impossibility of a direct determination from experiment,


combining contact mechanics [32–34] to obtain the contact area 𝐴
Among the various platforms to fabricate molecular junctions [1– with Rutherford backscattering (RBS) and/or nuclear reaction analysis
11], conducting probe atomic force microscopy (CP-AFM) [12–24] is an (NRA) to measure the surface coverage Σ is the state-of-the-art approach
approach pioneered by Frisbie’s group [12] that offers a series of ad- [25,31,35–39] to estimate 𝑁 = Σ𝐴.
vantages [25]. CP-AFM junctions consist of bundles containing a num- While the aforementioned nuclear methods enable the direct deter-
ber 𝑁 of molecules trapped between the metal-coated cantilever (AFM mination of the surface coverage Σ, models developed in contact me-
tip) and substrate covered by a self-assembled monolayer (SAM). The chanics pose certain problems to reliably estimate the contact area 𝐴
number of molecules 𝑁 contained in a junction and the related effec- between the AFM tip and the self-assembled monolayer (SAMs) under
tive contact area 𝐴 are an important subject in molecular electronics investigation. To deduce 𝐴, various models were developed in contact
research. The key role played by these quantities was thoroughly ex- mechanics [33,34,40–43]. To apply these models, SAM’s Young modu-
amined in the literature [26,27]. Without information on 𝑁 and 𝐴, a lus of elasticity 𝐸 is a key quantity needed. For significant previous work
comparison between transport properties of molecular junctions fab- on this quantity, see ref. [33] and citations therein. Studies focused on
ricated with different architectures is impossible. Comparing CP-AFM- 𝐸 of SAMs fabricated using alkane and alkane-based molecules of in-
based transport properties (e.g., low bias conductance 𝐺 or current 𝐼) terest for molecular electronics also exist [44–48]. Those studies could
with those measured, e.g., for single molecule STM (scanning tunnel- provide valuable raw data for the contact area 𝐴; unlike in refs. [25,31],
ing microscopy) break junctions or fabricated using crossed wire tech- to estimate 𝑁, assumptions had to be made there on SAMs coverage Σ,
niques [18,19,28–30] makes sense only if properties per molecule (i.e, a value not known from experiment. Returning to OPDn SAMs, as noted
𝐺1 ≡ 𝐺∕𝑁 or 𝐼1 ≡ 𝐼∕𝐴) can be estimated. The same applies to situations earlier [25] and elaborated in Section 2, the experimental determination
wherein a given physical junction is placed in different external condi- of 𝐸 for OPDn SAMs is unfortunately problematic.
tions. A mechanical force 𝐹 stretching a CP-AFM junction does not only So, the best one can do at present is to theoretically compute 𝐸 by
change 𝐺 = 𝐺(𝐹 ) but also 𝑁 = 𝑁(𝐹 ) [31]. To adequately quantify the investigating nanoelastic properties of the molecules of interest subject
impact of stretching, the conductance per molecule 𝐺1 (𝐹 ) = 𝐺(𝐹 )∕𝑁(𝐹 ) to compressive or tensile forces. This is the aim of the present theoreti-
should be analyzed.

E-mail address: ioan.baldea@pci.uni-heidelberg.de

https://doi.org/10.1016/j.apsadv.2021.100094
Received 2 October 2020; Received in revised form 24 December 2020; Accepted 16 March 2021
2666-5239/© 2021 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/)
I. Bâldea Applied Surface Science Advances 5 (2021) 100094

Fig. 1. Cartoon schematically depicting the indentation caused by a compres-


sive force in cases (a) of an AFM tip much stiffer than the SAM, (b) of an AFM tip
and SAM with comparable stiffness, and (c) of a SAM much stiffer than the AFM
tip. (Notice that, unlike Fig. 7, this figure implicitly assumes a SAM adsorbed
on an indeformable substrate.)

cal investigation on molecules OPDn≡ HS-(C6 H4 )𝑛 -SH (1 ≤ 𝑛 ≤ 4) of the


benchmark oligophenylene dithiol family [1,25,35,49–52].
While primarily aiming at providing values of 𝐸 to be subsequently
utilized in ongoing molecular electronics studies, the results reported
below also provide an explanation why a direct experimental determi-
nation of Young’s modulus of SAMs based on aromatic oligophenylene
molecules is currently too challenging. Naively, one could expect that
Fig. 2. Schematic representation of the method to compute the elastic constant
organic substances are not very stiff and hence a behavior sketched
of a molecule.
in Fig. 1a. For this reason, the lower bound estimate (Young modulus
𝐸 ≈ 58 GPa not much smaller than 𝐸𝐴𝑢 = 79 GPa) deduced in recent
experiments [25], — corresponding to Fig. 1b — could already be con- For each OPDn and OPn molecule considered, we first determined
sidered surprisingly large. What we are going to report below is even the molecular length 𝐿0 at equilibrium (zero external force, 𝐹 = 0).
more surprising and resembles to Fig. 1c. We found that OPDn SAMs Then, to simulate the effect of say, a tensile force, we optimized the
are much stiffer than recent experiments [25] have suggested. They are molecular geometry by fixing the molecular length 𝐿 at values larger
stiffer than steel and silicon, possibly also stiffer than Si3 N4 . than 𝐿0 , 𝐿0 < 𝐿1 < 𝐿2 < 𝐿3 < … , and determined the magnitude of
forces 𝐹1 < 𝐹2 < 𝐹3 < … exerted on the atoms at the molecular ends.
The calculated dependence of 𝐹𝑖 on 𝑥𝑖 ≡ 𝐿𝑖 − 𝐿0 turned out to be in-
2. Method deed linear (Hooke’s law applies), and this allowed to extract the spring
(force) constant 𝜅(= 𝐹𝑖 ∕𝑥𝑖 ).
We performed geometry optimizations for molecules subject to axial
mechanical forces (cf. Fig. 2). Quantum chemical computations carried 3. Results and discussion
out to this aim used the GAUSSIAN 16 package [53]. They were based
on the density functional theory using the Becke three-parameter B3LYP 3.1. Preliminary remarks
hybrid exchange-correlation functional [54,55]. For consistency with
our recent [56–59] and ongoing works on related systems, we used 6- As anticipated in Introduction, SAM’s modulus of elasticity 𝐸 ≡
311++G(D, P) basis sets [60,61] although including diffuse basis func- 𝐸𝑠 represents an important challenge for experimentalists. Insufficient
tions is in fact necessarily required only at larger molecular elongations signal-to-noise ratio makes it impossible to reliable estimate an effec-
beyond the linear (elastic) deformation regime investigated in this pa- tive Young modulus for OPDn/Au from tiny differences in indentation
per. depths measured with (OPDn/Au) and without (bare Au) SAM adsorbed

2
I. Bâldea Applied Surface Science Advances 5 (2021) 100094

on gold. To this, from a more general perspective, one should still add
the aggravating point that the exact determination of 𝐸 merely from
force-distance curves is impossible (see ref. [33], page 41 for details):
the mutual dependence between the slope of the contact line and the
jump-off-contact is expressed in terms of a parameter (𝜆 in ref. [25]) for
which (among other nontrivial things) information on SAM’s modulus
of elasticity 𝐸 is required [33]. see, e.g., ref. [33].
To avoid misunderstandings, one should emphasize here that the
aforementioned 𝐸 = 𝐸𝑠 as used in contact mechanics calculations is a
property of the SAM, which does not include the elastic interaction with
the AFM tip. Contact mechanics does account for the tip-SAM interaction
(as it should do), but it does it via the work of adhesion (𝛾 in ref. [25]).
The so-called “effective” Young’s modulus entering formulas of the var-
ious contact mechanics models [33,34,40–43] depends both on SAM’s
(𝐸 ≡ 𝐸𝑠 ) and tip’s (𝐸𝑡𝑖𝑝 ) moduli of elasticity; still, these are “intrinsic”
Young’s moduli, which refer to tip and SAM isolated (i.e. elastically de-
coupled) of each other.
Having said this, and given the fact that elastic properties of met-
als (gold in ref. [25]) used for (coating) AFM tips are known, what we
have to do is to focus on the elastic properties of monolayers of OPDn
molecules which, albeit not (mechanically) coupled to the AFM tip, still
have the same spatial structure as the real SAM.
In this context, one should note that OPDn molecules in SAMs uti-
lized to fabricate the recently investigated CP-AFM junctions [25,35] do
not resemble to those of ordinary (organic) solids or liquids. Both (el-
lipsometry and XPS [25,35,38]) experiments and theory [62–64] found
that OPDn molecules stand nearly vertical on metal. The average inter-
molecular spacing deduced from the coverage (Σ ≃ 3.3 molec/nm2 [25]) Fig. 3. DFT calculations [62–66] revealed that the atoms belonging to the two
amounts to 𝑑 ∼ 5.5 Å. crystallographically nonequivalent OPD2 molecules the unit cell of a densely
packed SAM with herringbone arrangement are sufficiently apart of each other.
In fact, the above average value 𝑑 ∼ 5.5 Å is sufficiently larger than
typical interatomic distances in organic solids to justify the main “ap-
proximation” of the present approach, i.e., to estimate 𝐸 based on calcu- an even less important role played by intermolecular interactions in the
lations to single molecules. Still, the reader less familiar with intermolec- latter case, which is our main focus here.
ular interactions in organic SAMs may miss additional support. This can To sum up, neither covalent nor hydrogen bonding or van der Waals
be inferred from recent microscopic studies of realistic, infinite, peri- interactions can be source of significant mechanical coupling between
odic OPDn SAMs [62–66]. They showed that the unit cell of periodic neighboring OPDn molecules, which are neutral and have no dipole
densely packed OPD2 SAMs with herringbone (hb) ordering comprises moment. So, we are dealing with monolayers wherein parallel OPDn
two crystallographically nonequivalent OPD2 molecules (Fig. 3). Results molecules are sufficiently apart of each other, and considering elastic
reported there: properties of strands of OPDn molecules isolated of each other is a legit-
(a) rule out any conventional chemical (covalent or ionic) bonds be- imate description.
tween neighboring molecules in the SAM: they are neutral (as re- Parenthetically, this one dimensional picture in SAM’s transverse di-
vealed by the atomic (Bader [67]) charges computed [64,66]), have rection is additionally supported by other pieces of recent experimental
neither net dipole moments nor unpaired electrons or lone pair elec- evidence [25,35] revealing a charge transport through individual OPDn
trons; moisolated isolated molecules weakly interacting among themselves in
(b) show that hydrogen bonding cannot come into question, either. transverse direction.
First, the closest (neutral H) atoms in the unit cell (cf. Fig. 3) are
spatially sufficiently well separated of each other; compare 𝑑𝐻𝐻 ≈ 3.2. Elastic properties of OPDn and OPn molecules
3 Åwith the significantly smaller van der Waals (vdW) radius 𝑑𝑣𝑑𝑊𝐻 =
The foregoing analysis made it clear that what we ultimately need
1.06 Å[68,69]. Second, adjacent OPDn molecules cannot share H
to compute are elastic properties of isolated OPDn molecules.
atoms; the H atoms closest to each other form covalent bonds with
Constrained optimization imposing a fixed values of the distance
C atoms which are neither electronegative nor electropositive. Re-
𝐿(X1 , X2 ) → 𝐿 = 𝐿0 + 𝑥 between the two (non-hydrogen) X1,2 atoms at
member that, in spite of the comparable atomic weight (18 versus
the two molecular ends (X1,2 =S for OPDn and X1,2 =C for OPn) larger
16), water and methane have very different boiling points(=measure
or smaller than the equilibrium value (𝐿0 ) straightforwardly allows the
of intermolecular interaction strengths): +100◦ C versus −161◦ C.
determination of the tensile or compressive forces 𝐹 = 𝐹 (𝑥) (of opposite
Water (whose O atom is electronegative) can form hydrogen bonds
direction and equal magnitude) exerted on the X1 and X2 end atoms.
while methane (whose C atom is neither electronegative nor elec-
Results of these quantum chemical calculations at small deforma-
tropositive) cannot;
tions are depicted in Figs. 4 and 5. The numerical values underlying
(c) also rule out significant intermolecular van der Waals interactions.
these figures are collected in Table 1. They reveal a linear dependence
This was explicitly showed in refs. [62,63], which compared elec-
(Hooke’s law) on 𝑥 of the force 𝐹 producing molecule’s mechanical de-
tronic structures of infinite two-dimensional OPDn SAMs computed
formation (cf. Fig. 4)
via DFT without and with vdW dispersion corrections.
𝐹 = 𝜅𝑥 (1)
Noteworthily, Fig. 3 and the results from the previous studies [62–
66] listed above pertain to OPDn SAMs with herringbone periodic- Its slope provides us with the molecule’s elastic (spring) constant 𝜅. As
ity, which are more densely packed than those mainly envisaged here visible in Fig. 4a, (straight) lines for longer OPDn molecules have larger
[25] (Σℎ𝑏 = 4.6 molec/nm2 versus Σ𝑒𝑥𝑝 = 3.3 molec/nm2 ). This implies slopes. This is in accord with the fact that, at a given deforming force

3
I. Bâldea Applied Surface Science Advances 5 (2021) 100094

Fig. 4. Results for the mechanical (here, 𝐹 > 0, tensile or compressing 𝐹 < 0) force as a function of molecular length 𝐿: (a) OPDn and (b) OPn.

Fig. 5. Results for the mechanical (here, 𝐹 > 0, tensile or compressing 𝐹 < 0) force as a function of molecular length 𝐿: (a) OPDn and (b) OPn.

Table 1 For comparison purposes, in Figs. 4b and 5 b we also present results


Elastic properties of OPDn and OPn molecules calculated microscopically as in- for the elastic properties of the parent (non-thiolated) molecules of the
dicated in the main text: spring constants (𝜅), specific stresses 𝐾, and Young’s oligophenylene series OPn≡ H-(C6 H4 )𝑛 -H (1 ≤ 𝑛 ≤ 4). The trend across
moduli 𝐸. The values of the Young’s modulus 𝐸 correspond to two values of the
the OPDn family noted above is also visible for OPn. Notice, in par-
surface coverages Σ: (a) value measured via Rutherford backscattering (RBS)
ticular, that OP1 (≡ benzene) has the lowest specific stress; the stiffest
and nuclear reaction analysis (NRA) for OPDn SAMs used to fabricate molecu-
lar CP-AFM junctions in ref. [25] (Σ = 3.3 molec/nm2 ) and (b) value for OPDn
interring C-C bond is missing there.
SAMs with herringbone structure (Σ = 4.63 molec/nm2 , cf. ref [62] and citations A variety of carbon-based molecular families M𝑛 [70–74] are known
therein). to exhibit even-odd alternation upon adding a repeat unit (phenyl ring
in our case) M𝑛 → M𝑛+1 . This is a particularly meaningful question be-
𝑎 𝑎 𝑏
Molecule 𝜅𝑂𝑃 𝐷𝑛 𝜅𝑂𝑃 𝑛 𝐾𝑂𝑃 𝐷𝑛 𝐾𝑂𝑃 𝑛 𝐸𝑂𝑃 𝐸𝑂𝑃 𝐸𝑂𝑃
𝐷𝑛 𝑛 𝐷𝑛 cause symmetry is found to alternate across the OPDn family; calcula-
OPD1 10.658 21.488 68.069 59.939 224.6 197.8 315.1 tions indicate that the odd-number members OPD1 and OPD3 possess
OPD2 6.755 9.856 72.331 70.154 238.7 231.5 334.9 C𝑖 symmetry while even-number members OPD2 and OPD4 possess C2
OPD3 4.997 6.511 75.137 74.525 248.0 245.9 347.9
symmetry. Nevertheless, the results depicted in Fig. 6 reveal that this is
OPD4 3.925 4.829 76.003 76.176 250.8 251.4 351.9
not the case. The dependence on 𝑛 is monotonic. Guided by the number
of OPDn and OPn constituents, we were led to consider an analytical
formula used for the fitting curves presented in Fig. 6, which allows us
𝐹 , homogeneous springs characterized by a length-independent specific to put the aforementioned monotonic dependence on 𝑛 in more quanti-
stress 𝐾 tative terms.
𝑥 Like in cases of the rods or bars schematically depicted in Fig. 2b,
𝐹 =𝐾 ; 𝐾 ≡ 𝜅𝐿0 (2)
𝐿0 for the characterization in terms of Young’s modulus of elasticity 𝐸 via
Hooke’s law, a cross section area 𝐴 needs to be assigned
respond with elongations proportional to their length: 𝑥 ∝ 𝐿0 , 𝜅 ∝ 1∕𝐿0 .
In fact, OPDn molecules do not behave like homogeneous springs be- 𝐴 𝜅𝐿0 𝐾
𝐹 = 𝐸 𝑥; 𝐸 = = (3)
cause the stiffness of their constituents is different: in the linear elastic 𝐿0 𝐴 𝐴
regime presently considered (i.e. at small 𝐹 ), the interring C-C bonds ⏟⏟⏟
𝜅
are notably stiffer than the aromatic phenyl rings, which are in turns
somewhat stiffer than the terminal C-S bonds. In the specific case considered, the effective cross section area can be
This results in values of the specific stress 𝐾 slightly decreasing from expressed in terms of the surface coverage Σ (cf. Fig. 2b). Thanks to
the longer OPD4 to the shorter OPD1, as observable by inspecting the RBS and NRA, for OPDn SAMs used to fabricate the CP-AFM junctions
slopes of the lines depicted in Fig. 5a. of ref. [25], this quantity is available: Σ ≈ 3.3 nm−2 [25]. This yields an

4
I. Bâldea Applied Surface Science Advances 5 (2021) 100094

Fig. 6. Results for the specific strengths 𝐾 (cf. eq (2)) and Young’s moduli 𝐸 (panels a and b, respectively) of oligophenylene dithiol molecules (OPDn) and their
parent (OPn) non-thiolated species calculated microscopically as described in the main text and fitted as indicated in the inset.

average area per molecule

𝐴 = 1∕Σ ≈ 30 nm2 (4)

The values of the Young’s moduli deduced via Eqs. (3) and (4) are
presented in Table 1. They show that the SAMs used to fabricate OPDn
CP-AFM junctions of ref. [25] are very stiff. They are stiffer than steel
(180 GPa for stainless AISI 302 and 200 GPa for structural ASTM-A36
[75]).
Let us also refer to two materials used for commercial AFM can-
tilevers: silicon and silicon nitride. These materials dominate the vast
majority of applications [33]. OPDn SAMs are stiffer than silicon: 𝐸Si ≈
130 − 185 GPa [33]. (Precise values of real materials depend on various
factors, e.g., precise composition and crystallographic orientation). Val-
ues for silicon nitride are in the range 𝐸Si3 N4 ≈ 160 − 290 GPa [33]. So,
using the coverages measured in ref. [25] we can also conclude that the
OPDn SAMs of ref. [25] can be even stiffer than silicon nitride.
Yet, this is not the whole issue. Notwithstanding that the OPDn
SAMs fabricated on polycrystalline gold in ref. [25] were found to be
characterized by exceptionally small statistical variations (hence im-
plicitly good coverage), the surface coverage of OPDn SAMs with her-
ringbone (hb) arrangement adsorbed on fcc Au (111) is higher (Σℎ𝑏 =
4.63 molec/nm2 [62]) than that of ref. [25] (Σ ≈ 3.3 molec/nm2 ). This
(hb) coverage corresponds to even higher values of 𝐸, which are also
included in Table 1. Inspection of the last column of Table 1 reveals that
OPDn SAMs with herringbone structure are definitely stiffer than Si3 N4 .
The above finding makes experimentalists’ difficulty anticipated in
Introduction understandable. They attempted to determine SAM’s 𝐸 by
subtracting the measured indentation 𝑥0𝐴𝑢 ≡ 𝑥0𝑡𝑖𝑝 + 𝑥0𝑠𝑢𝑏𝑠𝑡𝑟𝑎𝑡𝑒 of the bare
metal (i.e., without SAM, superscript 0) subject to a mechanical force
𝐹 from the indentation 𝑥𝑆𝐴𝑀∕𝐴𝑢 ≡ 𝑥𝑡𝑖𝑝 + 𝑥𝑆𝐴𝑀 + 𝑥𝑠𝑢𝑏𝑠𝑡𝑟𝑎𝑡𝑒 measured by
applying the same force 𝐹 on the SAM adsorbed on gold.
As sketched in Fig. 7, a mechanical force applied via AFM tip to
Fig. 7. As illustrated by means of two elastic springs in series (panel a), a me-
an OPDn SAM formed on gold substrate is similar to a force 𝐹 act-
chanical force applied to a stiff SAM yields an indentation hard to distinguish
ing on a stiff elastic spring (read SAM, 𝐸 ≈ 240 GPa) in series with a
from that in the absence of the SAM (compare panels b and c).
soft elastic spring (read gold tip and substrate in series, 𝐸𝐴𝑢 = 79 GPa).
The deformation of the latter is much larger than that of the former.
Rephrasing, the AFM tip does not significantly penetrate the SAM. The
interposed SAM basically serves to transmit the action of 𝐹 to the sub-
strate, which can be deformed if its Young’s modulus is small enough 4. Conclusion
— 𝐸𝑠𝑢𝑏𝑠𝑡𝑟𝑎𝑡𝑒 = 𝐸𝑡𝑖𝑝 ≡ 𝐸𝐴𝑢 ≪ 𝐸𝑆𝐴𝑀 —, as the case in experiment [25]. The
total deformation (“indentation”) 𝑥𝑆𝐴𝑀∕𝐴𝑢 cannot be distinguished from To the best of author’s knowledge, this is the first work reporting
the deformation 𝑥0𝐴𝑢 of the soft spring alone subject to the same force results for Young’s moduli of elasticity of SAMs based on benchmark
𝐹 ; distinguishing the tiny difference |𝑥𝑆𝐴𝑀∕𝐴𝑢 − 𝑥0𝐴𝑢 | ≈ 0 from zero re- species (as the case of the presently considered family of oligophenylene
quires a very good signal-to-noise ratio, which can hardly be reached in dithiols OPDn) routinely used for fabricating CP-AFM molecular junc-
experiment. tions. The fact that the present values 𝐸OPDn ≈ 240 ± 6 GPa are much

5
I. Bâldea Applied Surface Science Advances 5 (2021) 100094

𝑒𝑥𝑝
larger than the lower bound 𝐸OPDn ≈ 58 GPa extracted from recent ex- [11] Molecular Electronics: An Experimental and Theoretical Approach, I. Bâldea (Ed.),
periments [25] is noteworthy. Pan Stanford, 2015, doi:10.4032/9789814613910.
[12] D.J. Wold, C.D. Frisbie, Formation of metal-molecule-metal tunnel junctions: Micro-
On one side, it provides a rationale for the difficulty encountered by contacts to alkanethiol monolayers with a conducting AFM tip, J. Am. Chem. Soc.
experimentalists to measure a reliable value of 𝐸 [25]; the OPDn SAM 122 (12) (2000) 2970–2971, doi:10.1021/ja994468h.
high stiffness makes the deformation measurable for OPDn/Au in excess [13] D.J. Wold, C.D. Frisbie, Fabrication and characterization of metal-molecule-metal
junctions by conducting probe atomic force microscopy, J. Am. Chem. Soc. 123 (23)
to that for bare gold too small to be accurately estimated. (2001) 5549–5556, doi:10.1021/ja0101532. PMID: 11389638
On the other side and more importantly from a practical standpoint, [14] J.M. Beebe, V.B. Engelkes, L.L. Miller, C.D. Frisbie, Contact resistance in metal-
the presently reported values of 𝐸 make it possible to update/refine the molecule-metal junctions based on aliphatic sams: effects of surface linker
and metal work function, J. Am. Chem. Soc. 124 (38) (2002) 11268–11269,
numbers 𝑁 of OPDn molecules per junction. At a given AFM tip and
doi:10.1021/ja0268332. PMID: 12236731
load (typically compressive 𝐹 = 1 nN) which is necessarily applied to [15] D.J. Wold, R. Haag, M.A. Rampi, C.D. Frisbie, Distance dependence of electron tun-
render conduction through CP-AFM junctions possible, the contact area neling through self-assembled monolayers measured by conducting probe atomic
force microscopy: unsaturated versus saturated molecular junctions, J. Phys. Chem.
𝐴 for stiffer SAMs (read 𝐸OPDn ≈ 240 ± 6 GPa, the presently reported
B 106 (11) (2002) 2813–2816, doi:10.1021/jp013476t.
values, cf. Table 1) is significantly smaller than for softer SAMs (read [16] A. Salomon, D. Cahen, S. Lindsay, J. Tomfohr, V. Engelkes, C. Frisbie, Comparison
𝑒𝑥𝑝
𝐸OPDn ≈ 58 GPa, as used in ref. [25]). of electronic transport measurements on organic molecules, Adv. Mater. 15 (22)
Work is underway, but what one can state for sure is that the en- (2003) 1881–1890, doi:10.1002/adma.200306091.
[17] J.M. Beebe, V.B. Engelkes, J. Liu, J.J. Gooding, P.K. Eggers, Y. Jun, X. Zhu,
suing values of 𝑁 will be even smaller than those (𝑁 ∼ 80) estimated M.N. Paddon-Row, C.D. Frisbie, Length dependence of charge transport in
in ref. [25], which, surprisingly, already appeared much smaller and nanoscopic molecular junctions incorporating a series of rigid thiol-terminated nor-
reproducible than those (up to 𝑁 ∼ 103 [21]) claimed in earlier litera- bornylogs, J. Phys. Chem. B 109 (11) (2005) 5207–5215, doi:10.1021/jp044630p.
PMID: 16863186
ture [21] on CP-AFM junctions. Further, since what is directly accessible [18] J.M. Beebe, B. Kim, J.W. Gadzuk, C.D. Frisbie, J.G. Kushmerick, Transition from
in CP-AFM experiments is a junction’s transport property (e.g., conduc- direct tunneling to field emission in metal-molecule-metal junctions, Phys. Rev. Lett.
tance 𝐺𝑗𝑢𝑛𝑐 , current 𝐼𝑗𝑢𝑛𝑐 ), updated/refined 𝑁s’ also translate into up- 97 (2) (2006) 026801, doi:10.1103/PhysRevLett.97.026801.
[19] J.M. Beebe, B. Kim, C.D. Frisbie, J.G. Kushmerick, Measuring relative barrier heights
dated/refined values of transport property per molecule (𝐺1 = 𝐺𝑗𝑢𝑛𝑐 ∕𝑁, in molecular electronic junctions with transition voltage spectroscopy, ACS Nano 2
𝐼1 = 𝐼𝑗𝑢𝑛𝑐 ∕𝑁). This, in turn, makes a quantitative comparison with trans- (5) (2008) 827–832, doi:10.1021/nn700424u.
port properties measured using single-molecule (e.g., STM) testbeds [20] B. Kim, J.M. Beebe, Y. Jun, X.-Y. Zhu, C.D. Frisbie, Correlation between homo
alignment and contact resistance in molecular junctions: aromatic thiols ver-
more adequate.
sus aromatic isocyanides, J. Am. Chem. Soc. 128 (15) (2006) 4970–4971,
doi:10.1021/ja0607990.
[21] H.B. Akkerman, B. de Boer, Electrical conduction through single molecules and self-
Declaration of Competing Interest -assembled monolayers, J. Phys.: Condens. Matt. 20 (1) (2008) 013001.
[22] A. Tan, S. Sadat, P. Reddy, Measurement of thermopower and current-voltage char-
The authors declare that they have no known competing financial acteristics of molecular junctions to identify orbital alignment, Appl. Phys. Lett. 96
(1) (2010) 013110, doi:10.1063/1.3291521.
interests or personal relationships that could have appeared to influence
[23] B. Kim, S.H. Choi, X.-Y. Zhu, C.D. Frisbie, Molecular tunnel junctions based on 𝜋-
the work reported in this paper. conjugated oligoacene thiols and dithiols between ag, au, and pt contacts: Effect of
surface linking group and metal work function, J. Am. Chem. Soc. 133 (49) (2011)
19864–19877, doi:10.1021/ja207751w.
Acknowlgedgments [24] L. Luo, S.H. Choi, C.D. Frisbie, Probing hopping conduction in conjugated molec-
ular wires connected to metal electrodes, Chem. Mater. 23 (3) (2011) 631–645,
doi:10.1021/cm102402t.
The author thanks Dan Frisbie and Zuoti Xie for providing him
[25] Z. Xie, I. Bâldea, A.T. Demissie, C.E. Smith, Y. Wu, G. Haugstad, C.D. Frisbie, Ex-
unpublished experimental data. Financial support from the Deutsche ceptionally small statistical variations in the transport properties of metal-molecule-
Forschungsgemeinschaft (DFG grant BA 1799/3-2) and computational metal junctions composed of 80 oligophenylene dithiol molecules, J. Am. Chem.
support from the State of Baden-Württemberg through bw HPC/DFG Soc. 139 (16) (2017) 5696–5699, doi:10.1021/jacs.7b01918. PMID: 28394596
[26] F.C. Simeone, H.J. Yoon, M.M. Thuo, J.R. Barber, B. Smith, G.M. Whitesides, Defin-
through grants INST 40/467-1 FUGG and INST 40/575-1 FUGG ing the value of injection current and effective electrical contact area for egain-based
(JUSTUS-2) in the initial stage of this research is gratefully acknowl- molecular tunneling junctions, J. Am. Chem. Soc. 135 (48) (2013) 18131–18144,
edged. doi:10.1021/ja408652h. PMID: 24187999
[27] S. Park, S. Kang, H.J. Yoon, Power factor of one molecule thick films and
length dependence, ACS Central Sci. 5 (12) (2019) 1975–1982, doi:10.1021/acs-
References centsci.9b01042.
[28] J.G. Kushmerick, D.B. Holt, S.K. Pollack, M.A. Ratner, J.C. Yang, T.L. Schull,
[1] M.A. Reed, C. Zhou, C.J. Muller, T.P. Burgin, J.M. Tour, Conductance of a molecular J. Naciri, M.H. Moore, R. Shashidhar, Effect of bond-length alternation in molecular
junction, Science 278 (5336) (1997) 252–254, doi:10.1126/science.278.5336.252. wires, J. Am. Chem. Soc. 124 (36) (2002) 10654–10655, doi:10.1021/ja027090n.
[2] B. Xu, N.J. Tao, Measurement of single-molecule resistance by repeated formation [29] J.G. Kushmerick, D.B. Holt, J.C. Yang, J. Naciri, M.H. Moore, R. Shashidhar,
of molecular junctions, Science 301 (5637) (2003) 1221–1223, doi:10.1126/sci- Metal-molecule contacts and charge transport across monomolecular layers: Mea-
ence.1087481. surement and theory, Phys. Rev. Lett. 89 (2002) 086802, doi:10.1103/Phys-
[3] E. Lörtscher, H.B. Weber, H. Riel, Statistical approach to investigating transport RevLett.89.086802.
through single molecules, Phys. Rev. Lett. 98 (2007) 176807, doi:10.1103/Phys- [30] J.G. Kushmerick, Metal-molecule contacts, Mater. Today 8 (7) (2005) 26–30,
RevLett.98.176807. doi:10.1016/S1369-7021(05)70984-6.
[4] L. Venkataraman, J.E. Klare, C. Nuckolls, M.S. Hybertsen, M.L. Steigerwald, Depen- [31] Z. Xie, I. Bâldea, G. Haugstad, C.D. Frisbie, Mechanical deformation distinguishes
dence of single-molecule junction conductance on molecular conformation, Nature tunneling pathways in molecular junctions, J. Am. Chem. Soc. 141 (1) (2019) 497–
442 (7105) (2006) 904–907, doi:10.1038/nature05037. 504, doi:10.1021/jacs.8b11248.
[5] R.L. McCreery, A.J. Bergren, Progress with molecular electronic junctions: Meeting [32] K.L. Johnson, Contact Mechanics, Cambridge University Press, 1985,
experimental challenges in design and fabrication, Adv. Mater. 21 (43) (2009) 4303– doi:10.1017/CBO9781139171731.
4322, doi:10.1002/adma.200802850. [33] H.-J. Butt, B. Cappella, M. Kappl, Force measurements with the atomic force mi-
[6] C.A. Martin, R.H.M. Smit, H.S.J. van der Zant, J.M. van Ruitenbeek, A na- croscope: technique, interpretation and applications, Surf. Sci. Rep. 59 (1) (2005)
noelectromechanical single-atom switch, Nano Lett. 9 (8) (2009) 2940–2945, 1–152, doi:10.1016/j.surfrep.2005.08.003.
doi:10.1021/nl901355y. PMID: 19639963 [34] G. Haugstad, Atomic Force Microscopy, John Wiley & Sons, NJ, 2012,
[7] H. Song, Y. Kim, Y.H. Jang, H. Jeong, M.A. Reed, T. Lee, Observation of molecular doi:10.1002/9781118360668.
orbital gating, Nature 462 (7276) (2009) 1039–1043, doi:10.1038/nature08639. [35] Z. Xie, I. Bâldea, C. Smith, Y. Wu, C.D. Frisbie, Experimental and theoretical
[8] H. Song, M.A. Reed, T. Lee, Single molecule electronic devices, Adv. Mater. 23 (14) analysis of nanotransport in oligophenylene dithiol junctions as a function of
(2011) 1583–1608, doi:10.1002/adma.201004291. molecular length and contact work function, ACS Nano 9 (8) (2015) 8022–8036,
[9] H.J. Yoon, C.M. Bowers, M. Baghbanzadeh, G.M. Whitesides, The rate of charge doi:10.1021/acsnano.5b01629. PMID: 26190402
tunneling is insensitive to polar terminal groups in self-assembled monolayers [36] A.T. Demissie, G. Haugstad, C.D. Frisbie, Quantitative surface coverage measure-
in AgTSS(CH2)nM(CH2)mT//Ga2O3/EGaIn junctions, J. Am. Chem. Soc. 136 (1) ments of self-assembled monolayers by nuclear reaction analysis of carbon-12,
(2014) 16–19, doi:10.1021/ja409771u. PMID: 24350722 J. Phys. Chem. Lett. 7 (17) (2016) 3477–3481, doi:10.1021/acs.jpclett.6b01363.
[10] J.C. Cuevas, E. Scheer, Molecular Electronics: An Introduction to Theory and Exper- PMID: 27537072
iment, 2nd, World Scientific, 2017, doi:10.1142/10598. World Scientific Series in [37] Z. Xie, I. Bâldea, C.D. Frisbie, Why one can expect large rectification in molecular
Nanoscience and Nanotechnology: Vol. 15

6
I. Bâldea Applied Surface Science Advances 5 (2021) 100094

junctions based on alkane monothiols and why rectification is so modest, Chem. Sci. [54] A.D. Becke, Density-functional exchange-energy approximation with correct asymp-
9 (2018) 4456–4467, doi:10.1039/C8SC00938D. totic behavior, Phys. Rev. A 38 (1988) 3098–3100, doi:10.1103/PhysRevA.38.3098.
[38] Z. Xie, I. Bâldea, C.D. Frisbie, Determination of energy level alignment in molecu- [55] A.D. Becke, A new mixing of Hartree-Fock and local density-functional theories, J.
lar tunnel junctions by transport and spectroscopy: self-consistency for the case of Chem. Phys. 98 (2) (1993) 1372–1377, doi:10.1063/1.464304.
oligophenylene thiols and dithiols on ag, au, and pt electrodes, J. Am. Chem. Soc. [56] I. Bâldea, Evidence that molecules in molecular junctions may not be subject to the
141 (8) (2019) 3670–3681, doi:10.1021/jacs.8b13370. entire external perturbation applied to electrodes, Langmuir 36 (5) (2020) 1329–
[39] Z. Xie, I. Bâldea, C.D. Frisbie, Energy level alignment in molecular tunnel junctions 1337, doi:10.1021/acs.langmuir.9b03430.
by transport and spectroscopy: self-consistency for the case of alkyl thiols and dithi- [57] I. Bâldea, Extensive quantum chemistry study of neutral and charged C4 N
ols on ag, au, and pt electrodes, J. Am. Chem. Soc. 141 (45) (2019) 18182–18192, chains, ACS Earth Space Chem. 4 (3) (2020) 434–448, doi:10.1021/ac-
doi:10.1021/jacs.9b08905. PMID: 31617711 searthspacechem.9b00321.
[40] K.L. Johnson, K. Kendall, A.D. Roberts, D. Tabor, Surface energy and the con- [58] I. Bâldea, Profiling C4 N radicals of astrophysical interest, Mon. Not. R. Astron. Soc.
tact of elastic solids, Proc. Roy. Soc. Lond. A 324 (1558) (1971) 301–313, 493 (2020) 2506–2510, doi:10.1093/mnras/staa455.
doi:10.1098/rspa.1971.0141. JKR model of contact mechanics [59] I. Bâldea, Profiling astrophysically relevant MgC4 H chains. an attempt to aid
[41] B. Derjaguin, V. Muller, Y. Toporov, Effect of contact deformations on astronomical observations, Mon. Not. R. Astron. Soc. 498 (2020) 4316–4326,
the adhesion of particles, J. Colloid Interface Sci. 53 (2) (1975) 314–326, doi:10.1093/mnras/staa2354.
doi:10.1016/0021-9797(75)90018-1. DMT model of contact mechanics [60] G.A. Petersson, A. Bennett, T.G. Tensfeldt, M.A. Al-Laham, W.A. Shirley,
[42] D. Dugdale, Yielding of steel sheets containing slits, J. Mech. Phys. Solids 8 (2) J. Mantzaris, A complete basis set model chemistry. I. The total energies of closed-
(1960) 100–104, doi:10.1016/0022-5096(60)90013-2. Maugis-Dugdale model of shell atoms and hydrides of the first-row elements, J. Chem. Phys. 89 (4) (1988)
contact mechanics 2193–2218, doi:10.1063/1.455064.
[43] D. Maugis, Adhesion of spheres: the JKR-DMT transition using a Dugdale model, J. [61] G.A. Petersson, M.A. Al-Laham, A complete basis set model chemistry. II. Open-shell
Colloid. Interf. Sci. 150 (1) (1992) 243–269, doi:10.1016/0021-9797(92)90285-T. systems and the total energies of the first-row atoms, J. Chem. Phys. 94 (9) (1991)
[44] F.W. DelRio, C. Jaye, D.A. Fischer, R.F. Cook, Elastic and adhesive properties of 6081–6090, doi:10.1063/1.460447.
alkanethiol self-assembled monolayers on gold, Appl. Phys. Lett. 94 (13) (2009) [62] I. Bâldea, A surprising way to control the charge transport in molecular electronics:
131909, doi:10.1063/1.3111440. the subtle impact of the coverage of sam of floppy molecules adsorbed on electrodes,
[45] Y. Zhang, X. Qiu, P. Gordiichuk, S. Soni, T.L. Krijger, A. Herrmann, R.C. Chiechi, Faraday Discuss. 204 (2017) 35–52, doi:10.1039/C7FD00101K.
Mechanically and electrically robust self-assembled monolayers for large- [63] I. Bâldea, A novel route to get functionality in nanoelectronics: controlling the charge
area tunneling junctions, J. Phys. Chem. C 121 (27) (2017) 14920–14928, transport by the subtle impact of the coverage of self-assembled monolayers on the
doi:10.1021/acs.jpcc.7b03853. conformation of floppy molecules adsorbed on metallic electrodes, Appl. Surf. Sci.
[46] R. Henda, Simulation of self-assembled monolayers under normal stress, AIChE J. 472 (2019) i16–21, doi:10.1016/j.apsusc.2018.07.155.
46 (6) (2000) 1275–1279, doi:10.1002/aic.690460620. [64] I. Bâldea, Impact of molecular conformation on transport and transport-
[47] X. Quan, J.D. Madura, G.R. Hutchison, Self-assembled monolayer piezoelectrics: related properties at the nanoscale, Appl. Surf. Sci. 487 (2019) 593–600,
electric-field driven conformational changes, 2017. doi:10.1016/j.apsusc.2019.05.112.
[48] Q. Tu, H.S. Kim, T.J. Oweida, Z. Parlak, Y.G. Yingling, S. Zauscher, Interfacial me- [65] I. Bâldea, Floppy molecules as candidates for achieving optoelectronic molecular
chanical properties of graphene on self-assembled monolayers: Experiments and sim- devices without skeletal rearrangement or bond breaking, Phys. Chem. Chem. Phys.
ulations, ACS Appl. Mater. Interfaces 9 (11) (2017) 10203–10213, doi:10.1021/ac- 19 (2017) 30842–30851, doi:10.1039/C7CP06428D.
sami.6b16593. PMID: 28230343 [66] I. Bâldea, A sui generis electrode-driven spatial confinement effect responsi-
[49] S. Guo, J. Hihath, I. Diez-Pérez, N. Tao, Measurement and statistical analysis ble for strong twisting enhancement of floppy molecules in closely packed
of single-molecule current-voltage characteristics, transition voltage spectroscopy, self-assembled monolayers, Phys. Chem. Chem. Phys. 20 (2018) 23492–23499,
and tunneling barrier height, J. Am. Chem. Soc. 133 (47) (2011) 19189–19197, doi:10.1039/C8CP04974B.
doi:10.1021/ja2076857. [67] R.F.W. Bader, Atoms in Molecules: A Quantum Theory, Oxford University Press: New
[50] A. Tan, J. Balachandran, S. Sadat, V. Gavini, B.D. Dunietz, S.-Y. Jang, P. Reddy, York, 1994.
Effect of length and contact chemistry on the electronic structure and thermoelectric [68] A. Bondi, van der waals volumes and radii, J. Phys. Chem. 68 (3) (1964) 441–451,
properties of molecular junctions, J. Am. Chem. Soc. 133 (23) (2011) 8838–8841, doi:10.1021/j100785a001.
doi:10.1021/ja202178k. [69] S.S. Batsanov, Van der waals radii of elements, Inorganic Materials 37 (2001) 871–
[51] A. Tan, J. Balachandran, B.D. Dunietz, S.-Y. Jang, V. Gavini, P. Reddy, Length de- 885, doi:10.1023/A:1011625728803.
pendence of frontier orbital alignment in aromatic molecular junctions, Appl. Phys. [70] Q. Fan, G.V. Pfeiffer, Theoretical study of linear C𝑛 (n=6-10) and HC𝑛
Lett. 101 (24) (2012) 243107, doi:10.1063/1.4769986. H (n=2-10) molecules, Chem. Phys. Lett. 162 (6) (1989) 472–478,
[52] S. Guo, G. Zhou, N. Tao, Single molecule conductance, thermopower, and transition doi:10.1016/0009-2614(89)87010-1.
voltage, Nano Lett. 13 (9) (2013) 4326–4332, doi:10.1021/nl4021073. [71] F. Tao, S.L. Bernasek, Understanding odd-even effects in organic self-assembled
[53] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheese- monolayers, Chem. Rev. 107 (5) (2007) 1408–1453, doi:10.1021/cr050258d. PMID:
man, G. Scalmani, V. Barone, G.A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A.V. 17439290
Marenich, J. Bloino, B.G. Janesko, R. Gomperts, B. Mennucci, H.P. Hratchian, J.V. [72] L. Yuan, D. Thompson, L. Cao, N. Nerngchangnong, C.A. Nijhuis, One carbon mat-
Ortiz, A.F. Izmaylov, J.L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. ters: the origin and reversal of odd-even effects in molecular diodes with self-
Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V.G. Zakrzewski, assembled monolayers of Ferrocenyl-Alkanethiolates, J. Phys. Chem. C 119 (31)
J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, (2015) 17910–17919, doi:10.1021/acs.jpcc.5b04797.
J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. [73] Z. Wu, S. Xu, H. Lu, A. Khamoshi, G.-B. Liu, T. Han, Y. Wu, J. Lin, G. Long, Y. He,
Throssell, J. J. A. Montgomery, J.E. Peralta, F. Ogliaro, M.J. Bearpark, J.J. Heyd, Y. Cai, Y. Yao, F. Zhang, N. Wang, Even-odd layer-dependent magnetotransport
E.N. Brothers, K.N. Kudin, V.N. Staroverov, T.A. Keith, R. Kobayashi, J. Normand, of high-mobility q-valley electrons in transition metal disulfides, Nat. Commun. 7
K. Raghavachari, A.P. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, Cossi, M. Millam, (2016) 12955, doi:10.1038/ncomms12955.
Klene, C. Adamo, R. Cammi, J.W. Ochterski, R.L. Martin, K. Morokuma, O. Farkas, [74] I. Bâldea, Alternation of singlet and triplet states in carbon-based chain molecules
J.B. Foresman, D.J. Fox, Gaussian, Inc., Wallingford ct, Gaussian 16, revision b.01, and its astrochemical implications: Results of an extensive theoretical study, Adv.
2016. Theor. Simul. 2 (9) (2019) 1900084, doi:10.1002/adts.201900084.
[75] Cf. https://www.engineeringtoolbox.com/young-modulus-d_417.html.

You might also like