London Dispersion Directs On-Surface Self-Assembly of [121

Aug 28, 2017 - London dispersion (LD) acts between all atoms and molecules in nature, but the role of LD interactions in the self-assembly of molecula...
3 downloads 6 Views 1MB Size
Subscriber access provided by Grand Valley State | University

Article

London Dispersion Directs On-Surface SelfAssembly of [121]Tetramantane Molecules Daniel Ebeling, Marina Sekutor, Marvin Stiefermann, Jalmar Tschakert, Jeremy E. P. Dahl, Robert M.K. Carlson, Andre Schirmeisen, and Peter R. Schreiner ACS Nano, Just Accepted Manuscript • DOI: 10.1021/acsnano.7b05204 • Publication Date (Web): 28 Aug 2017 Downloaded from http://pubs.acs.org on August 29, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Nano is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

London Dispersion Directs On-Surface SelfAssembly of [121]Tetramantane Molecules Daniel Ebeling,a,‡,* Marina Šekutor,b,‡ Marvin Stiefermann,a Jalmar Tschakert,a Jeremy E. P. Dahl,c Robert M. K. Carlson,c André Schirmeisen,a,* and Peter R. Schreinerb,* a

Institute of Applied Physics, Justus-Liebig University, Heinrich-Buff-Ring 16, 35392

Giessen, Germany, [email protected], [email protected] b

Institute of Organic Chemistry, Justus-Liebig University, Heinrich-Buff-Ring 17, 35392

Giessen, Germany, [email protected] c

Stanford Institute for Materials and Energy Sciences, Stanford, CA 94305, USA



Both contributors are considered first authors.

TOC graphic

1 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 27

KEYWORDS: bond imaging / CO-functionalized tips / diamondoids / low temperature noncontact atomic force microscopy / van der Waals interactions

Abstract. London dispersion (LD) acts between all atoms and molecules in nature, but the role of LD interactions in the self-assembly of molecular layers is still poorly understood. In this study direct visualization of single molecules using atomic force microscopy with COfunctionalized tips revealed exact adsorption structure of bulky and highly polarizable [121]tetramantane molecules on Au(111) and Cu(111) surfaces.

We could therefore

determine the absolute molecular orientation of bulky, completely sp3 hybridized molecules on metal surfaces. Moreover, we demonstrated how LD drives this on-surface self-assembly of [121]tetramantane hydrocarbons, resulting in the formation of a highly ordered 2D lattice. Our experimental findings were underpinned by a systematic computational study, which allowed us to quantify the energies associated with LD interactions and to analyze intermolecular close contacts and attractions in more detail.

To observe a single molecule and recognize its individual atoms is the chemist's ultimate fascination, and astonishingly, what once was only a musing became in recent years an exciting reality. Numerous nanostructures have been prepared using on-surface synthesis on metals, with scanning tunneling microscopy (STM) and atomic force microscopy (AFM) emerging as the main tools for product identification and the analysis of surface phenomena.14

The AFM technique, using a scanning probe tip terminated by a CO molecule, results in a

significant increase of resolution, to the point that individual atoms in molecules can be visualized.5-7 Moreover, the ability to pick up small molecules with the tip provides ways to control single-molecule reactions.8-11 The advances in this field are steadily on the rise,12, 13 2 ACS Paragon Plus Environment

Page 3 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

ranging from elucidation of molecular structures14 to step-wise analysis of intermediates and products formed by on-surface reactions.15-19 Since many self-assembly processes on surfaces are still poorly understood, techniques capable of sub-molecular resolution could help shed some light on the forces behind them. One such force is London dispersion (LD).20,

21

Since LD conceptually arises from

instantaneous induced dipoles, it is independent of interactions between permanent dipoles, it keeps molecules together, and therefore is a key interaction in the formation of condensed matter.22,

23

For increasingly larger and hence more polarizable structures, LD gains more

importance as the attractive effect grows rapidly with molecular size.21 For instance, bulky groups can act as dispersion energy donors (DEDs)24 that stabilize extremely long C–C bonds that would otherwise be fleetingly existent.25 LD interactions have also been reported to play an important role for self-assembly processes of molecular structures on surfaces.26-29 However, in most examples LD coexists side-by-side with electrostatic interactions or hydrogen bonding, making it difficult to assess and quantify the overall LD contributions. We therefore chose [121]tetramantane for our study, which serves as an archetype system to explore LD due to the absence of other interactions, since recent studies showed that hydrocarbons in general are primarily stabilized by the mechanism of oscillating dipoles (in other words by LD), with electrostatic interactions having only of a minor effect.30-32 Our aim was to determine the influence LD has on the self-assembly of bulky tetramantane molecules and to quantify such attractive interactions in two dimensions. [121]Tetramantane belongs to the class of diamondoid compounds, which are hydrogenterminated cage molecules that structurally resemble the diamond crystal lattice (Fig. 1).33, 34 Its first synthesis was accomplished in the 1970s by McKervey et al. albeit in too low a yield for any practical application.35 Recent isolation of diamondoids from oil36 enabled advances in selective functionalization of higher diamondoids37 and deposition of functionalized 3 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 27

[121]tetramantane derivatives on metal surfaces resulted in the formation of self-assembled monolayers with fascinating electronic properties.38-40

Figure 1.

Structures of lower diamondoids and [121]tetramantane, hydrogen-terminated

nanodiamonds.

An important aspect for studying the self-assembly process in detail is a precise identification of molecular orientation. For bulky, completely sp3 hybridized molecules this is especially difficult since they may easily re-orient on the surface due to small differences in adsorption energy.

For example, Wang et al. reported the first single-molecule STM study of

[121]tetramantane on a Au(111) surface,41 finding that individual molecules oriented very differently on the surface upon tip manipulation. The authors inferred the orientation of hydrocarbon cages from comparison of density functional theory (DFT) computations with STM data. However, this approach is difficult since STM data depend on the local density of states, which is related to the molecular structure in a non-trivial way. Hence, the molecular structure is not measured directly. In addition, the Crommie group developed an STM–based vibrational spectroscopy technique (IRSTM) capable of providing spectroscopic evidence for the adsorbed assemblies consisting of adamantane and tetramantane sub-monolayers on a Au(111) surface.42, 43 By probing IR responses of such diamondoid assemblies they could not only differentiate between chemical structures but also conclude that intermolecular interactions between tetramantane molecules act more strongly than the interactions between tetramantanes and the gold surface, as evidenced by different tetramantane vibrational 4 ACS Paragon Plus Environment

Page 5 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

resonances.42 Despite these recent advances, direct determination of orientation by visual inspection of single, completely sp3-hybridized molecules remained elusive. Here we should stress the inherent difference between sp2 and sp3 (hydro)carbons, which is especially obvious when considering two-dimensional materials. An illustrative example is a comparison of graphene (sp2 material) with diamond (sp3 material). In case of graphene each carbon atom covalently binds with three neighboring carbons, forming three σ bonds and a conducting π band from the fourth valence electron and ultimately producing a flat sheet. 44 On the other hand, a diamond carbon atom forms four covalent bonds with neighboring carbons in a tetrahedral fashion and has no free electrons.45 As a consequence, graphene and diamond have different physico-chemical properties, but they also differ in their structural features, i.e., sp2 vs. sp3 structural arrangement. Coming back to analogous hydrocarbons (aromatic vs. diamondoid molecules), it is no surprise that their structure and, consequently, their possible on-surface orientation when forming monolayers on metal surfaces differs as well. With the aim to broaden the understanding of sp3-hydrocarbon self-assembly on metal surfaces and to quantify LD contributions responsible for the formation of highly ordered monolayers, our study combines experimental and computational approaches. The AFM technique with a CO-terminated tip enabled us to observe individual [121]tetramantane molecules with atomic resolution and to study the LD-driven assembly by directly determining their precise position and orientation. Using computational tools we could not only confirm the experimental findings, but also quantify the attractive intermolecular interactions. Such a systematic approach was necessary to gain insight into the role that LD plays in on-surface assembly of bulky hydrocarbons and molecules with dispersion energy donors (DEDs).

5 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 27

Results and Discussion. We deposited [121]tetramantane on two different metal surfaces, namely Au(111) and Cu(111). On both materials [121]tetramantane forms well-ordered large scale islands at room temperature (Fig. 2a and Fig. S1 in the Supplementary information, SI). The molecular islands can be fitted with regular 2D lattices over large distances on the order of several tens of nm (blue circles in Fig. 2a). The observed lattice parameters for the two metal surfaces are similar but not identical (Fig. 2). This leads to packing densities of 1.53 and 1.47 molecules nm–2 on Au(111) and Cu(111), respectively.

Figure 2. Self-assembly of [121]tetramantane on an Au(111) surface. (a) STM overview image of large scale 2D island and corresponding single line profile (solid and dashed red lines). The color scale is split into two parts to reveal the herringbone reconstruction of the Au(111) surface (left part) and the molecular island (right part) side by side. White arrows indicate lattice directions of the Au(111) surface. Blue circles represent a regular 2D lattice. Identical structures have been observed on different islands and different surfaces (average lattice parameters: Au(111) - vector a = (10.1 ± 0.1) Å, vector b = (7.6 ± 0.1) Å, angle = (58 ± 1)°, Cu(111) - a = (10.2 ± 0.1) Å, b = (7.7 ± 0.1) Å, α = (60 ± 1)°, Fig. S1). (b) Highresolution STM topography image. Zoom-in location is marked in (a) by a blue box. (c) High-resolution AFM frequency shift image revealing submolecular resolution. Parameters: 6 ACS Paragon Plus Environment

Page 7 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

STM: 200 mV, 10 pA; AFM: Δz = ‒236 pm (with respect to tunneling gap at 200 mV, 10 pA), Q = 8500, oscillation amplitude = 120 pm. All measurements performed at 5 K in UHV with CO-functionalized tip.

It should be noted that the 2D [121]tetramantane island (blue circles) does not align to any of the Au(111) surface vectors (white arrows in Fig. 2a). A closer look reveals that the obtained 2D lattice parameters for [121]tetramantane on Au(111) do not coincide with any regular overlayer structures. In other words, the observed molecular overlayer is incommensurate with the underlying Au lattice. Furthermore, we observe a virtually unaffected herringbone reconstruction of the Au(111) surface beneath the molecular layer (bright lines that shine through the layer in Fig. 2a and Fig. S2). Since any significant molecule-substrate interaction would inevitably lead to a stress relaxation between the Au atoms in the surface layer,46, 47 the interaction between [121]tetramantane molecules and the Au surface atoms must be very small. On Cu(111) the observed 2D overlayer is in accordance with a (3 × 4) superstructure, and hence aligns perfectly with the Cu lattice. At first glance this would indicate relatively strong molecule-substrate interactions.47-49

However, the observed 2D lattice parameters and

molecular packing densities on Au(111) and Cu(111) are very similar, and on both surfaces we observe identical types of intermolecular arrangement (as discussed below, Fig. 4). Hence, we rationalize that the presence of the (3 × 4) superstructure on Cu(111) is caused by a random coincidence of the molecular dimensions with the dimensions of the Cu(111) lattice rather than by strong molecule-substrate interactions. Furthermore, there are remarkable similarities between the lattice parameters of adsorbed 2D [121]tetramantane islands and bulk [121]tetramantane crystals.50 The vector c of the 3D monoclinic lattice structure has, e.g., a length of 10.33 Å, which is close to the determined 2D 7 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 27

vectors of a = (10.1 ± 0.1) Å and a = (10.2 ± 0.1) Å for Au(111) and Cu(111), respectively. Moreover, within (110) planes of the monoclinic 3D lattice the tetramantane molecules are packed with a 2D density of 1.46 molecules nm–2. This almost exactly matches the observed 2D packing density on Cu(111) and deviates by less than 5% for Au(111), which also indicates relatively weak molecule-substrate interactions. To reveal the precise adsorption structure and orientation of [121]tetramantane on Au(111) we performed high-resolution scans with CO-functionalized tips at 5 K. In Fig. 2b and 2c STM and AFM scans of the region marked with the blue box in Fig. 2a are shown. Submolecular resolution is obtained in the AFM image, with each depicted molecule being composed of several bright spots that appear in a specific pattern. It is remarkable that all molecules in the imaged region (around 80) are oriented in the same way except for a single molecule located at the rim of the island (red dashed circles). Next we determined the precise adsorption structure of the tetramantane molecules. In the beginning the bond imaging technique has been mainly applied to aromatic molecules, which adsorb planar to the surface.5,

13, 51, 52

For these systems the method directly allows for a

detailed interpretation of the chemical structure of the studied molecules. For 3D objects, however, image interpretation is not as straightforward. Since the standard bond imaging technique relies on 2D constant height scanning, the method is very sensitive to any height difference in the sample topography. Hence, to reveal structures of bulky molecules or angled adsorption structures, more sophisticated imaging schemes have to be utilized.19, 53-56

8 ACS Paragon Plus Environment

Page 9 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 3. Adsorption structure of [121]tetramantane. (a)-(c) AFM images of a small cluster of [121]tetramantane on Au(111) obtained at three different imaging heights, i.e., Δz = 360, 330, and 270 pm (with respect to tunneling gap at 200 mV, 150 pA). Parameters: Q = 21300, oscillation amplitude = 70 pm. Each molecule is composed of five bright spots, which are arranged in an Olympic ring-like pattern (orange circles). (d)-(g) Top and side views of [121]tetramantane adsorbed on Au(111). Hydrogens in the imaging and surface planes (gray planes) are marked as orange and blue spheres, respectively. Red spheres indicate two

9 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 27

hydrogens just below the imaging plane (and corresponding ones just above the surface plane), which determine the on-surface handedness of the adsorbed molecule. All other hydrogens are displayed as bright blue sticks.

Here, we performed constant height imaging at several different tip-sample distances to evaluate the formation of the image contrast with decreasing tip-sample distance and to obtain a 3D frequency shift map,57 which can be used to determine the tilting angle of the molecule. Fig. 3a-c shows three AFM images, which have been captured at tip-sample distances of Δz = 360, 330, and 270 pm (with respect to tunneling gap at 200 mV, 150 pA). The closer the COtip is approached to the surface, the more bright features gradually appear in the images. Ultimately, each molecule is represented by five bright spots, which are arranged in an Olympic ring-like pattern (orange circles in Fig. 3a-c). These bright spots are caused by repulsive interactions between hydrogen atoms in the imaging plane and the CO-tip, which lead to positive frequency shifts of the tuning fork sensor. Note that bright lines between the five bright spots, which appear when the tip is brought very close to the molecule (Fig. 3c), arise from a tilting of the CO-tip.58 Hence, these lines should not be misinterpreted as “bonds” between hydrogen atoms. The adsorption structure of [121]tetramantane on Au(111) is shown in Fig. 3d-g. Hydrogen atoms, which correspond to the imaged five-atom pattern, are displayed by orange spheres. Due to the symmetry of the molecule, a counterpart of the five-atom pattern is adsorbed to the Au(111) surface (blue spheres). The AFM images reveal differences in brightness of the five features, which are caused by a slight tilting of the molecule. We determined the tilting angle from the 3D frequency shift map (see Figs. S3 and S4 for details). The resulting tilting angles are approx. 5° and 12° with respect to the x- and y-axes, respectively (Fig. 3f, g). On Cu(111) we find a similar behavior of the image contrast with decreasing imaging heights. Hence, we 10 ACS Paragon Plus Environment

Page 11 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

conclude that the molecular tilting is rather independent of the surface material and, e.g., caused by the non-symmetric arrangement of atoms within the adsorption plane (and neighboring atoms just above this plane). Note that [121]tetramantane is, in our case, a prochiral molecule when attached to a surface, i.e., while being achiral in the gas phase, the molecules adopt an on-surface chirality by adsorbing on the surface plane.59 Therefore, an identification of the five-atom pattern is not sufficient for an absolute determination of molecular orientation.

A closer look at the

structure reveals that each [121]tetramantane molecule contains four of the mentioned fiveatom patterns, which can be obtained by rotating the molecule around its longitudinal axis (cf. x-axis in Fig. 3d-g). These patterns can be divided into two different types, which determine the on-surface chirality of the adsorbed molecules (in the following denoted as “Msurf” and “Psurf”-type). To distinguish between the types, two specific hydrogen atoms located just below the imaging plane are depicted as red spheres. These red atoms can be located either on the left or on the right side of the five-atom pattern (cf. orange and blue pattern in Fig. 3e, which are Msurf- and Psurf-type, respectively). The two red atoms are not directly revealed in our constant height scans. However, when the CO-tip is brought very close to the molecule during imaging, it can even sense the two red hydrogen atoms just below the imaging plane. In Fig. 3c these red atoms appear as a bright halo next to the highest atom within the five-atom pattern, as indicated by red arrows. This feature directly allows for determination of the surface-induced handedness of the molecules as the brightest atom within the five-atom pattern always indicates its on-surface handedness. Corresponding overlays of Msurf- and Psurf-type molecular models are depicted in Fig. 3b. After equilibration of the system for sufficiently long times at adequate temperatures (e.g., several minutes at room temperature), we observed that [121]tetramantane forms large scale well-ordered islands where nearly 100% of the molecules were perfectly arranged with each 11 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 27

other (Fig. 2c). However, even for smaller islands, which form at temperatures below 100 K within several seconds, there is a strong tendency for the molecules to cluster in the energetically most favorable way. Even within islands of 20–30 molecules, clusters on the order of 8–10 molecules with nearly perfect arrangement can be found (Fig. S5). Figs. 4a and 4b show the observed arrangement of molecules within the ordered clusters. The lattice parameters, i.e., lattice vectors and angles, can be precisely determined by large scale STM overview scans (Fig. 2). The exact orientation of single molecules within the 2D lattice is, however, only accessible using the bond imaging method by identifying the orientation of the five-atom pattern and the on-surface handedness of the molecules (which is obvious from the tilting of the molecule as discussed above, Fig. 3).

12 ACS Paragon Plus Environment

Page 13 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 4. Different types and on-surface chirality of tetramantane islands. (a), (c) Two different types (A and B) of structures are considered, which are composed of Msurf- and Psurftype molecules. (b), (d) For each type (A and B) one mirror-symmetric structure with identical binding energy exists. The observed lattice vectors and the corresponding angles for each surface material are given in caption of Fig. 2. While lattice parameters are identical, the orientational arrangement of the molecules is different for type A and B as indicated by 13 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 27

dashed lines between specific neighboring atoms of the molecules. (e) Type A (Psurf) island on Cu(111) surface. Parameters: Q = 20100, oscillation amplitude = 160 pm. (f) Two type A (Psurf) islands on Au(111) surface separated by grain boundary (dashed white line). (g) Type A (Msurf) island on Au(111) (same island as in Fig. 2). Parameters for (f,g): Q = 8500, oscillation amplitude = 120 pm.

In order to quantify LD contributions and support our experimental findings, we undertook a detailed computational study of LD-bound complexes initially consisting of two [121]tetramantanes.

After testing a number of geometries that arise from different

orientations of the tetramantane cage and by using several computational methods to verify our results (see SI for details), we found that the energetically most favorable tetramantane orientations were those that maximize the number of close contacts between the hydrocarbons. The computed energetic stabilization attributed to LD thereby amounts to about ‒10 kcal mol‒1. Another important finding was that functionals uncorrected for LD lead to very weekly bound or even dissociated structures and only dispersion-corrected computational methods give quantifiable LD-bound complexes, confirming the crucial role that LD plays in intermolecular interaction between the cages. These results further validate our rationale that LD is chiefly responsible for the experimentally observed 2D lattice selfassembly process. Additionally, we undertook a DFT study of four interacting [121]tetramantane molecules (see SI for details). We considered the case where only the orientation of the five-atom pattern with respect to the 2D lattice is known but information about the on-surface handedness of the molecules is missing. This is illustrated in Fig. 4a and 4c, which display two types of structures (A and B) that exhibit identical lattice parameters and packing densities but are composed of Msurf- and Psurf-type molecules, respectively. Note that for each type of structure 14 ACS Paragon Plus Environment

Page 15 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

a mirror symmetric counterpart with identical binding energy exists (cf. Fig. 4b, d). However, the binding energies of A- and B-type structures are not equal. This is revealed by the red dashed lines in Fig. 4a-d. While the molecules within the rows of both types of structures are organized in the same way, the arrangement between top and bottom rows significantly differs from each other. More detailed information about type A and B structures, including distances for the closest intermolecular contacts is given in the Supporting Information (Figures S6 and S7). The difference in binding energy of the two structural types is directly reflected by our computations.

While tetramers of type A are certainly stabilized by

dispersion interactions (by ~ 20 kcal mol‒1 = 0.87 eV), tetramers of type B are dominated by short range repulsive forces and are therefore energetically unfavorable. This finding is in agreement with our experimental observations since no structures of type B have been found on any surface material. A-type structures (of Msurf- and Psurf-type), however, have been observed on Au(111) and Cu(111) surfaces (see exemplary images in Fig. 4e-g). In addition, we studied the influence of a small tilting angle of the molecules with respect to the x-axis that has been observed in the experiments. As shown in Fig. 3f, the side of the fiveatom pattern connected to the two red hydrogen atoms is tilted upwards (+5°).

In the

computations we also studied the case when this side is tilted downwards (–5°, see Fig. S8 for details), which has a noticeable influence on the binding energy (Table 1). In particular, for structures of type B, the opposite tilting of –5° would even lead to a stabilization of the tetramer. Nevertheless, type B (–5°) tetramers are far less stable than all type A structures (for both, +5° and –5°). These results are remarkable in the context that all computations were conducted without taking the metal surface into account. Certainly, the substrate has some influence on island formation, which is, e.g., reflected by the slightly different lattice parameters and packing densities that have been observed on Au(111) and Cu(111) (Fig. 2). Nevertheless, for both 15 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 27

type A structures (on Au and Cu) the computed energy gain caused by intermolecular dispersion interactions is on the same order of magnitude (~ 20 kcal mol‒1). This energy gain clearly arises from the precise intermolecular arrangement as obvious from the differences between type A and B structures (and tilting angles). Therefore, we can follow that even small structural changes can drastically influence the binding energy and, furthermore, that the selfassembly process is driven by LD interactions.

Table 1. Computed interaction energies of type A and type B [121]tetramantane tetramers with different tilting angles.a Type

ΔE / kcal mol‒1

Type

ΔE / kcal mol‒1

Au(111)

A (+5°)

‒19.3

A (‒5°)

‒19.1

Au(111)

B (+5°)

73.6

B (‒5°)

‒4.4

Cu(111)

A (+5°)

‒20.6

A (‒5°)

‒19.5

Cu(111)

B (+5°)

33.7

B (‒5°)

‒7.0

a

Interaction energies are defined as a difference between the energy of the tetramer and the energy of four

[121]tetramantane molecules. Computations were performed at the B3LYP-D3(BJ)/6-31G(d,p) level of theory (see SI for details).

Conclusions. To summarize, we used STM and AFM imaging to visualize the sub-molecular structure of [121]tetramantane molecules and to explore their self-assembly on Cu(111) and Au(111) surfaces. Small diamondoids serve as an archetype system to study the role of LD interactions during structure formation since other binding contributions, such as hydrogen bonding, dipole-dipole, π-π, electrostatic interactions, or metal coordination, which predominate self-assembly processes in many other cases, can be neglected. The application of the bond imaging technique allowed a precise and unambiguous determination of the

16 ACS Paragon Plus Environment

Page 17 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

adsorption structure of bulky, completely sp3 hybridized molecules. It also allowed for a means to investigate the dependence of the organization process on their 2D arrangement. A detailed computational study supports the AFM findings and we successfully quantified the effect LD has on intermolecular interactions between tetramantanes, thus establishing LD as a driving force for directed self-assembly for such class of compounds. Although many highresolution AFM studies of surface phenomena have been reported, our investigation focuses on an entirely non-aromatic system and contributes to explorations of sp3 hybridized molecules capable of self-assembly on metal surfaces.

Methods. Sample preparation. [121]Tetramantane was isolated from petroleum and purified by multiple HPLC separations as described previously.36 Cu(111) and Au(111) crystals (Mateck, Germany) were cleaned by multiple cycles of sputtering with argon ions and subsequent annealing. To achieve small molecular clusters of 10–30 molecules the sample was precooled in the STM at 5 K. For evaporation the sample was removed from the STM with a precooled wobble stick and held for 30–60 s in front of a home-built sublimation device,47 which was kept at approx. 30 °C.

Large scale islands were obtained after

equilibrating the sample for several hours at room temperature. STM/AFM measurements. All measurements were performed at 5 K with a commercial ultrahigh vacuum, low temperature STM/AFM (ScientaOmicron, Germany) using tuning fork sensors and CO-functionalized tips.

The electrochemically etched tungsten tips were

conditioned by voltage pulses and indentations into the sample surface, prior to the functionalization with single CO molecules using standard recipes.60 All images have been obtained by constant height scanning using the frequency modulation mode. Small drifts of the scanner in z-direction during scanning were regularly adjusted by pausing the scan and switching on the STM feedback. 17 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 27

Computational details. All geometry optimizations were performed with the GAUSSIAN09 program package using several DFT approaches (see SI for details) and the obtained minima were verified by frequency computations. Since the B3LYP functional does not account for LD, it needs to be augmented by a correction such as Grimme’s D3-dispersion formulation in connection with Becke-Johnson (BJ) damping. This approach allows for an analysis of the LD contribution in the studied [121]tetramantane complexes. Single-point energies on the DFT-optimized structures (B3LYP-D3(BJ)/6-31G(d,p) level of theory) were computed with the ORCA 4.0.0 program using RI-MP2 and DLPNO-CCSD(T) methods and complete basis set extrapolation was also performed. All computation methods we used (except for the uncorrected B3LYP which does not account for LD and therefore cannot provide quantifiable dimers) showed comparable trends in interaction energies for the LD-bound complexes (Table S1). We found that the energetically most stable complexes engage in greatest number of close contacts between tetramantanes, confirming that LD significantly acts between the cages and is responsible for preferable orientation of molecules.

Another validity

confirmation of our computational approach comes from Local Energy Decomposition (LED) analysis which is implemented in ORCA 4.0.0 for the DLPNO-CCSD(T) method. Dispersion contributions from decomposition of CCSD pair energies were in agreement with energies obtained by other methods we used (Table S1).

Complete computational details and

references are provided in the SI.

ASSOCIATED CONTENT Supporting Information.

Additional STM/AFM images, computational details and

Cartesian coordinates are available free of charge via the Internet at http://pubs.acs.org. The authors declare no competing financial interest. 18 ACS Paragon Plus Environment

Page 19 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected]; [email protected]; [email protected]. ORCID Daniel Ebeling: 0000-0001-5829-170X Marina Šekutor: 0000-0003-1629-3672 Peter R. Schreiner: 0000-0002-3608-5515 Author Contributions ‡ D.E.

and M.Š. contributed equally to the work.

ACKNOWLEDGMENT We acknowledge financial support from the Deutsche Forschungsgemeinschaft (DFG) via the GRK (Research Training Group) 2204 "Substitute Materials for Sustainable Energy Technologies". This project was also supported by the Laboratory of Materials Research (LaMa) of JLU and the LOEWE program of excellence of the Federal State of Hessen (project initiative STORE-E). The work was also supported in part (for J.E.P.D., R.M.K.C., and P.R.S) by the US Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, under contract DE-AC02-76SF00515. M.Š. thanks the Alexander-von-Humboldt Foundation for a Humboldt Research Fellowship for Postdoctoral Researchers.

Computations were conducted on the LOEWE-CSC high-

performance computer of the State of Hesse.

19 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 27

REFERENCES 1.

Stepanow, S.; Lingenfelder, M.; Dmitriev, A.; Spillmann, H.; Delvigne, E.; Lin, N.;

Deng, X.; Cai, C.; Barth, J. V.; Kern, K., Steering Molecular Organization and Host-Guest Interactions Using Two-Dimensional Nanoporous Coordination Systems. Nat. Mater. 2004, 3, 229‒233. 2.

Grill, L.; Dyer, M.; Lafferentz, L.; Persson, M.; Peters, M. V.; Hecht, S., Nano-

Architectures by Covalent Assembly of Molecular Building Blocks. Nat. Nanotechnol. 2007, 2, 687‒691. 3.

Cai, J.; Ruffieux, P.; Jaafar, R.; Bieri, M.; Braun, T.; Blankenburg, S.; Muoth, M.;

Seitsonen, A. P.; Saleh, M.; Feng, X.; Mullen, K.; Fasel, R., Atomically Precise Bottom-up Fabrication of Graphene Nanoribbons. Nature 2010, 466, 470‒473. 4.

Sengupta, S.; Ebeling, D.; Patwardhan, S.; Zhang, X.; von Berlepsch, H.; Böttcher, C.;

Stepanenko, V.; Uemura, S.; Hentschel, C.; Fuchs, H.; Grozema, F. C.; Siebbeles, L. D. A.; Holzwarth, A. R.; Chi, L.; Würthner, F., Biosupramolecular Nanowires from Chlorophyll Dyes with Exceptional Charge-Transport Properties. Angew. Chem., Int. Ed. 2012, 51, 6378‒ 6382. 5.

Gross, L.; Mohn, F.; Moll, N.; Liljeroth, P.; Meyer, G., The Chemical Structure of a

Molecule Resolved by Atomic Force Microscopy. Science 2009, 325, 1110‒1114. 6.

Moll, N.; Gross, L.; Mohn, F.; Curioni, A.; Meyer, G., The Mechanisms Underlying

the Enhanced Resolution of Atomic Force Microscopy with Functionalized Tips. New J. Phys. 2010, 12, 125020. 7.

Altman, E. I.; Baykara, M. Z.; Schwarz, U. D., Noncontact Atomic Force Microscopy:

An Emerging Tool for Fundamental Catalysis Research. Acc. Chem. Res. 2015, 48, 2640‒ 2648. 20 ACS Paragon Plus Environment

Page 21 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

8.

ACS Nano

Eigler, D. M.; Schweizer, E. K., Positioning Single Atoms with a Scanning Tunnelling

Microscope. Nature 1990, 344, 524‒526. 9.

Meyer, G.; Neu, B.; Rieder, K.-H., Controlled Lateral Manipulation of Single

Molecules with the Scanning Tunneling Microscope. Appl. Phys. A 1995, 60, 343‒345. 10.

Morgenstern, K.; Lorente, N.; Rieder, K.-H., Controlled Manipulation of Single

Atoms and Small Molecules Using the Scanning Tunnelling Microscope. Phys. Status Solidi B 2013, 250, 1671‒1751. 11.

Aragonès, A. C.; Haworth, N. L.; Darwish, N.; Ciampi, S.; Bloomfield, N. J.; Wallace,

G. G.; Diez-Perez, I.; Coote, M. L., Electrostatic Catalysis of a Diels–Alder Reaction. Nature 2016, 531, 88-91. 12.

Jarvis, S. P., Resolving Intra- and Inter-Molecular Structure with Non-Contact Atomic

Force Microscopy. Int. J. Mol. Sci. 2015, 16, 19936-19959. 13.

Pavliček, N.; Gross, L., Generation, Manipulation and Characterization of Molecules

by Atomic Force Microscopy. Nat. Rev. Chem. 2017, 1, 0005. 14.

Schuler, B.; Meyer, G.; Gross, L.; Pena, D.; Mullins, O. C., Unraveling the Molecular

Structures of Asphaltenes by Atomic Force Microscopy. J. Am. Chem. Soc. 2015, 137, 9870‒ 9876. 15.

De Oteyza, D. G.; Gorman, P.; Chen, Y.-C.; Wickenburg, S.; Riss, A.; Mowbray, D.

J.; Etkin, G.; Pedramrazi, Z.; Tsai, H.-Z.; Rubio, A.; Crommie, M. F.; Fischer, F. R., Direct Imaging of Covalent Bond Structure in Single-Molecule Chemical Reactions. Science 2013, 340, 1434‒1437. 16.

Pavliček, N.; Schuler, B.; Collazos, S.; Moll, N.; Perez, D.; Guitian, E.; Meyer, G.;

Pena, D.; Gross, L., On-Surface Generation and Imaging of Arynes by Atomic Force Microscopy. Nat. Chem. 2015, 7, 623‒628.

21 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

17.

Page 22 of 27

Riss, A.; Paz, A. P.; Wickenburg, S.; Tsai, H.-Z.; De Oteyza, D. G.; Bradley, A. J.;

Ugeda, M. M.; Gorman, P.; Jung, H. S.; Crommie, M. F.; Rubio, A.; Fischer, F. R., Imaging Single-Molecule Reaction Intermediates Stabilized by Surface Dissipation and Entropy. Nat. Chem. 2016, 8, 678‒683. 18.

Schuler, B.; Fatayer, S.; Mohn, F.; Moll, N.; Pavliček, N.; Meyer, G.; Pena, D.; Gross,

L., Reversible Bergman Cyclization by Atomic Manipulation. Nat. Chem. 2016, 8, 220‒224. 19.

Zint, S.; Ebeling, D.; Schlöder, T.; Ahles, S.; Mollenhauer, D.; Wegner, H. A.;

Schirmeisen, A., Imaging Successive Intermediate States of the On-Surface Ullmann Reaction on Cu(111): Role of the Metal Coordination. ACS Nano 2017, 11, 4183‒4190. 20.

London, F., The General Theory of Molecular Forces. Trans. Faraday Soc. 1937, 33,

8‒26. 21.

Wagner, J. P.; Schreiner, P. R., London Dispersion in Molecular Chemistry-

Reconsidering Steric Effects. Angew. Chem., Int. Ed. 2015, 54, 12274‒12296. 22.

Jiemchooroj, A.; Sernelius, B. E.; Norman, P., C6 Dipole-Dipole Dispersion

Coefficients for the n-Alkanes: Test of an Additivity Procedure. Phys. Rev. A 2004, 69, 044701. 23.

Israelachvili, J. N., In Intermolecular and Surface Forces (Third Edition), Academic

Press: San Diego, 2011; p 107‒132. 24.

Grimme, S.; Schreiner, P. R., Steric Crowding Can Stabilize a Labile Molecule:

Solving the Hexaphenylethane Riddle. Angew. Chem., Int. Ed. 2011, 50, 12639‒12642. 25.

Fokin, A. A.; Chernish, L. V.; Gunchenko, P. A.; Tikhonchuk, E. Y.; Hausmann, H.;

Serafin, M.; Dahl, J. E. P.; Carlson, R. M. K.; Schreiner, P. R., Stable Alkanes Containing Very Long Carbon–Carbon Bonds. J. Am. Chem. Soc. 2012, 134, 13641‒13650. 26.

De Feyter, S.; De Schryver, F. C., Self-Assembly at the Liquid/Solid Interface:  STM

Reveals. J. Phys. Chem. B 2005, 109, 4290‒4302.

22 ACS Paragon Plus Environment

Page 23 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

27.

ACS Nano

Kepčija, N.; Zhang, Y.-Q.; Kleinschrodt, M.; Björk, J.; Klyatskaya, S.; Klappenberger,

F.; Ruben, M.; Barth, J. V., Steering On-Surface Self-Assembly of High-Quality Hydrocarbon Networks with Terminal Alkynes. J. Phys. Chem. C 2013, 117, 3987‒3995. 28.

Ascolani, H.; van der Meijden, M. W.; Cristina, L. J.; Gayone, J. E.; Kellogg, R. M.;

Fuhr, J. D.; Lingenfelder, M., Van der Waals Interactions in the Self-Assembly of 5Amino[6]helicene on Cu(100) and Au(111). Chem. Commun. 2014, 50, 13907‒13909. 29.

Gao, H.-Y.; Wagner, H.; Held, P. A.; Du, S.; Gao, H.-J.; Studer, A.; Fuchs, H., In-

Plane Van der Waals Interactions of Molecular Self-Assembly Monolayer. Appl. Phys. Lett. 2015, 106, 081606. 30.

Echeverría, J.; Aullón, G.; Danovich, D.; Shaik, S.; Alvarez, S., Dihydrogen Contacts

in Alkanes Are Subtle but Not Faint. Nat. Chem. 2011, 3, 323‒330. 31.

Danovich, D.; Shaik, S.; Neese, F.; Echeverría, J.; Aullón, G.; Alvarez, S.,

Understanding the Nature of the CH···HC Interactions in Alkanes. J. Chem. Theory Comput. 2013, 9, 1977‒1991. 32.

Wang, C.; Mo, Y.; Wagner, J. P.; Schreiner, P. R.; Jemmis, E. D.; Danovich, D.;

Shaik, S., The Self-Association of Graphane Is Driven by London Dispersion and Enhanced Orbital Interactions. J. Chem. Theory Comput. 2015, 11, 1621‒1630. 33.

Schwertfeger, H.; Fokin, A. A.; Schreiner, P. R., Diamonds Are a Chemist's Best

Friend: Diamondoid Chemistry Beyond Adamantane. Angew. Chem., Int. Ed. 2008, 47, 1022‒ 1036. 34.

Gunawan, M. A.; Hierso, J.-C.; Poinsot, D.; Fokin, A. A.; Fokina, N. A.; Tkachenko,

B. A.; Schreiner, P. R., Diamondoids: Functionalization and Subsequent Applications of Perfectly Defined Molecular Cage Hydrocarbons. New J. Chem. 2014, 38, 28‒41.

23 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

35.

Page 24 of 27

Burns, W.; McKervey, M. A.; Mitchell, T. R. B.; Rooney, J. J., A New Approach to

the Construction of Diamondoid Hydrocarbons. Synthesis of anti-Tetramantane. J. Am. Chem. Soc. 1978, 100, 906‒911. 36.

Dahl, J. E.; Liu, S. G.; Carlson, R. M. K., Isolation and Structure of Higher

Diamondoids, Nanometer-Sized Diamond Molecules. Science 2003, 299, 96‒99. 37.

Schreiner, P. R.; Fokina, N. A.; Tkachenko, B. A.; Hausmann, H.; Serafin, M.; Dahl,

J. E. P.; Liu, S.; Carlson, R. M. K.; Fokin, A. A., Functionalized Nanodiamonds: Triamantane and [121]Tetramantane. J. Org. Chem. 2006, 71, 6709‒6720. 38.

Yang, W. L.; Fabbri, J. D.; Willey, T. M.; Lee, J. R. I.; Dahl, J. E.; Carlson, R. M. K.;

Schreiner, P. R.; Fokin, A. A.; Tkachenko, B. A.; Fokina, N. A.; Meevasana, W.; Mannella, N.; Tanaka, K.; Zhou, X. J.; van Buuren, T.; Kelly, M. A.; Hussain, Z.; Melosh, N. A.; Shen, Z.-X., Monochromatic Electron Photoemission from Diamondoid Monolayers. Science 2007, 316, 1460‒1462. 39.

Clay, W. A.; Liu, Z.; Yang, W.; Fabbri, J. D.; Dahl, J. E.; Carlson, R. M. K.; Sun, Y.;

Schreiner, P. R.; Fokin, A. A.; Tkachenko, B. A.; Fokina, N. A.; Pianetta, P. A.; Melosh, N.; Shen, Z.-X., Origin of the Monochromatic Photoemission Peak in Diamondoid Monolayers. Nano Lett. 2009, 9, 57‒61. 40.

Narasimha, K. T.; Ge, C.; Fabbri, J. D.; Clay, W.; Tkachenko, B. A.; Fokin, A. A.;

Schreiner, P. R.; Dahl, J. E.; Carlson, R. M. K.; Shen, Z. X.; Melosh, N. A., Ultralow Effective Work Function Surfaces Using Diamondoid Monolayers. Nat. Nanotechnol. 2016, 11, 267‒272. 41.

Wang, Y.; Kioupakis, E.; Lu, X.; Wegner, D.; Yamachika, R.; Dahl, J. E.; Carlson, R.

M. K.; Louie, S. G.; Crommie, M. F., Spatially Resolved Electronic and Vibronic Properties of Single Diamondoid Molecules. Nat. Mater. 2008, 7, 38‒42.

24 ACS Paragon Plus Environment

Page 25 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

42.

ACS Nano

Pechenezhskiy, I. V.; Hong, X.; Nguyen, G. D.; Dahl, J. E. P.; Carlson, R. M. K.;

Wang, F.; Crommie, M. F., Infrared Spectroscopy of Molecular Submonolayers on Surfaces by Infrared Scanning Tunneling Microscopy: Tetramantane on Au(111). Phys. Rev. Lett. 2013, 111, 126101. 43.

Sakai, Y.; Nguyen, G. D.; Capaz, R. B.; Coh, S.; Pechenezhskiy, I. V.; Hong, X.;

Wang, F.; Crommie, M. F.; Saito, S.; Louie, S. G.; Cohen, M. L., Intermolecular Interactions and Substrate Effects for an Adamantane Monolayer on a Au(111) Surface. Phys. Rev. B 2013, 88, 235407. 44.

Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K., The

Electronic Properties of Graphene. Rev. Mod. Phys. 2009, 81, 109‒162. 45.

Rafi, K., Diamond as a Unique High-Tech Electronic Material: Difficulties and

Prospects. J. Phys. D: Appl. Phys. 2007, 40, 6467. 46.

Gimzewski, J. K.; Modesti, S.; Gerber, C.; Schlittler, R. R., Observation of a New

Au(111) Reconstruction at the Interface of an Adsorbed C60 Overlayer. Chem. Phys. Lett. 1993, 213, 401‒406. 47.

Zint, S.; Ebeling, D.; Ahles, S.; Wegner, H. A.; Schirmeisen, A., Subsurface-

Controlled Angular Rotation: Triphenylene Molecules on Au(111) Substrates. J. Phys. Chem. C 2016, 120, 1615‒1622. 48.

Weckesser, J.; Cepek, C.; Fasel, R.; Barth, J. V.; Baumberger, F.; Greber, T.; Kern, K.,

Binding and Ordering of C60 on Pd(110): Investigations at the Local and Mesoscopic Scale. J. Chem. Phys. 2001, 115, 9001‒9009. 49.

Ebeling, D.; Bradler, S.; Roling, B.; Schirmeisen, A., 3-Dimensional Structure of a

Prototypical Ionic Liquid–Solid Interface: Ionic Crystal-Like Behavior Induced by Molecule– Substrate Interactions. J. Phys. Chem. C 2016, 120, 11947‒11955.

25 ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

50.

Page 26 of 27

Roberts, P. J.; Ferguson, G., anti-Tetramantane, a Large Diamondoid Fragment. Acta

Crystallogr., Sect. B 1977, 33, 2335‒2337. 51.

Gross, L.; Mohn, F.; Moll, N.; Schuler, B.; Criado, A.; Guitián, E.; Peña, D.; Gourdon,

A.; Meyer, G., Bond-Order Discrimination by Atomic Force Microscopy. Science 2012, 337, 1326‒1329. 52.

Zhang, J.; Chen, P.; Yuan, B.; Ji, W.; Cheng, Z.; Qiu, X., Real-Space Identification of

Intermolecular Bonding with Atomic Force Microscopy. Science 2013, 342, 611‒614. 53.

Albrecht, F.; Pavliček, N.; Herranz-Lancho, C.; Ruben, M.; Repp, J., Characterization

of a Surface Reaction by Means of Atomic Force Microscopy. J. Am. Chem. Soc. 2015, 137, 7424‒7428. 54.

Moreno, C.; Stetsovych, O.; Shimizu, T. K.; Custance, O., Imaging Three-

Dimensional Surface Objects with Submolecular Resolution by Atomic Force Microscopy. Nano Lett. 2015, 15, 2257‒2262. 55.

Majzik, Z.; Cuenca, A. B.; Pavliček, N.; Miralles, N.; Meyer, G.; Gross, L.;

Fernández, E., Synthesis of a Naphthodiazaborinine and Its Verification by Planarization with Atomic Force Microscopy. ACS Nano 2016, 10, 5340‒5345. 56.

Ebeling, D.; Zhong, Q.; Ahles, S.; Chi, L.; Wegner, H. A.; Schirmeisen, A., Chemical

Bond Imaging Using Higher Eigenmodes of Tuning Fork Sensors in Atomic Force Microscopy. Appl. Phys. Lett. 2017, 110, 183102. 57.

Hölscher, H.; Langkat, S. M.; Schwarz, A.; Wiesendanger, R., Measurement of Three-

Dimensional Force Fields with Atomic Resolution Using Dynamic Force Spectroscopy. Appl. Phys. Lett. 2002, 81, 4428‒4430. 58.

Hapala, P.; Kichin, G.; Wagner, C.; Tautz, F. S.; Temirov, R.; Jelínek, P., Mechanism

of High-Resolution STM/AFM Imaging with Functionalized Tips. Phys. Rev. B 2014, 90, 085421.

26 ACS Paragon Plus Environment

Page 27 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

59.

ACS Nano

Stetsovych, O.; Švec, M.; Vacek, J.; Chocholoušová, J. V.; Jančařík, A.; Rybáček, J.;

Kosmider, K.; Stará, I. G.; Jelínek, P.; Starý, I., From Helical to Planar Chirality by OnSurface Chemistry. Nat. Chem. 2017, 9, 213‒218. 60.

Bartels, L.; Meyer, G.; Rieder, K. H.; Velic, D.; Knoesel, E.; Hotzel, A.; Wolf, M.;

Ertl, G., Dynamics of Electron-Induced Manipulation of Individual CO Molecules on Cu(111). Phys. Rev. Lett. 1998, 80, 2004‒2007.

27 ACS Paragon Plus Environment