Structural Evolutions of the Self-Assembled N-decyldecanamide on Au

hexangular porous networks (we refer to as type I), and the unit cell parameters .... transform from phase I to phase III or phase IV when the molecul...
0 downloads 0 Views 566KB Size
Subscriber access provided by University of South Dakota

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Structural Evolutions of the Self-Assembled N-decyldecanamide on Au(111) Chencheng Peng, Haiping Lin, Kaifeng Niu, Junjie Zhang, Jinqiang Ding, Biao Yang, Nan Cao, Haiming Zhang, Youyong Li, Qing Li, and Lifeng Chi J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b06517 • Publication Date (Web): 05 Sep 2018 Downloaded from http://pubs.acs.org on September 15, 2018

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 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 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.

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 23 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

The Journal of Physical Chemistry

Structural Evolutions of the Self-Assembled N-decyldecanamide on Au(111) Chencheng Peng, Haiping Lin, Kaifeng Niu, Junjie Zhang, Jinqiang Ding, Biao Yang, Nan Cao, Haiming Zhang, Youyong Li, Qing Li,* and Lifeng Chi*

Institute of Functional Nano & Soft Materials (FUNSOM) and Jiangsu Key Laboratory for Carbon-Based Functional Materials & Devices, Soochow University, Suzhou 215123, P. R. China

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

ABSTRACT: The self-assembly of N-decyldecanamide (DDA) on Au(111) is investigated by means of scanning tunneling microscopy (STM) under ultrahigh vacuum conditions. Four self-assembled phases have been observed by varying the molecular coverage and the evaporation temperature. At lower coverage, the self-assembled phase with lowest Gibbs free energy is preferred, suggesting the self-assembly at low coverage is thermodynamically driven. At fixed evaporation temperature, the increasing of the molecular coverage results in the phase transitions from sparse packing to dense ones. Based on the systematic study of the self-assembly structures of DDA on the Au(111) surfaces, a phase diagram is eventually obtained. This study provides means to control the phase transitions on Au(111) surfaces by altering the balance of the influence of thermodynamic and kinetic effects.

2

ACS Paragon Plus Environment

Page 2 of 23

Page 3 of 23 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

The Journal of Physical Chemistry

INTRODUCTION Molecular self-assembly on metal surfaces has been considered to have potential application in functional biosensors and nanomaterials.1-5 The nature of the molecular self-assembly on surfaces is determined by the balance of the molecule-molecule and molecule-substrate interactions.6-11 Generally speaking, molecules tend to aggregate with each other, such that the minimum Gibbs free energy can be achieved when the system is in equilibrium. For the cases in which molecule-molecule or molecule-substrate interaction is relatively strong, the kinetic effect needs to be taken into consideration since the surface mobility is suppressed.12-17 The competition of the thermodynamic and kinetic effects leads to diverse self-assembled structures on surfaces which can be tuned by various external parameters, such as the concentration of the solution,18-21 the substrate,22,23 solvent effect,24,25 procedure of the thermal treatment26-29 and so forth. Considering the importance and the complicity, controlling of the molecular self-assembly on surfaces has acquired extensive interest in recent decades.30-34 Here, we report the systematical study on the self-assembly phenomenon of N-decyldecanamide (DDA) molecules (the structural models with two configurations are shown in Scheme 1) on the Au(111) surfaces. The experiments are conducted under ultrahigh vacuum (UHV), so that the influence of the solvent is prevented. Aided by the scanning tunneling microscopy (STM) imaging, we obtain a series of self-assembly structures with single-molecule resolution. By carefully analyzing the structures prepared with different coverages and molecular evaporation temperatures, 3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

a phase diagram is drawn. Density functional theory (DFT) calculations show that the low coverage phase is thermodynamically favorable. At higher coverage, the increasing of the collision probability of the molecules leads to denser phases, despite the larger Gibbs free energy in those phases.

Scheme 1. The Ball-Stick Model of N-decyldecanamide (DDA).

METHODS STM Measurements. All STM measurements were performed in an ultrahigh vacuum (UHV) system (base pressure better than 1 × 10−10 Torr). Experiments were conducted with commercial Unisoku low temperature scanning tunneling microscopy (LT-STM 1500S). All the STM measurements were acquired at 77 K. The Au(111) substrates were cleaned by 3 repeated cycles of argon ion sputtering (1 keV) and annealing (450 ℃, 15 mins) before the deposition of organic molecules. The DDA molecules were synthesized by Gerhard Erker’s group in the University of Münster. DDA molecules were deposited by organic molecular beam epitaxy (OMBE, kentax UHV equipment company). A commercial Pt-Ir tip (purchased from UNISOKU company) was carefully prepared by the e-beam heating. All images were processed and analyzed by WSxM.35 4

ACS Paragon Plus Environment

Page 4 of 23

Page 5 of 23 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

The Journal of Physical Chemistry

Theoretical Calculations. All calculations in this work were carried out with the periodic plane-wave DFT code VASP 5.3.36,37 along with standard library ultrasoft pseudopotentials. The dispersion-corrected PBE-D3 functional38,39 was used to model these self-assembled systems. The PBE-D3 functional corresponds to the gradient-corrected exchange and correlation functional of Perdew-Burke-Ernzerhof (PBE)38 coupled with the third-generation post-SCF dispersion corrections proposed by Grimme.39,40 The Atomic Simulation Environment (ASE)41 was used to set up the simulation cells (vacuum>20 A). Each slab was constructed as a perfect (111)-fcc surface with PBE-optimized lattice constants of 4.16 A for bulk Au. Four metal layers, sufficient to describe the ABCABC stacking scheme and shown to be sufficient to describe the systems42,43 were employed in the calculations. The energy of a DDA molecule in different phases is defined as EDDA = (Eslab/DDA-Eslab)/N

(1)

where Eslab/DDA is the total energy of the combined system, Eslab is the total energy of the clean slab, and N is the number of DDA molecules in the unit cells. For geometry optimizations, we use an energy cutoff of 450 eV and a 1 × 1 × 1 Monkhorst-Pack grid. Final single-point calculations employed the same energy cutoff, but a more densely sampled (4 × 4 × 1) k-point grid, in order to resolve the adsorption energy with the desired accuracy.

RESULTS AND DISCUSSION 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

1. Coverage Dependent Self-Assembly. Figure 1 gives the STM topographic images after depositing DDA molecules on the Au(111) surfaces held at room temperature. Figure 1a-1c show the representative structure of the self-assembled islands with different molecular coverages when the DDA molecules are evaporated at the temperature of 40 ℃. At lower coverage (0.02 ± 0.01 monolayer (ML)), each DDA molecular component is observed to bend at the amide linkage, as depicted in Figure 1a. Similar bending of the long-chain hydrocarbon derivatives has been reported previously on various surfaces.44-46 The DDA molecules aggregate into hexangular porous networks (we refer to as type I), and the unit cell parameters are: a = 3.54 ± 0.02 nm, b = 3.54 ± 0.02 nm, θ = 60 ± 2°. Figure 1e provides the relaxed structural model, in which three DDA molecules form a trimer unit via the N-H…O hydrogen bonds (The substrate is involved in the DFT calculations. The computational details are described in the methods part). The trimers aggregate and form hexangular nanostructure via van der Waals interactions between alkyl chains of the adjacent molecules. The calculated pore-to-pore distance is 3.53 nm, which is in nice agreement with the experimental observations (3.54 ± 0.02 nm). Structural evolutions take place when the molecular coverage increases. Figure 1b gives a representative STM image of the sample with 0.1 ± 0.02 ML DDA molecules deposited on the Au(111) surface. A large-scale well-ordered rhomboid pore structure (we refer to as type II) is formed, in which each DDA molecule maintains its bent configuration. Two bent DDA monomers interact with each other in a head-to-head fashion via two N-H…O hydrogen bonds between the amide groups. The formed 6

ACS Paragon Plus Environment

Page 6 of 23

Page 7 of 23 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

The Journal of Physical Chemistry

dimers interact with the adjacent ones through van der Waals interactions between the alkyl chains. The unit cell of the rhomboid pore structure is given in Figure 1b. The size of the unit cell is a = 2.00 ± 0.02 nm, b = 1.70 ± 0.02 nm, and the angle between a and b directions is 90 ± 2°. Figure 1f shows the optimized structure model. The calculated size of the unit cell (a = 2.02 nm, b = 1.70 nm, θ = 90°) agrees well with the experimental observations.

Figure 1. Self-assembly structures after depositing different amount of DDA molecules on the Au(111) surfaces held at room temperature. (a)-(c) STM images after depositing DDA molecules on the Au(111) surfaces, with the evaporation temperature of 40 ℃. The molecular coverage is 0.02 ± 0.01 ML for (a), 0.1 ± 0.02 ML for (b) and 0.25 ± 0.05 ML for (c). (d) STM images after depositing 0.25 ± 0.05 ML DDA molecules on the Au(111) surface, with the evaporation temperature of 60 ℃. The scanning parameters are It = 500 pA, Vb = -1 V for (a), It = 20 pA, Vb = -1 V for (b), It = 400 pA, Vb = -1 V for (c) and It = 500 pA, Vb = -1 V for (d). (e)-(h) DFT optimized structural models of the self-assembly phases given in (a)-(d), respectively. The scale bar is 5 nm for (a), (b), (c) and 2nm for (d). 7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

A new phase emerges when the molecular coverage increases up to 0.25 ± 0.05 ML. In this phase, the DDA molecules exhibit two kinds of configurations: the bent configuration and the straight ones. The unit cell of the close-packed phase (we refer to as type III) is shown in Figure 1c. The unit cell has a size of (2.35 ± 0.02) × (2.03 ± 0.02) nm2, and an angle of 90 ± 2°. As shown in the DFT optimized structural model (Figure 1g), two DDA molecules with bent configurations interact with each other and form a dimer via van der Waals interactions between the alkyl chains. The dimers interact with the adjacent straight molecules via N-H…O hydrogen bonds between the amide groups. The calculated unit cell is 2.35 × 2.04 nm2, and the angle between a and b directions is 90°, in nice agreement with the STM image shown in Figure 1c. No phase transitions are observed by further increasing the molecular coverage if the evaporation temperature maintains at 40 ℃. A new self-assembly phase, however, emerges when the evaporation temperature is elevated to 60 ℃. The new phase exhibits a compact stripe self-assembly structure (we refer to as type IV), in which all the DDA components show their straight configurations. As shown in Figure 1d, the DDA molecules have nearly identical orientation and they interact with each other via both the N-H…O hydrogen bonds between the amide groups and van der Waals interactions between the alkyl chains. The unit cell is given in Figure 1d, which exhibits a quasi-square shape (a = 2.94 ± 0.01 nm, b = 0.50 ± 0.01 nm, θ = 85 ± 1°). The observation is in agreement with the previous studies.47 From the DFT calculations (Figure 1h), the parameter of the unit cell is a = 2.94 nm, b = 0.50 nm,

8

ACS Paragon Plus Environment

Page 8 of 23

Page 9 of 23 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

The Journal of Physical Chemistry

and the angle between a and b directions is 85°, agrees well with the STM measurements. 2. Phase Diagram of the Self-Assembly. We also get the STM images of the samples prepared with the molecular evaporation temperature of 60 ℃ and 80 ℃, respectively, as shown in Figure S1. In summary, four kinds of self-assembly structures of DDA molecules are observed under UHV on the Au(111) surfaces. By systematically monitoring the assembly phases by varying the molecular coverage and the evaporation temperature, a phase diagram is obtained. As shown in Figure 2, when the molecular coverage is relatively low (e.g. lower than 0.07 ML), a hexangular porous structure (Type I) is preferred. It has been reported that the influence of kinetic factors on the self-assembly is suppressed at lower coverage in various systems, both under vacuum and in solutions.14,39-42 As a consequence, the molecules tend to aggregate into a structure that meets the requirement for the lowest Gibbs free energy. In order to confirm that the self-assembly structure shown in Figure 1a is thermodynamically driven, DFT calculations were performed. The calculated total energies of each phase are summarized in Figure 3. As seen, the total energy of DDA molecules of the hexangular phase is -8159.40 kcal/mol, lower than that of the other three kinds of phases, confirming that the formation of the hexangular phase is thermodynamically most favorable. In order to estimate the contributions of the hydrogen bonds to the molecular interaction, we have performed additional DFT calculations with and without the dispersion correction of each assembly structures in gas phase. As described by the supporting information (Table S1), the contribution of 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

the hydrogen bonds in the self-assembly phases is in the range of ~23% to ~38%. The calculated hydrogen bond energy is in good agreement with previous reports (-6.46 kcal/mol).52,47 As seen in Figure 2, at a determined evaporation temperature, which means the evaporation rate of the molecules keeps unaltered, the self-assembly structures tend to transform from phase I to phase III or phase IV when the molecular coverage increases. Similar phase evolutions have been reported in solutions,18,21 and they attributed such concentration dependent phase evolutions to that the denser packing is preferred at higher coverage. Ideally, the critical concentration that leads to the phase transition in solution is that the loose self-assembled structure occupied the entire surface.21 Under UHV, similarly, both the increment of the molecular coverage and the evaporating rate results in the increasing of the collision probability of the molecules on surfaces. As a consequence, the self-assembled phase is not solely determined by the minimum of the free energy, leading to the formation of denser packed molecules on the surface. As shown in Table 1, the measured molecular density is 0.552 molecules/nm2 for phase I, 0.588 molecules/nm2 for phase II, 0.628 molecules/nm2 for phase III and 0.682 molecules/nm2 for phase IV. The sequence agrees nicely with the experimental observations that denser packing is observed at higher molecular coverage or evaporation temperature, despite that those phases possess higher free energies.

10

ACS Paragon Plus Environment

Page 10 of 23

Page 11 of 23 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

The Journal of Physical Chemistry

Figure 2. (a) Coverage and evaporation temperature dependence of the self-assembly structures of DDA molecules on the Au(111) surfaces. (b) The phase diagram extracted from (a).

Figure 3. The calculated total energies (kcal/mol) for the optimized self-assembled structures of the DDA molecules for the porous, rhomboid, close-packed and stripe phases, respectively.

Finally, we annealed the samples in order to decrease the molecular coverage via the considerable molecular desorption. Upon decreasing the molecular coverage, the phases with relatively high coverage (rhomboid, close-packed, stripe phases) undergo a thorough transition to the porous phase. For example, Figure 4a gives the 11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 23

corresponding large scale STM image of a rhomboid phase (Type II phase). Annealing the sample at 200 ℃ for 20 mins leads to significant decrease of the molecular coverage by the considerable molecular desorption, as shown in Figure 4b. Close investigation reveals that the all the remaining self-assembly domains (as highlighted by the white boxes) show porous structures, as depicted by the high resolution STM topographic image (Figure 4c). Similar control experiments have been performed on the close-packed (Type III) and stripe (Type IV) phases. For both cases, porous phases are observed after sufficient annealing. The controlling experiments suggest that the porous phase at lower molecular coverage is thermodynamically favorable. Table 1. Unit Cell Parameters and Molecular Density of Each Self-Assembly Phase of DDA Molecules on Au(111) Phases

Number (n) of

Unit cell parameters

molecules per

Density ρ (molecules/nm2)

a (nm)

b (nm)

θ (°)

unit cell

℃ Porous

3.54 ± 0.02

3.54 ± 0.02

60 ± 2

6

0.552

℃ Rhomboid

2.00 ± 0.02

1.70 ± 0.02

90 ± 2

2

0.588

℃ Close packed

2.35 ± 0.02

2.03 ± 0.02

90 ± 2

3

0.628

℃ Stripe

2.94 ± 0.01

0.50 ± 0.01

85 ± 1

1

0.682

Figure 4. Transferring from the rhomboid phase to the porous phase by thermal annealing. (a) STM image (It = 20 pA, Vb = -1 V) of a Au(111) sample covered with 12

ACS Paragon Plus Environment

Page 13 of 23 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

The Journal of Physical Chemistry

0.1 ± 0.02 ML DDA molecules. (b) STM image (It = 20 pA, Vb = -1 V) of the sample after annealing the sample given in (a) at 200 ℃ for 20 mins. (c) High resolution STM image of white boxes (It = 500 pA, Vb = -1 V) of (b). The scale bar is 10 nm for (a), 100 nm for (b) and 2 nm for (c).

CONCLUSIONS Abundant self-assembly phases have been observed utilizing scanning tunneling microscopy after depositing DDA molecules on Au(111) surfaces. By systematically varying the molecular coverage and the evaporation rate, a phase diagram of the self-assembly structures is obtained. Combing with the DFT calculations, we found out that the phase at lower coverage is thermodynamically determined. At higher coverage or higher evaporation rate, the influence of the kinetic effect cannot be neglected, which leads to the formation of denser packings. This systematic study for controllable phase evolutions on metal surfaces provides fundamental insight into the competition of thermodynamic and kinetic effects in the self-assembly.

ASSOCIATED CONTENT Supporting Information The STM images of the sample prepared with various evaporation temperatures; The contribution of the hydrogen bonds in the total energy. AUTHOR INFORMATION Corresponding Authors 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 23

*E-mail: [email protected]

*E-mail: [email protected]

ACKNOWLEDGMENTS We acknowledge the Collaborative Innovation Center of Suzhou Nano Science & Technology and the Priority Academic Program Development of Jiangsu Higher Education Institutions. We thank Prof. Gerhard Erker in the University of Münster for providing the DDA molecules. This work was supported by National Science Foundation of China (21790053, 21622306, 91545127, 21403149, 21872099), National

Major

State

Basic

Research

Development

Program

of

China

(2017YFA0205002, 2014CB932600) and Natural Science Foundation of Jiangsu Province (BK20140305).

14

ACS Paragon Plus Environment

Page 15 of 23 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

The Journal of Physical Chemistry

REFERENCES (1) Mas-ballesté, R.; Gómeznavarro, C.; Góme-zherrero, J.; Zamora, F. 2D Materials: To Graphene and Beyond. Nanoscale 2011, 3, 20-30. (2) Barth, J. V. Molecular Architectonic on Metal Surfaces. Annu. Rev. Phys. Chem. 2007, 58, 375-407. (3) Wan, L. J. Fabricating and Controlling Molecular Self-Organization at Solid Surfaces: Studies by Scanning Tunneling Microscopy. Acc. Chem. Res. 2006, 39, 334-342. (4) Chaki, N. K.; Vijayamohana, K. Self-Assembled Monolayers as a Tunable Platform for Biosensor Applications. Biosens. Bioelectron. 2002, 17, 1-12. (5) Arya, S. K.; Solanki, P. R.; Datta, M.; Malhotra, B. D. Recent Advances in Self-Assembled Monolayers Based Biomolecular Electronic Devices. Biosens. Bioelectron. 2009, 24, 2810-2817. (6) De Feyter, S.; De Schryver, F. C. Self-Assembly at the Liquid/Solid Interface: STM Reveals. J. Phys. Chem. B 2005, 109, 4290-4302. (7) De Feyter, S.; De Schryver, F. C. Two-Dimensional Supramolecular Self-Assembly Probed by Scanning Tunneling Microscopy. Chem. Soc. Rev. 2003, 32, 139-150. (8) De Feyter, S.; Gesquière, A.; Abdelmottaleb, M. M.; Grim, P. C. M.; De Schryver, F. C.; Meiners, C.; Sieffert, M.; Valiyaveettil, S.; Müllen, K. Scanning Tunneling Microscopy: A Unique Tool in the Study of Chirality, Dynamics, and Reactivity in Physisorbed Organic Monolayers. Acc. Chem. Res. 2000, 33, 520-531. 15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

(9) Giancarlo, L. C.; Flynn, G. W. Raising Flags:  Applications of Chemical Marker Groups to Study Self-Assembly, Chirality, and Orientation of Interfacial Films by Scanning Tunneling Microscopy. Acc. Chem. Res. 2000, 33, 491-501. (10) Kühnle, A. Self-Assembly of Organic Molecules at Metal Surfaces. Curr. Opin. Colloid. Interface Sci. 2015, 20, 227-234. (11) Barth, J. V.; Costantini, G.; Kern, K. Engineering Atomic and Molecular Nanostructures at Surfaces. Nature 2005, 437, 671-679. (12) Bellec, A.; Arrigoni, C.; Schull, G.; Douillard, L.; Fiorini-Debuisschert, C.; Mathevet, F.; Kreher, D.; Attias, A-J.; Charra, F. Solution-Growth Kinetics and Thermodynamics of Nanoporous Self-Assembled Molecular Monolayers. J. Chem. Phys. 2011, 134, 124702. (13) Blunt. M. O.; Adisoejoso, J.; Tahara, K.; Katayama, K.; Auweraer, M. V. D.; Tobe, Y.; De Feyter, S. Temperature-Induced Structural Phase Transitions in a TwoDimensional Self-Assembled Network. J. Am. Chem. Soc. 2013, 135, 12068-12075. (14) Gutzler, R.; Sirtl, T.; Diensmaier, J. F.; Mahata, K.; Heckl, W. M.; Schmittel, M.; Lackinger, M. Reversible Phase Transitions in Self-Assembled Monolayers at the Liquid-Solid Interface: Temperature-Controlled Opening and Closing of Nanopores. J. Am. Chem. Soc. 2010, 132, 5084-5090. (15) Wang, C.; Jana, P. K.; Zhang, H. M.; Mu, Z. C.; Kehr, G.; Blömker, T.; Eeker, G.; Fuchs, H.; Heuer, A.; Chi, L. F. Controlling Two-Phase Self-Assembly of an Adenine Derivative on HOPG via Kinetic Effects. Chem. Commun. 2014, 50, 9192-9195. 16

ACS Paragon Plus Environment

Page 16 of 23

Page 17 of 23 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

The Journal of Physical Chemistry

(16) Li, Y.; Lin, N. Combined Scanning Tunneling Microscopy and Kinetic Monte Carlo Study on Kinetics of Cu-Coordinated Pyridyl-Porphyrin Supramolecular Self-Assembly on a Au(111) Surface. Phys. Rev. B 2011, 84, 125418. (17) Sirtl, T.; Schlögl, S.; Rastgoo-Lahrood, A.; Jelic, J.; Neogi, S.; Schmittel, M.; Heckl, W. M.; Reuter, K.; Lackinger, M. Control of Intermolecular Bonds by Deposition Rates at Room Temperature: Hydrogen Bonds versus Metal Coordination in Trinitrile Monolayers. J. Am. Chem. Soc. 2013, 135, 691-695. (18) Tahara, K.; Okuhata, S.; Adisoejoso, J.; Lei, S. B.; Fujita, T.; De Feyter, S.; Tobe, Y. 2D Networks of Rhombic-Shaped Fused Dehydrobenzo[12]Annulenes: Structural Variations under Concentration Co ntrol. J. Am. Chem. Soc. 2009, 131, 17583-17590. (19) Lei, S. b.; Tahara, K.; De Schryver, F. C.; Auweraer, M. V. D.; Tobe, Y.; De Feyter, S. One Building Block, Two Different Supramolecular Surface-Confined Patterns: Concentration in Control at the Solid-Liquid Interface. Angew. Chem. Int. Ed. 2008, 47, 2964-2968. (20) Kampschulte, L.; Werblowsky, T. L.; Kishore. R. S. K.; Schmittel, M.; Heckl, W. M.; Lackinger, M. Thermodynamical Equilibrium of Binary Supramolecular Networks at the Liquid-Solid Interface. J. Am. Chem. Soc. 2008, 130, 8502-8507. (21) Shen, X. L.; Wei, X. D.; Tan, P. L.; Yu, Y. G.; Yang, B.; Gong, Z. M.; Zhang, H. M.; Lin, H. P.; Li, Y. Y.; Li, Q.; et al. Concentration-Controlled Reversible Phase Transitions in Self-Assembled Monolayers on HOPG Surfaces. Small 2015, 19, 2284-2290. 17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

(22) Balandina, T.; Tahara, K.; Sändig, N.; Blunt, M. O.; Adisoejoso, J.; Lei, S. B.; Zerbetto, F.; Tobe, Y.; De Feyter, S. Role of Substrate in Directing the Self-Assembly of Multicomponent Supramolecular Networks at the Liquid-Solid Interface. ACS Nano 2012, 9, 8381-8389. (23) Piot, L.; Marchenko, A.; Wu, J. S.; Müllen, K.; Fichou, D. Structural Evolution of Hexa-peri-Hexabenzocoronene Adlayers in Heteroepitaxy on n-Pentacontane Template Monolayers. J. Am. Chem. Soc. 2005, 127, 16245-16250. (24) Mamdouh, W.; Uji-i, H.; Ladislaw, J. S.; Dulcey, A. E.; Percec, V.; De Schryver, F. C.; De Feyter, S. Solvent Controlled Self-Assembly at the Liquid-Solid Interface Revealed by STM. J. Am. Chem. Soc. 2006, 128, 317-325. (25) Tahara, K.; Furukawa, S.; Uji-i, H.; Uchino, T.; Ichikawa, T.; Zhang, J.; Mamdouh, W.; Sonoda, M.; De Schryver, F. C.; De Feyter, S.; et al. Two-Dimensional Porous Molecular Networks of Dehydrobenzo[12]Annulene Derivatives via Alkyl Chain Interdigitation. J. Am. Chem. Soc. 2006, 128, 16613-16625. (26) Merz, L.; Parschau, M.; Zoppi, L.; Baldridge, K. K.; Siegel, J. S.; Ernst, K-H. Reversible Phase Transitions in a Buckybowl Monolayer. Angew. Chem. Int. Ed. 2009, 48, 1966-1969. (27) Ruben, M.; Payer, D.; Landa, A.; Comosso, A.; Gattinoni, C.; Lin, N.; Collin, J-P.; Sauvage, J-P.; Vita, A. D.; Kern, K. 2D Supramolecular Assemblies of Benzene-1,3,5-triyl-tribenzoic Acid: Temperature-Induced Phase Transformations

18

ACS Paragon Plus Environment

Page 18 of 23

Page 19 of 23 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

The Journal of Physical Chemistry

and Hierarchical Organization with Macrocyclic Molecules. J. Am. Chem. Soc. 2006, 128, 15644-15651. (28) Walch, H.; Gutzler, R.; Sirtl, T.; Eder, G.; Lackinger, M. Material- and Orientation-Dependent Reactivity for Heterogeneously Catalyzed Carbon-Bromine Bond Homolysis. J. Phys. Chem. C 2010, 114, 12604-12609. (29) Kühnle, A.; Linderoth, T. R.; Schunack, M.; Besenbacher, F. L-Cysteine Adsorption Structures on Au(111) Investigated by Scanning Tunneling Microscopy under Ultrahigh Vacuum Condition. Langmuir 2006, 22, 2156-2160. (30) Kudernac, T.; Lei, S. B.; Elemans, J. A. A. W.; De Feyter, S. Two-dimensional Supramolecular Self-Assembly: Nanoporous Networks on Surfaces. Chem. Soc. Rev. 2009, 38, 402-421. (31) Lackinger, M.; Griessl, S.; Kampschulte, L.; Jamitzky, F.; Heckl, W. M. Dynamics of Grain Boundaries in Two-Dimensional Hydrogen-Bonded Molecular Networks. Small 2005, 5, 532-539. (32) Liang, H. L.; He, Y.; Ye, Y. C.; Xu, X. G.; Cheng, F.; Sun, W.; Shao, X.; Wang, Y. F.; Li, J. L.; Wu, K. Two-Dimensional Molecular Porous Networks Constructed by Surface Assembling. Coord. Chem. Rev. 2009, 253, 2959-2979. (33) Elemans, J. A. A. W.; Lei, S. B.; De Feyter, S. Molecular and Supramolecular Networks on Surfaces: From Two-Dimensional Crystal Engineering to Reactivity. Angew. Chem. Int. Ed. 2009, 48, 7298-7332.

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

(34) Uemura, S.; Tanoue, R.; Yilmaz, N.; Ohira, A.; Kunitake, M. Molecular Dynamics in Two-Dimensional Supramolecular Systems Observed by STM. Materials 2010, 3, 4252-4276. (35) Horcas, I.; Fernandez, R.; Gomez-Rodriguez, J. M.; Colchero, J.; Gomez-Herrero, J.; Baro, A. M. WSxM: A Software for Scanning Probe Microscopy and a Tool for Nanotechnology. Rev. Sci. Instrum. 2007, 78, 013705. (36) Kresse, G.; Hafner, J. Ab. initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558-561 (37) Kresse, G.; Furthmuller, J. Efficient Iterative Schemes for ab initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169-11186. (38) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865-3868. (39) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate ab initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (40) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456-1465. (41) Bahn, S. R.; Jacobsen, K. W. An Object-Oriented Scripting Interface to a Legacy Electronic Structure Code. Comput. Sci. Eng. 2002, 4, 56-66.

20

ACS Paragon Plus Environment

Page 20 of 23

Page 21 of 23 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

The Journal of Physical Chemistry

(42) Hanke, F.; Haq, S.; Raval, R.; Persson, M. Heat-to-Connect: Surface Commensurability Directs Organometallic One-Dimensional Self-Assembly. ACS Nano 2011, 5, 9093-9103. (43) Dyer, M. S.; Robin, A.; Haq, S.; Raval, R.; Persson, M.; Klimes, J. Understanding the Interaction of the Porphyrin Macrocycle to Reactive Metal Substrates: Structure, Bonding and Adatom Capture. ACS Nano 2011, 5, 1831-1838. (44) Cai, Y. G.; Bernasek, S. L. Structures Formed by the Chiral Assembly of Racemic Mixtures of Enantiomers: Iodination Products of Elaidic and Oleic Acids. J. Phys. Chem. B 2005, 109, 4514-4519. (45) Tao, F.; Goswami, J.; Bernasek, S. L. Self-Assembly and Odd-Even Effects of cis-Unsaturated Carboxylic Acids on Highly Oriented Pyrolytic Graphite. J. Phys. Chem. B 2006, 110, 4199-4206. (46) Kim, K.; Plass, K. E.; Matzger, A. J. Kinetic and Thermodynamic Forms of a Two-Dimensional Crystal. Langmuir 2003, 19, 7149-7152. (47) Zou, B.; Dreger, K.; Mück-Lichtenfeld, C.; Grimme, S.; Schäfer, H. J.; Fuchs, H.; Chi, L. F. Simple and Complex Lattices of N-Alkyl Fatty Acid Amides on a Highly Oriented Pyrolytic Graphite Surface. Langmuir 2005, 21, 1364-1370. (48) Santos, M. S.; Tavares, F. W.; Biscaia, E. C. Molecular Thermodynamics of Micellization: Micelle Size Distributions and Geometry Transitions. Braz. J. Chem. Eng. 2016, 33, 515-523.

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

(49) Enache, M.; Maggini, L.; Llanes-Pallas, A.; Jung, T. A.; Bonifazi, D.; Stöhr, M. Coverage-Dependent Disorder-to-Order Phase Transformation of a Uracil Derivative on Ag(111). J. Phys. Chem. C 2014, 118, 15286-15291. (50) Lackinger, M.; Heckl, W. M. Carboxylic Acids: Versatile Building Blocks and Mediators for Two-Dimensional Supramolecular Self-Assembly. Langmuir 2009, 25, 11307-11321. (51) Gutzler, R.; Cardenas, L.; Rosei, F. Kinetics and Thermodynamics in Surface-Confined Molecular Self-Assembly. Chem. Sci. 2011, 2, 2290-2300. (52) Tsuzuki, S.; Lüthi, H. P. Interaction Energies of van der Waals and Hydrogen Bonded Systems Calculated Using Density Functional Theory: Assessing the PW91 Model. J. Chem. Phys. 2011, 114, 3949-3957.

22

ACS Paragon Plus Environment

Page 22 of 23

Page 23 of 23 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

The Journal of Physical Chemistry

TOC GRAPHIC

23

ACS Paragon Plus Environment