Next Article in Journal
Bismuth Oxychloride Nanomaterials Fighting for Human Health: From Photodegradation to Biomedical Applications
Previous Article in Journal
Layered Double Hydroxides (LDHs) as New Consolidants for Cultural Heritage Masonry
Previous Article in Special Issue
Molecular Self-Assembly of an Unusual Dinuclear Ruthenium(III) Complex Based on the Nucleobase Guanine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Monte Carlo Simulations of the Metal-Directed Self-Assembly of Y-Shaped Positional Isomers

by
Karolina Nieckarz
1 and
Damian Nieckarz
2,*
1
Faculty of Chemistry, Maria Curie-Skłodowska University in Lublin, Maria Curie-Skłodowska Square 3, 20-031 Lublin, Poland
2
Department of Theoretical Chemistry, Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie-Skłodowska University in Lublin, Maria Curie-Skłodowska Square 3, 20-031 Lublin, Poland
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(4), 492; https://doi.org/10.3390/cryst12040492
Submission received: 26 February 2022 / Revised: 26 March 2022 / Accepted: 30 March 2022 / Published: 1 April 2022
(This article belongs to the Special Issue Self-Assembled Complexes: “Love at First Sight”)

Abstract

:
The rational fabrication of low-dimensional materials with a well-defined topology and functions is an incredibly important aspect of nanotechnology. In particular, the on-surface synthesis (OSS) methods based on the bottom-up approach enable a facile construction of sophisticated molecular architectures unattainable by traditional methods of wet chemistry. Among such supramolecular constructs, especially interesting are the surface-supported metal–organic networks (SMONs), composed of low-coordinated metal atoms and π -aromatic bridging linkers. In this work, the lattice Monte Carlo (MC) simulation technique was used to extract the chemical information encoded in a family of Y-shaped positional isomers co-adsorbed with trivalent metal atoms on a flat metallic surface with (111) geometry. Depending on the intramolecular distribution of active centers (within the simulated molecular bricks, we observed a metal-directed self-assembly of two-dimensional (2D) openwork patterns, aperiodic mosaics, and metal–organic ladders. The obtained theoretical findings could be especially relevant for the scanning tunneling microscopy (STM) experimentalists interested in a surface-assisted construction of complex nanomaterials stabilized by directional coordination bonds.

Graphical Abstract

1. Introduction

In arts and crafts, the openwork ornamentation made by cutting out holes with well-defined geometrical shapes in a preselected material (fabric, clay, wood, etc.), is particularly popular. Thanks to the holes, a beam of light is introduced into the artistic composition, which shapes the object with a play of chiaroscuro and subtly emphasizes its fragile internal structure [1]. Interestingly, many one-atom-thick molecular assemblies, such as surface-supported metal–organic networks (SMONs), exhibit a non-trivial, openwork structure [2,3,4,5,6,7]. This is due to the presence of empty cavities (nanopores) within their structure that could differ in shape, spatial distribution, and chirality [8,9,10]. Moreover, the openwork supramolecular architectures stabilized by directional ligand metal coordination bonds are not only aesthetically pleasing but could also find practical applications in nanotechnology (as selective molecular sieves, atomically defined adsorbents, surface protective coatings, etc.), nanotribology, and heterogeneous catalysis [11,12,13,14].
To date, numerous molecular precursors with a different symmetry and predefined shape, size, and intramolecular distribution of functional groups have been synthesized and successfully applied as building blocks of a plethora of SMONs with tailorable morphology, porosity, and functions [2,4,5,9,11,15,16,17,18,19]. Among them, especially prominent is a family of small star-shaped molecules with C 3 -symmetrical backbones comprising interconnected phenyl moieties. In recent decades, such carbon-based molecular bricks made possible the on-surface synthesis (OSS) of diverse metal–organic networks [4,13,16,18,19,20,21,22], sophisticated planar tessellations [23], ladders [24], nanoribbons, and small isolated aggregates [25,26]. However, much less attention so far has been paid to the analogous carbon-based molecules with a reduced symmetry [27,28,29]. For these molecular building blocks, the outcome of metal-directed self-assembly is usually hard to predict. It is especially true in the case of conformationally flexible molecules with high structural heterogeneity, which could exist in different chemical forms, such as tautomers, conformers, enantiomers, etc. [30,31].
In our previous work, we demonstrated that the tripodal ligands with properly assigned interaction centers could be valuable building blocks of the novel, metal–organic architectures [32]. However, the performed computer simulations were limited only to the molecular bricks with C 3 -symmetrical backbones. Therefore, in this contribution, we extend our theoretical investigations on a family of Y-shaped positional isomers whose arms differ in length and are equipped with active centers (functional groups) embedded in para and/or meta positions. To that end, we have applied a classical lattice Monte Carlo (MC) method, which allows for a precise investigation of complex molecular assemblies [33,34,35]. Moreover, the lattice MC simulations, based on discrete intermolecular potentials, enable us to avoid tedious, time-consuming force calculations [36,37]. The main objective of our work is to extract the chemical information encoded in the studied molecules and exploit the gained hints to steer the morphology of surface-confined patterns, comprising Y-shaped bridging linkers, and trivalent metal atoms. The obtained results could be especially useful in the context of surface-assisted fabrication of novel openwork nanomaterials, metal–organic ladders, and aperiodic mosaics with untrivial topologies.

2. Materials and Methods

To explore the surface-supported self-assembly of positional isomers AJ (see Figure 1) the coarse-grained Monte Carlo (MC) simulations were performed in the N L T ensemble, where N = l + m is the total number of adsorbed entities (where l = 600 is the number of Y-shaped ligands, and m = l metal atoms), L stands for the linear size of the simulated system ( L = 130 ), and T is temperature. The MC computer calculations were performed on a rhombic fragment of flat triangular lattice comprising L × L energetically equivalent adsorption sites (vertices of a triangular lattice) at low surface coverage, defined as θ = ( 5 l + m ) L 2 (if not stated otherwise). In the adopted model, the two-dimensional (2D) triangular lattice represents an atomically smooth crystalline M ( 111 ) surface with no defects and corrugations, where M = Cu ,   Ag , or Au . The distance between two neighboring adsorption sites on a triangular lattice was arbitrarily set to 1 and, for the sake of simplicity, expressed in dimensionless units. To minimize the influence of negative edge-effects on the morphology of the resulting molecular patterns, imposed were periodic boundary conditions in all planar directions of a 2D lattice. Additionally, to avoid the formation of unwanted point structural defects (missing ligands/metal atoms, self-interstitial defect, etc.), the annealing procedure was applied. Namely, the simulated overlayer was linearly cooled down from the initial temperature, T = 1.001 , to the target temperature T = 0.001 in 10 3 equal intervals d T = 0.001 , comprising 6 N · 10 3 MC steps each (one MC step is a single displacement/in-plane rotation trial of randomly selected entity adsorbed on a triangular lattice).
The MC simulations started with a random distribution of l tripodal ligands and m trivalent metal atoms on a triangular lattice in an optimal l / m = 1 : 1 stoichiometric ratio. Then, in the next step, the investigated metal–organic overlayer was energetically equilibrated in a series of 6 N · 10 6 MC steps. For this purpose, one component of the adsorbed overlayer was selected at random, and its total potential energy E 1 was determined. Then, the chosen ligand/metal atom was displaced within the 2D lattice or rotated in place (only if the Y-shaped molecule was selected) by a magnitude of acute angle ± 60. If this operation was successful, its total potential energy E 2 in a new position Ω was calculated. Otherwise, the tested entity was left in the original configuration on a lattice, and the above outlined procedure was successively iterated. The ligand/metal atom displacement attempt was accepted/rejected based on the Metropolis MC criterion [38,39] involving a calculation of transition probability p = m i n { 1 , e x p ( Δ E / k B T ) } , where Δ E = E 2 E 1 is the potential energy difference between the new and the old position of the probed entity, k B denotes the reduced Boltzmann constant ( k B = 1 in our model), and T is a system temperature. In the case where E 2 E 1 , the new position Ω was accepted unconditionally. Otherwise, the random number r ( 0 , 1 ) was generated and compared with a transition probability p . If a random number r satisfied inequality r p , the new position Ω was accepted. In the latter case ( r > p ), the tested configuration was rejected, and the ligand/metal atom was moved back to the original position on a triangular lattice.
In the adopted theoretical model, each of the π -aromatic positional isomers AJ is represented by a group of in-plane interconnected hard spheres (segments) with properly assigned active centers (black arrows in Figure 1), while metal atoms are modeled as single flat segments. In our coarse-grained approach, a single gray segment corresponds to one phenyl ring of real surface isomer AJ, while segment–segment connections represent C–C sigma bonds linking aromatic moieties of the studied molecules. Additionally, each gray segment present within the simulated linker AJ covers only one adsorption site on a triangular lattice (localized, monolayer adsorption). The same is true in the case of a single metal atom. Contrary to the classical Langmuir adsorption model, the desorption of components of the investigated metal–organic overlayers is prohibited ( N L T ensemble). Moreover, lateral interactions between physisorbed Y-shaped ligands and metal atoms on the adjacent adsorption sites of a triangular lattice are possible—the in-plane ligand metal interactions were assumed to be short-ranged (limited to the neighboring adsorption sites on a lattice), anisotropic, equal in strength ( ω = 1.0 , negative value means attraction), and reversible. We assumed that a single coordination bond could only be formed if the metal atom m and linker l met on a triangular lattice in a proper spatial configuration; one of the outer segments of the Y-shaped linker and the metal atom cover two adjacent adsorption sites (vertices of a triangular lattice), and interaction directions (black arrows) assigned to them are oppositely directed ( ) and collinear. To solely extract the chemical information encoded in the geometric shape and the intramolecular distribution of active centers of the positional isomers AJ, we assumed that the potential energy of an anisotropic coordination bond prevails over other interactions possible in the studied metal–organic overlayers. For this purpose, all potential energies associated with the ligand surface, metal surface, metal metal, and ligand ligand side-to-side interactions were set to zero. Additionally, to compare the structural properties of the ordered 2D metal–organic patterns, we introduced the structural parameter called density   ρ . This dimensionless parameter was defined as the ratio of occupied adsorption sites inside a unit cell of arbitrary selected phase and the total area A of this unit cell, that is ρ = ( 5 x + y ) A 1 , where x and y , respectively, denote the number of Y-shaped linkers and metal atoms within a unit cell.

3. Results and Discussion

Initial MC simulations were conducted for the C 1 -symmetric positional isomer A, the arms of which are functionalized exclusively in para positions. Such intramolecular distribution of identical active centers makes linker A achiral, which is manifested in a relatively simple morphology of the emerging supramolecular patterns. However, even such a trivial functionalization scheme combined with a reduced symmetry of molecule A directed the surface-confined self-assembly into two polymorphic superstructures with distinct morphological properties such as density ρ , pore size/shape, and their spatial distribution (see Figure 2a,b).
As seen in Figure 2a, the first polymorph ( ρ = 0.371 ) is characterized by a rhombic unit cell ( 7 × 7 ,   60 / 120 o ) comprising three linkers A and three metal atoms in a 1:1 stoichiometric ratio. The cavities present within the corresponding network are periodically distributed and arranged into a flower-like pattern with a central hexagonal void (7 adsorption sites) and six petals (12 adsorption sites each) with the shape of a truncated equilateral triangle. In comparison to the discussed phase, the brick wall-shaped polymorphic network shown in Figure 2b has a ~ 15 % higher density ( ρ = 0.433 ). Its parallelogram unit cell ( 2 3 × 19 ,   67 / 113 o ) contains only one linker A and one metal atom. Moreover, all Y-shaped ligands A present within this homogenous superstructure are oriented unidirectionally. The adjacent pores are also oriented along the same direction of a triangular lattice and comprise 10 adsorption sites each. Although the discussed 2D polymorphs are almost energetically equivalent, the surface-confined self-assembly of a heteroporous network illustrated in Figure 2a is privileged due to entropic reasons. Namely, after a visual inspection of 50 independent system replicas, it revealed that the two-porous phase was formed selectively in ~ 66 % of them. This result could be reasonably explained by comparing the number of spatial orientations of linkers A within the polymorphic phases—in the superstructure shown in Figure 2a each bridging linker A could be incorporated into the growing metal–organic network in three different ways, whereas in the case of a uniporous network (Figure 2b), only a unidirectional arrangement of ligands A could result in formation of a defect-free overlayer. Interestingly, similar crystalline polymorphs were observed experimentally in the case of 3,3′,5,5′-tetrabromo-2,2′,4,4′,6,6′-hexafluorobiphenyl (Br4F6BP) molecules [40], brominated tetrathienoanthracene [41], as well as a theoretical model [42]. In the light of the above-listed scientific reports, the discussed two-directional self-assembly is no exception but rather the next example of a dichotomous formation of complex metal–organic patterns.
Contrary to the achiral molecule A, the next studied positional isomer B is functionalized at one (S)-meta and two para positions. As shown in Figure 3a, this intramolecular distribution of active centers (functional groups G) allows for a highly selective formation of the openwork metal–organic network with huge cavities (55 adsorption sites) surrounded concentrically by smaller pores comprising 12 and 2 adsorption sites each. The gear-like shaped voids have diameter d = 8 (corner-to-corner) and a unique set of teeth differing in length, that is ( 1.1.3 ) 6 . In the adopted notation, an individual number denotes the length of a single tooth, while the number sequence corresponds to their order of occurrence at the edges of the largest pore. The characteristic feature of the heteroporous phase is a relatively low density ρ = 0.344 and an exceptionally large rhombic unit cell ( 11 × 11 ,   60 / 120 o ) , comprising six molecules B and six metal atoms in a 1:1 stoichiometric ratio.
Embedding an active center on the second short arm of tritopic molecule B in the (S)-meta position gives a subsequent positional isomer, denoted by the letter C. In this molecule, the (S)-meta active center is pointed towards its long arm, monofunctionalized in the para position. Interestingly, even such a tiny structural modification strongly affects the topology of the resulting supramolecular network. As illustrated in Figure 3b, the infinite metal–organic phase exhibits basic structural features of the adsorbed overlayer from Figure 3a. The gear-like voids ( 1.1.2 ) 6 of the considered superstructure encompass 31 adsorption sites and are separated by the adjacent small pores comprising 12 and 3 free adsorption sites. Additionally, with a relatively small diameter of the largest pores ( d = 6 ) , a substantial density follows, increasing the crystalline structure ( ρ = 0.416 ) as well as the side length reduction of its unit cell ( 10 × 10 ,   60 / 120 o ) . Surprisingly, a displacement of the para active center attached to the short arm of linker C to the (S)-meta position (within the same outer segment) directed the metal-assisted self-assembly into the subsequent open porous network (see Figure 4a) with vast cavities, characteristic for the overlayer illustrated in Figure 3a. However, within the discussed adlayer, the gear-like voids ( 1.1.2 ) 6 are separated by alternatively arranged small pores with a star-shaped (four adsorption sites) and needle-like geometry (three adsorption sites). The rhombic unit cell of the corresponding supramolecular structure contains six surface isomers D and six metal atoms in a 1:1 stoichiometric ratio.
In contrast to the above-discussed molecular assemblies, in the case of isomer E, which arms are functionalized in different positions—in para/(R)-meta/(S)-meta mode counting anticlockwise—we did not observe an emerging open porous phase with translational symmetry. Instead, the isolated metal–organic ladders were obtained. In the exemplary snapshot shown in Figure 4b, each ladder comprises alternately arranged small pores comprising four and two adsorption sites, while the metal atoms incorporated into its structure are arranged in a zig-zag fashion (blue line in the magnified fragment of Figure 4b). Although the metal–organic ladders are randomly distributed on a triangular lattice, all of them are pointed unidirectionally. Such a preferred spatial configuration is not surprising, as it provides unperturbed elongation (metal-directed growth) of individual ladders.
As shown in Figure 5a, also in the case of positional isomer F, the obtained metal–organic phase differs largely in morphology from the supramolecular architectures depicted in Figure 2 and Figure 4a. Namely, its main feature is the lack of translational order. As shown in the magnified fragment of Figure 4a, the multiporous phase is built of three types of empty cavities marked in different colors: pink (10 adsorption sites), chalk (7 adsorption sites), and light-green (4 adsorption sites). Interestingly, the basic morphological properties of the considered overlayer are also conserved at a two times higher surface coverage (see Figure 5b).
The aperiodicity of the overlayers shown in Figure 5a,b originates exclusively from a non-regular distribution of bridging linkers F, whereas trivalent metal atoms occupy vertices of a well-defined triangular lattice with a primitive unit cell (see Figure 6b). Another unique feature of the aperiodic mosaics from Figure 5a,b are the randomly distributed flower-like rosettes with closed and open walls. As shown in the magnified fragment of Figure 6a, the central point of a closed hexagonal rosette is a windmill-like three-fold coordination node surrounded by two coordination shells comprising a periodical array of metal atoms and molecules F. However, this perfect spatial arrangement of bridging linkers/metal atoms could be easily perturbed without generating any point structural defects. Therefore, in the mosaic superstructure, many open-walled rosettes could be easily identified. Within the aperiodic metal–organic network from Figure 5a, the most abundant are irregular pores marked in chalk color ( ~ 50%), while the light-green and pink elongated cavities are almost equally populated and constitute ~ 27% and ~ 23% of all empty pores present in the considered overlayer (averages from five independent system replicas).
Even after the cursory visual inspection of the aperiodic mosaics illustrated in Figure 5a,b, it becomes clear that the metal-directed self-assembly of molecules F could produce also other metal–organic patterns with a highly periodic distribution of bridging linkers F and metal atoms. Small fragments of such artificially created connections are shown in Figure 7. Especially interesting among them are the overlayers composed of flower-like rosettes (Figure 7a,b). The first of them ( ρ = 0.533 ) is built of adjacent hexagonal rosettes with central windmill-like coordination nodes oriented clockwise (see Figure 7a). The second polymorph, with a huge rhombic unit cell ( 234 × 234   60 / 120 o ), also comprises closed hexagonal rosettes, but some of their central windmill-like nodes are oriented anticlockwise (Figure 7b). Moreover, one could imagine a morphologically similar phase (with equal density ρ ), where the windmill-like coordination nodes are randomly oriented. The third artificial overlayer, shown in Figure 7c, resembles an open porous network obtained for molecule C (see Figure 3b). However, within this phase, the gear-like cavities ( 1.1.2 ) 6 with a diameter d = 6 (corner-to-corner), are surrounded concentrically by smaller pores comprising four and six adsorption sites. The last crystalline pattern is homogenous in terms of porosity ( ρ = 0.533 ) and comprises only irregular voids containing seven adsorption sites each (Figure 7d).
The next studied molecule, denoted by the letter G, is characterized by a highly symmetrical distribution of active centers (see Figure 8). Namely, the long arm of anchor-like ligand G is equipped with an active center in the para position, whereas the remaining two shorter arms (equal in length) are functionalized in meta positions with opposite in-plane orientations: (R) and (S). Consequently, both meta active centers are pointed to the para functionalized arm of C 1 -symmetric molecule G. Such a unique functionalization scheme allows this isomer to self-assemble with trivalent metal atoms into a highly aperiodic superstructure with no structural defects (see Figure 8a). As shown in the magnified fragment of Figure 8a, the complex mosaic phase is composed of six different cavities. Among them are kite-like voids (○, ○–○, ○–○–○) encompassing 6, 12, and 18 free adsorption sites; elongated saw-like pores ● (nine adsorption sites), and ●–● (15 adsorption sites); and small needle-like voids, comprising only three adsorption sites. As in the case of the mosaic overlayers depicted in Figure 5a,b, the aperiodicity of the considered metal–organic network originates only from a non-regular distribution of anchor-like linkers G, while the isolated metal atoms occupy vertices of a periodic triangular lattice with primitive unit cell ( 2 3 × 2 3 ,   60 / 120 o ) . Surprisingly, at high surface coverage ( θ ~ 0.53 ), some fraction of molecules G is engaged in the formation of metal–organic ladders coexisting with an aperiodic phase (see Figure 8b). As shown in a magnified fragment of Figure 8b, the metal–organic ladders are close-packed and oriented along the same direction of a triangular lattice. However, the adjacent ladders comprising small needle-like voids do not create a periodic crystalline phase, as some of them are translated in-plane with respect to the others. A closer visual inspection of an aperiodic superstructure shown in Figure 8a reveals that similar ladder-like metal–organic connections are also present in its structure. However, the lengths of these ordered connections are relatively small; most of the metal–organic ladders incorporated into the phase illustrated in Figure 8a comprise only from two to five parallel oriented linkers G.
Unlike the aperiodic mosaics created by the positional isomers F and G, the supramolecular pattern obtained for molecule H is highly regular. As seen in Figure 9a, the open porous network comprises three types of nanocavities. Among them are large gear-like shaped voids ( 1.2.2 ) 6 with a diameter d = 2 13 (corner-to-corner), hexagonal pores (seven adsorption sites), and truncated rhombus-like small voids (four adsorption sites). Inside the rhombic unit cell ( 111 × 111 ,   69 / 111 o ) of the crystalline overlayer, there are six molecules H and six metal atoms. Moreover, the multiporous phase has low density ( ρ = 0.346 ) , comparable with the metal–organic array composed of surface isomers B mixed with trivalent metal atoms in a 1:1 molar ratio (see Figure 3a). A very similar topology corresponds also to the defect-free overlayer comprising linkers I. As shown in Figure 9b, the infinite metal–organic phase ( ρ = 0.381 ) is composed of large voids ( 1.2.2 ) 6 with diameter d = 2 13 (corner-to-corner), windmill-like pores (nine adsorption sites), and very small cavities encompassing only two adsorption sites. The rhombic unit cell of the considered network ( 109 × 109 ,   60 / 120 o ) contains six ligands I and six metal atoms in a 1:1 stoichiometric ratio, which agrees perfectly with the overall ligand/metal proportion in the system under study.
For the last examined molecule (denoted as J), with arms functionalized selectively in (S)-meta positions, we observed the formation of two polymorphs differing in terms of porosity. As seen in Figure 10a,b, both metal–organic networks are very similar to those obtained for achiral linker A (see Figure 2a,b). Namely, the multiporous overlayer shown in Figure 10a is characterized by a rhombic unit cell ( 37 × 37 ,   60 / 120 o ) comprising three ligands J and three trivalent metal atoms in a 1:1 stoichiometric ratio. Within the open porous crystalline structure ( ρ = 0.562 ), the periodically distributed windmill-like pores (nine adsorption sites) could be easily identified. Their edges are decorated by adjacent small voids enclosing four and six free adsorption sites. The second brick wall-like polymorph with a parallelogram unit cell ( 3 × 4 ,   60 / 120 o ) is built of unidirectionally oriented voids including six adsorption sites each (see Figure 10b). Interestingly, such close-packed arrangement of relatively small voids makes this metal–organic superstructure the densest among all 2D crystals discussed in this work ( ρ = 0.577 ) .

4. Conclusions

Exploiting the MC simulations, we examined the metal-directed self-assembly of flat π -aromatic molecules AJ with a reduced symmetry. The obtained results indicate that the dichotomous self-assembly is a unique feature of tritopic molecules A and J, in which terminal segments are functionalized selectively in para/(S)-meta positions. For the positional isomers with a more complex functionalization pattern (linkers B, C, D, H, and I), we observed the formation of metal–organic networks with large gear-like shaped voids decorated by smaller cavities, differing in size and shape. Interestingly, only in the case of the simulated system comprising linkers E, did we observe the formation of isolated metal–organic ladders instead of an open porous phase. However, the most intriguing supramolecular patterns emerged in the adsorbed systems comprising ligands F and G mixed with trivalent metal atoms. Namely, these simple building blocks were able to create complex mosaics with an aperiodic distribution of bridging linkers and a highly periodic arrangement of metal atoms.
To conclude, the performed calculations indicate that a subtle modification of intramolecular distribution of active centers embedded in termini of Y-shaped molecules AJ strongly affects the final topology of self-assembled overlayers, where anisotropic coordination bonds play a decisive role. In this context, the tritopic positional isomers investigated herein could be considered valuable building blocks of novel surface-confined supramolecular architectures.

Author Contributions

Conceptualization, D.N.; Data curation, K.N.; Funding acquisition, D.N.; Methodology, D.N.; Project administration, D.N.; Software, K.N.; Supervision, D.N.; Visualization, K.N.; Writing—original draft, D.N.; Writing—review and editing, K.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Centre, Poland, research grant 2018/31/D/ST4/01443, SONATA 14.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grünbaum, B.; Shephard, G.C. Tilings and Patterns, 2nd ed.; Dover Publications: Mineola, NY, USA, 2015. [Google Scholar]
  2. Schlickum, U.; Klappenberger, F.; Decker, R.; Zoppellaro, G.; Klyatskaya, S.; Ruben, M.; Kern, K.; Brune, H.; Barth, J.V. Surface-confined metal-organic nanostructures from Co-directed assembly of linear terphenyl-dicarbonitrile linkers on Ag(111). J. Phys. Chem. C 2010, 114, 15602–15606. [Google Scholar] [CrossRef] [Green Version]
  3. Adhikari, R.; Kuliga, J.; Ruppel, M.; Jux, N.; Marbach, H.; Steinrück, H.P. Self-assembled 2D-coordination Kagome, quadratic, and close-packed hexagonal lattices formed from a cyano-functionalized benzoporphyrin on Cu(111). J. Phys. Chem. C 2021, 125, 7204–7212. [Google Scholar] [CrossRef]
  4. Song, Y.; Wang, Y.; Jin, Q.; Zhou, K.; Shi, Z.; Liu, P.N.; Ma, Y. Self-assembly and local manipulation of Au-pyridyl coordination networks on metal surfaces. ChemPhysChem 2017, 18, 2088–2093. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, J.; Li, J.; Xu, Z.; Zhou, X.; Xue, Q.; Wu, T.; Zhong, M.; Li, R.; Sun, R.; Shen, Z.; et al. On-surface preparation of coordinated lanthanide-transition-metal clusters. Nat. Commun. 2021, 12, 1619. [Google Scholar] [CrossRef]
  6. Song, L.; Yang, B.; Liu, F.; Niu, K.; Han, Y.; Wang, J.; Zheng, Y.; Zhang, H.; Li, Q.; Chi, L. Synthesis of two-dimensional metal-organic frameworks via dehydrogenation reactions on a Cu(111) surface. J. Phys. Chem. C 2020, 124, 12390–12396. [Google Scholar] [CrossRef]
  7. Schlickum, U.; Decker, R.; Klappenberger, F.; Zoppellaro, G.; Klyatskaya, S.; Ruben, M.; Silanes, I.; Arnau, A.; Kern, K.; Brune, H.; et al. Metal-organic honeycomb nanomeshes with tunable Cavity size. Nano Lett. 2007, 7, 3813–3817. [Google Scholar] [CrossRef] [Green Version]
  8. Zhang, Y.; Zhang, X.; Li, Y.; Zhao, S.; Hou, S.; Wu, K.; Wang, Y. Packing Sierpiński triangles into two-dimensional crystals. J. Am. Chem. Soc. 2020, 142, 17928–17932. [Google Scholar] [CrossRef]
  9. Li, S.W.; Zhang, R.X.; Kang, L.X.; Li, D.Y.; Xie, Y.L.; Wang, C.X.; Liu, P.N. Steering metal-organic network structures through conformations and configurations on surfaces. ACS Nano 2021, 15, 18014–18022. [Google Scholar] [CrossRef]
  10. Bischoff, F.; He, Y.; Seufert, K.; Stassen, D.; Bonifazi, D.; Barth, J.V.; Auwärter, W. Tailoring large pores of porphyrin networks on Ag(111) by metal-organic coordination. Chem. Eur. J. 2016, 22, 15298–15306. [Google Scholar] [CrossRef] [Green Version]
  11. Dong, L.; Gao, Z.; Lin, N. Self-assembly of metal-organic coordination structures on surfaces. Progr. Surf. Sci. 2016, 91, 101–135. [Google Scholar] [CrossRef]
  12. Wakayama, Y. On-surface molecular nanoarchitectonics: From self-assembly to directed assembly. Jpn. J. Appl. Phys. 2016, 55, 1102AA. [Google Scholar] [CrossRef]
  13. Zhang, R.; Lyu, G.; Chen, C.; Lin, T.; Liu, J.; Liu, P.N.; Lin, N. Two-dimensional superlattices of Bi nanoclusters formed on a Au(111) surface using porous supramolecular templates. ACS Nano 2015, 9, 8547–8553. [Google Scholar] [CrossRef] [PubMed]
  14. Gutzler, R.; Stepanow, S.; Grumelli, D.; Lingenfelder, M.; Kern, K. Mimicking enzymatic active sites on surfaces for energy conversion chemistry. ACC Chem. Res. 2015, 48, 2132–2139. [Google Scholar] [CrossRef]
  15. Li, C.; Zhang, X.; Li, N.; Wang, Y.; Yang, J.; Gu, G.; Zhang, Y.; Hou, S.; Peng, L.; Wu, K.; et al. Construction of Sierpiński triangles up to the fifth order. J. Am. Chem. Soc. 2017, 139, 13749–13753. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, J.; Lin, T.; Shi, Z.; Xia, F.; Dong, L.; Liu, P.N.; Lin, N. Structural transformation of two-dimensional metal-organic coordination networks driven by intrinsic in-plane compression. J. Am. Chem. Soc. 2011, 133, 18760–18766. [Google Scholar] [CrossRef]
  17. Kumar, A.; Banerjee, K.; Foster, A.S.; Liljeroth, P. Two-dimensional band structure in honeycomb metal-organic frameworks. Nano Lett. 2018, 18, 5596–5602. [Google Scholar] [CrossRef]
  18. Umbach, T.R.; Bernien, M.; Hermanns, C.F.; Krüger, A.; Sessi, V.; Fernandez-Torrente, I.; Stoll, P.; Pascual, J.I.; Franke, K.J.; Kuch, W. Ferromagnetic coupling of mononuclear Fe centers in a self-assembled metal-organic network on Au(111). Phys. Rev. Lett. 2012, 109, 267207. [Google Scholar] [CrossRef] [Green Version]
  19. Shi, Z.; Liu, J.; Lin, T.; Xia, F.; Liu, P.N.; Lin, N. Thermodynamics and selectivity of two-dimensional metallo-supramolecular self-assembly resolved at molecular scale. J. Am. Chem. Soc. 2011, 133, 6150–6153. [Google Scholar] [CrossRef]
  20. Vijayaraghavan, S.; Ecija, D.; Auwärter, W.; Joshi, S.; Seufert, K.; Drach, M.; Nieckarz, D.; Szabelski, P.; Aurisicchio, C.; Bonifazi, D.; et al. Supramolecular assembly of interfacial nanoporous networks with simultaneous expression of metal-organic and organic-bonding motifs. Chem. Eur. J. 2013, 19, 14143–14150. [Google Scholar] [CrossRef]
  21. Ecija, D.; Vijayaraghavan, S.; Auwärter, W.; Joshi, S.; Seufert, K.; Aurisicchio, C.; Bonifazi, D.; Barth, J.V. Two-dimensional short-range disordered crystalline networks from flexible molecular modules. ACS Nano 2012, 6, 4258–4265. [Google Scholar] [CrossRef]
  22. Shu, C.H.; He, Y.; Zhang, R.X.; Chen, J.L.; Wang, A.; Liu, P.N. Atomic-scale visualization of stepwise growth mechanism of metal-alkynyl networks on surfaces. J. Am. Chem. Soc. 2020, 142, 16579–16586. [Google Scholar] [CrossRef] [PubMed]
  23. Yan, L.; Kuang, G.; Zhang, Q.; Shang, X.; Liu, P.N.; Lin, N. Self-assembly of a binodal metal-organic framework exhibiting a demi-regular lattice. Faraday Discuss. 2017, 204, 111–121. [Google Scholar] [CrossRef] [PubMed]
  24. Nijs, T.; Klein, Y.M.; Mousavi, S.F.; Ahsan, A.; Nowakowska, S.; Constable, E.C.; Housecroft, C.E.; Jung, T.A. The different faces of 4′-pyrimidinyl functionalized 4,2′:6′,4′′-terpyridines: Metal organic assemblies from solution and on Au(111) and Cu(111) surface platforms. J. Am. Chem. Soc. 2018, 140, 2933–2939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ammon, M.; Sander, T.; Maier, S. On-surface synthesis of porous carbon nanoribbons from polymer chains. J. Am. Chem. Soc. 2017, 139, 12976–12984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ammon, M.; Haller, M.; Sorayya, S.; Maier, S. On-surface synthesis of porous carbon nanoribbons on silver: Reaction kinetics and the influence of the surface structure. ChemPhysChem 2019, 20, 2333–2339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Šimėnas, M.; Ibenskas, A.; Tornau, E.E. Coronene molecules in hexagonal pores of tricarboxylic acids: A Monte Carlo study. J. Phys. Chem. C 2015, 119, 20524–20534. [Google Scholar] [CrossRef]
  28. Barnard, R.A.; Dutta, A.; Schnobrich, J.K.; Morrison, C.N.; Ahn, S.; Matzger, A.J. Two-dimensional crystals from reduced symmetry analogues trimesic acid. Chem. Eur. J. 2015, 21, 5954–5961. [Google Scholar] [CrossRef] [Green Version]
  29. Szabelski, P.; Nieckarz, D.; Rżysko, W. Influence of molecular shape and interaction anisotropy on the self-assembly of tripod building blocks on solid surfaces. Colloids Surf. A 2017, 532, 522–529. [Google Scholar] [CrossRef]
  30. Szabelski, P.; Nieckarz, D.; Rżysko, W. Structure formation in 2D assemblies comprising functional tripod molecules with reduced symmetry. J. Phys. Chem. C 2017, 121, 25104–25117. [Google Scholar] [CrossRef]
  31. Ivanova, D.; Deneva, V.; Nedeltcheva, D.; Kamounah, F.S.; Gergov, G.; Hansen, P.E.; Kawauchi, S.; Antonov, L. Tautomeric transformations of piroxicam in solution: A combined experimental and theoretical study. RSC Adv. 2015, 5, 31852–31860. [Google Scholar] [CrossRef] [Green Version]
  32. Gdula, K.; Nieckarz, D. On-surface self-assembly of metal-organic architectures: Insights from computer simulations. J. Phys. Chem. C 2020, 124, 20066–20078. [Google Scholar] [CrossRef]
  33. Nieckarz, K.; Szabelski, P.; Nieckarz, D. Monte Carlo simulations of the self-assembly of hierarchically organized metal-organic networks on solid surfaces. Surf. Sci. 2022, 719, 122041. [Google Scholar] [CrossRef]
  34. Fadeeva, A.; Gorbunov, V.A.; Solovyeva, O.S.; Stishenko, P.V.; Myshlyavtsev, A.V. Homologous series of flower phases in metal-organic networks in Au(111) surface. J. Phys. Chem. C 2020, 124, 11506–11515. [Google Scholar] [CrossRef]
  35. Ibenskas, A.; Tornau, E.E. Modeling of high-temperature ordered structures with weak intermolecular C-H···F and C-H···N bonds. J. Phys. Chem. C 2021, 125, 19560–19569. [Google Scholar] [CrossRef]
  36. Nieckarz, D.; Szabelski, P. Simulation of the self-assembly of simple molecular bricks into Sierpiński triangle. Chem. Comm. 2014, 50, 6843–6845. [Google Scholar] [CrossRef]
  37. Lisiecki, J.; Szabelski, P. Surface-confined metal-organic precursors comprising naphthalene-like derivatives with differently distributed halogen substituents: A Monte Carlo model. J. Phys. Chem. C 2020, 124, 20280–20293. [Google Scholar] [CrossRef]
  38. Metropolis, N.; Rosenbluth, A.W.; Rosenbluth, M.N.; Teller, A.H. Equation of state calculations by fast computing machines. J. Chem. Phys. 1953, 21, 1087–1092. [Google Scholar] [CrossRef] [Green Version]
  39. Frenkel, D.; Smit, B. Understanding Molecular Simulation; Academic Press: London, UK, 2002. [Google Scholar]
  40. Lischka, M.; Michelitsch, G.S.; Martsinovich, N.; Eichhorn, J.; Rastgoo-Lahrood, A.; Strunskus, T.; Breuer, R.; Reuter, K.; Schmittel, M.; Lackinger, M. Remote functionalization in surface-assisted dehalogenation by conformational mechanics: Organometallic self-assembly of 3,3′,5,5′-tetrabromo-2,2′,4,4′,6,6′-hexafluorobiphenyl on Ag(111). Nanoscale 2018, 10, 12035–12044. [Google Scholar] [CrossRef] [Green Version]
  41. Gutzler, R.; Cardenas, L.; Lipton-Duffin, J.; El Garah, M.; Dinca, L.E.; Szakacs, C.E.; Fu, C.; Gallagher, M.; Vondráček, M.; Rybachuk, M.; et al. Ullmann-type coupling of brominated tetrathienoanthracene on copper and silver. Nanoscale 2014, 6, 2660–2668. [Google Scholar] [CrossRef] [Green Version]
  42. Nieckarz, D.; Rżysko, W.; Szabelski, P. On-surface self-assembly of tetratopic molecular building blocks. Phys. Chem. Chem. Phys. 2018, 20, 23363–23377. [Google Scholar] [CrossRef]
Figure 1. A survey of Y-shaped positional isomers AJ and their modeled counterparts adsorbed on a triangular lattice. The black arrows attached to the terminal segments of the simulated molecules denote the positions of active centers (functional groups of real organic molecules: G = nitrile/propiolonitrile moiety, ring-nitrogen atom), able to interact reversibly with the coadsorbed trivalent metal atom (red, isolated segment). For clarity, in the case of prochiral units (BF and HJ) only one surface isomer, denoted arbitrary as (S), is shown. The vertices of a triangular lattice represent isolated adsorption sites present in an energetically homogenous metallic surface with (111) geometry.
Figure 1. A survey of Y-shaped positional isomers AJ and their modeled counterparts adsorbed on a triangular lattice. The black arrows attached to the terminal segments of the simulated molecules denote the positions of active centers (functional groups of real organic molecules: G = nitrile/propiolonitrile moiety, ring-nitrogen atom), able to interact reversibly with the coadsorbed trivalent metal atom (red, isolated segment). For clarity, in the case of prochiral units (BF and HJ) only one surface isomer, denoted arbitrary as (S), is shown. The vertices of a triangular lattice represent isolated adsorption sites present in an energetically homogenous metallic surface with (111) geometry.
Crystals 12 00492 g001
Figure 2. (a) The open porous metal–organic network composed of 600 metal atoms (red isolated segments) and 600 achiral ligands A in optimal 1:1 stoichiometric ratio at surface coverage θ   ~   0.21 , T = 0.001 ; (b) homogenous brick wall-like overlayer comprising 600 metal atoms and 600 molecules A at surface coverage θ   ~   0.21 , T = 0.001 . In the magnified fragments of the corresponding phases, the rhombic unit cells are marked (solid red lines). The solid black lines mark the edges of empty pores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. The vertices of a triangular lattice (adsorption sites) are not shown for the sake of clarity.
Figure 2. (a) The open porous metal–organic network composed of 600 metal atoms (red isolated segments) and 600 achiral ligands A in optimal 1:1 stoichiometric ratio at surface coverage θ   ~   0.21 , T = 0.001 ; (b) homogenous brick wall-like overlayer comprising 600 metal atoms and 600 molecules A at surface coverage θ   ~   0.21 , T = 0.001 . In the magnified fragments of the corresponding phases, the rhombic unit cells are marked (solid red lines). The solid black lines mark the edges of empty pores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. The vertices of a triangular lattice (adsorption sites) are not shown for the sake of clarity.
Crystals 12 00492 g002
Figure 3. (a) Multiporous metal–organic network composed of 600 metal atoms and 600 ligands B in a 1:1 stoichiometric ratio; (b) openwork phase comprising 600 metal atoms and 600 ligands C. In the magnified fragments of the metal–organic patterns, rhombic unit cells are marked (solid red lines). The solid black lines mark the edges of nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. For the largest gear-like shaped nanopores, their diameters d (corner-to-corner) are marked.
Figure 3. (a) Multiporous metal–organic network composed of 600 metal atoms and 600 ligands B in a 1:1 stoichiometric ratio; (b) openwork phase comprising 600 metal atoms and 600 ligands C. In the magnified fragments of the metal–organic patterns, rhombic unit cells are marked (solid red lines). The solid black lines mark the edges of nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. For the largest gear-like shaped nanopores, their diameters d (corner-to-corner) are marked.
Crystals 12 00492 g003
Figure 4. (a) Highly crystalline metal–organic phase composed of 600 metal atoms and 600 ligands D in a 1:1 stoichiometric ratio; (b) adsorbed overlayer comprising 600 metal atoms and 600 ligands E in a 1:1 stoichiometric ratio. In a magnified fragment of the open porous network, a rhombic unit cell is marked (solid red line). The solid black lines mark the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. The solid blue line denotes the zig-zag pattern created by metal atoms cementing the metal–organic ladder.
Figure 4. (a) Highly crystalline metal–organic phase composed of 600 metal atoms and 600 ligands D in a 1:1 stoichiometric ratio; (b) adsorbed overlayer comprising 600 metal atoms and 600 ligands E in a 1:1 stoichiometric ratio. In a magnified fragment of the open porous network, a rhombic unit cell is marked (solid red line). The solid black lines mark the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. The solid blue line denotes the zig-zag pattern created by metal atoms cementing the metal–organic ladder.
Crystals 12 00492 g004
Figure 5. (a) Exemplary aperiodic metal–organic network composed of 600 metal atoms and 600 ligands F in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.21 , T = 0.001 ; (b) adsorbed overlayer comprising 1200 metal atoms and 1200 ligands F in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.42 , T = 0.001 . The solid black lines in the bottom-left inset mark the edges of empty nanopores differing in size and shape. Inside the nanopores marked in different colors (pink, chalk, and light-green) the numbers of empty adsorption sites are given.
Figure 5. (a) Exemplary aperiodic metal–organic network composed of 600 metal atoms and 600 ligands F in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.21 , T = 0.001 ; (b) adsorbed overlayer comprising 1200 metal atoms and 1200 ligands F in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.42 , T = 0.001 . The solid black lines in the bottom-left inset mark the edges of empty nanopores differing in size and shape. Inside the nanopores marked in different colors (pink, chalk, and light-green) the numbers of empty adsorption sites are given.
Crystals 12 00492 g005
Figure 6. (a) The random distribution of nanopores within the metal–organic network built of 600 metal atoms and 600 ligands F. Nanopores differing in size and shape are marked in different colors (pink, chalk, and light-green). The solid black oval in the magnified fragment of the mosaic superstructure encompasses a hexagonal closed rosette with a periodic arrangement of metal atoms and linkers F. The incomplete solid black ovals denote the positions of open-walled rosettes; (b) supramolecular phase from Figure 5a, where some bridging ligands F were removed to better visualize a highly periodic distribution of metal atoms (red isolated segments). The solid red line indicates a primitive rhombic unit cell ( 13 × 13   60 / 120 o ).
Figure 6. (a) The random distribution of nanopores within the metal–organic network built of 600 metal atoms and 600 ligands F. Nanopores differing in size and shape are marked in different colors (pink, chalk, and light-green). The solid black oval in the magnified fragment of the mosaic superstructure encompasses a hexagonal closed rosette with a periodic arrangement of metal atoms and linkers F. The incomplete solid black ovals denote the positions of open-walled rosettes; (b) supramolecular phase from Figure 5a, where some bridging ligands F were removed to better visualize a highly periodic distribution of metal atoms (red isolated segments). The solid red line indicates a primitive rhombic unit cell ( 13 × 13   60 / 120 o ).
Crystals 12 00492 g006
Figure 7. Small fragments of artificially created metal–organic overlayers composed of ligands F and trivalent metal atoms. The solid red lines delimit unit cells of the crystalline supramolecular patterns. The solid black lines mark the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores the numbers of empty adsorption sites are given. (a,b) The flower-like networks with density ρ = 0.533 and rhombic unit cells differing in side lengths: ( 117 × 117   60 / 120 o   ) and ( 234 × 234   60 / 120 o ) . The black curved arrows denote a handedness of individual closed rosettes; (c) the multiporous phase with a rhombic unit cell ( 91 × 91 ,   60 / 120 o ) and density ρ = 0.457 ; (d) the homoporous phase with a rhombic unit cell ( 39 × 39 ,   60 / 120 o ) and density equal to ρ = 0.533 .
Figure 7. Small fragments of artificially created metal–organic overlayers composed of ligands F and trivalent metal atoms. The solid red lines delimit unit cells of the crystalline supramolecular patterns. The solid black lines mark the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores the numbers of empty adsorption sites are given. (a,b) The flower-like networks with density ρ = 0.533 and rhombic unit cells differing in side lengths: ( 117 × 117   60 / 120 o   ) and ( 234 × 234   60 / 120 o ) . The black curved arrows denote a handedness of individual closed rosettes; (c) the multiporous phase with a rhombic unit cell ( 91 × 91 ,   60 / 120 o ) and density ρ = 0.457 ; (d) the homoporous phase with a rhombic unit cell ( 39 × 39 ,   60 / 120 o ) and density equal to ρ = 0.533 .
Crystals 12 00492 g007
Figure 8. (a) The aperiodic metal–organic network comprising 1200 metal atoms and 1200 anchor-like ligands G in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.42 , T = 0.001 ; (b) adsorbed overlayer composed of 1500 metal atoms and 1500 molecules ligands G in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.53 , T = 0.001 . The solid black lines mark the edges of empty nanopores differing in size and shape.
Figure 8. (a) The aperiodic metal–organic network comprising 1200 metal atoms and 1200 anchor-like ligands G in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.42 , T = 0.001 ; (b) adsorbed overlayer composed of 1500 metal atoms and 1500 molecules ligands G in a 1:1 stoichiometric ratio at surface coverage θ   ~   0.53 , T = 0.001 . The solid black lines mark the edges of empty nanopores differing in size and shape.
Crystals 12 00492 g008
Figure 9. (a) A crystalline multiporous superstructure comprising 600 metal atoms and 600 ligands H in a 1:1 stoichiometric ratio; (b) openwork metal–organic phase composed of 600 metal atoms and 600 ligands I in a 1:1 stoichiometric ratio. In the magnified fragments of the ordered metal–organic arrays, the rhombic unit cells are marked (solid red lines). The solid black lines delimit the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. For the largest gear-like shaped nanopores, their diameters d (corner-to-corner) are marked.
Figure 9. (a) A crystalline multiporous superstructure comprising 600 metal atoms and 600 ligands H in a 1:1 stoichiometric ratio; (b) openwork metal–organic phase composed of 600 metal atoms and 600 ligands I in a 1:1 stoichiometric ratio. In the magnified fragments of the ordered metal–organic arrays, the rhombic unit cells are marked (solid red lines). The solid black lines delimit the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given. For the largest gear-like shaped nanopores, their diameters d (corner-to-corner) are marked.
Crystals 12 00492 g009
Figure 10. The polymorphic metal–organic superstructures obtained for molecule J: (a) a multiporous overlayer comprising 600 metal atoms and 600 ligands J in a 1:1 stoichiometric ratio; (b) a homogenous network composed of 600 metal atoms and 600 ligands J in a 1:1 stoichiometric ratio. In the magnified fragments of the 2D patterns, the rhombic unit cells are marked (solid red lines). The solid black lines delimit the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given.
Figure 10. The polymorphic metal–organic superstructures obtained for molecule J: (a) a multiporous overlayer comprising 600 metal atoms and 600 ligands J in a 1:1 stoichiometric ratio; (b) a homogenous network composed of 600 metal atoms and 600 ligands J in a 1:1 stoichiometric ratio. In the magnified fragments of the 2D patterns, the rhombic unit cells are marked (solid red lines). The solid black lines delimit the edges of empty nanopores differing in size and shape. Inside the highlighted nanopores, the numbers of empty adsorption sites are given.
Crystals 12 00492 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nieckarz, K.; Nieckarz, D. Monte Carlo Simulations of the Metal-Directed Self-Assembly of Y-Shaped Positional Isomers. Crystals 2022, 12, 492. https://doi.org/10.3390/cryst12040492

AMA Style

Nieckarz K, Nieckarz D. Monte Carlo Simulations of the Metal-Directed Self-Assembly of Y-Shaped Positional Isomers. Crystals. 2022; 12(4):492. https://doi.org/10.3390/cryst12040492

Chicago/Turabian Style

Nieckarz, Karolina, and Damian Nieckarz. 2022. "Monte Carlo Simulations of the Metal-Directed Self-Assembly of Y-Shaped Positional Isomers" Crystals 12, no. 4: 492. https://doi.org/10.3390/cryst12040492

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop