Next Article in Journal
A Review on Progress, Challenges, and Prospects of Material Jetting of Copper and Tungsten
Previous Article in Journal
Construction of an Amethyst-like MoS2@Ni9S8/Co3S4 Rod Electrocatalyst for Overall Water Splitting
Previous Article in Special Issue
Hierarchical Self-Assembly and Conformation of Tb Double-Decker Molecular Magnets: Experiment and Molecular Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Dimensional Hetero- to Homochiral Phase Transition from Dynamic Adsorption of Barbituric Acid Derivatives

1
TITANS, SPEC, CEA, CNRS, Université Paris-Saclay, 91191 Gif sur Yvette, France
2
ITODYS, CNRS UMR 7086, Université Paris Cité, 15 rue Jean Antoine de Baïf, 75013 Paris, France
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(16), 2304; https://0-doi-org.brum.beds.ac.uk/10.3390/nano13162304
Submission received: 19 July 2023 / Revised: 8 August 2023 / Accepted: 8 August 2023 / Published: 10 August 2023

Abstract

:
Barbituric acid derivative (TDPT) is an achiral molecule, and its adsorption on a surface results in two opposite enantiomerically oriented motifs, namely TDPT-Sp and Rp. Two types of building blocks can be formed; block I is enantiomer-pure and is built up of the same motifs (format SpSp or RpRp) whereas block II is enantiomer-mixed and composes both motifs (format SpRp), respectively. The organization of the building blocks determines the formation of different nanoarchitectures which are investigated using scanning tunneling microscopy at a liquid/HOPG interface. Sophisticated, highly symmetric “nanowaves” are first formed from both building blocks I and II and are heterochiral. The “nanowaves” are metastable and evolve stepwisely into more close-packed “nanowires” which are formed from enantiomer-pure building block I and are homochiral. A dynamic hetero- to homochiral transformation and simultaneous multi-scale phase transitions are demonstrated at the single-molecule level. Our work provides novel insights into the control and the origin of chiral assemblies and chiral transitions, revealing the various roles of enantiomeric selection and chiral competition, driving forces, stability and molecular coverage.

1. Introduction

Chirality is a natural phenomenon which presents an essential scientific issue in fields ranging from the origin of life to enantioselective/specific catalysis in chemistry and from nonlinear optics to surface nanostructures, etc. Due to the rapidly expanding number of investigations of large molecules with various supramolecular organizations on solid surfaces, two-dimensional (2D) chirality [1] as a phenomenological concept has been increasingly observed and interpreted on the nanoscale. Chiral building blocks can be formed from chiral, prochiral or achiral molecules in the presence of surfaces [2,3,4,5], where chirality can be propagated into long-range nanostructures [6,7,8,9]. Organizational chirality in nanoarchitectures frequently involves 2D mirror symmetry and is produced through the self-assembly of organic molecules (chiral or achiral) or building blocks [6,7,8,9,10,11,12,13,14,15,16].
Chiral nanostructures can be directly visualized by means of Scanning Tunneling Microscopy (STM) under different environments in an ultrahigh vacuum (UHV) [4,12,15,16,17,18,19,20] or at a solid/liquid interface [3,5,21,22,23,24,25]. The main scientific and practical interest started with the simple recognition of chirality and then shifted towards the induction and engineering of more complex chiral assemblies [15,16,17,18,19,20,21,22,23,24,25,26,27].
In clean UHV environments, molecules form large enantiomer islands on metal surfaces, and metal–organic interface-directed chiral transfer occurs [28]. Controlled coverage [29,30,31,32] and temperature [33] have served as driving forces for chiral selectivity. Very recently, on-surface reaction has been revealed as a novel means to generate 2D chiral frameworks [34]. Rotation and dynamics of chirality-specific motors, though challenging, have been successfully recorded [35,36]. Although access to different metal surfaces or temperatures at the solid/liquid interface is limited, extra factors such as solvents, molecular concentration or pH can be exploited for chiral propagation and control. For example, chiral assemblies can be induced by mixing an achiral building block with a chiral modifier [21,22,23,37,38]. Solvents can be used to tune 2D chirality [39]. A “chiral memory concept” has been reported [40] where chirality remains even after the chiral modifier has been removed.
On the other hand, it is very important to understand the correlation between the different types of enantiomerically mixed or pure chiralities and chiral-selective phase transitions, but this remains very challenging [29,31,41] and will therefore be the focus of this work. We report a supramolecular self-assembly of prochiral barbituric acid derivatives on a highly oriented pyrolytic graphite (HOPG) substrate. Scanning tunneling microscopy images reveal that the molecules first self-assemble into sophisticated “nanowaves” at the liquid/HOPG interface where the highly symmetric structure is enantiomer-mixed and is heterochiral. The “nanowaves” are metastable and evolve stepwisely into more close-packed “nanowires” which are enantiomer-pure and homochiral. A hetero- to homochiral phase transition is demonstrated at the single-molecule level. The enantioselective competition induces a supramolecular reorganization, which demonstrates the occurrence of multi-scale phase transitions.

2. Materials and Methods

5-(4-Tetradecyloxybenzylidene)pyrimidine-2,4,6-trione (TDPT, formula C25H36N2O4, Figure 1a) was synthesized according to the known procedure described in previously published articles [42]. 4-Hydroxybenzaldehyde was first treated with NaH in DMF, and the phenolate obtained was condensed with 1-bromotetradecane to provide 1 in excellent yield (95%) and pure enough to be used directly in the next step. Reaction of 1 (Figure S1) with barbituric acid in the presence of a catalytic quantity of piperidinium benzoate gave 2 in good yield (64%).
TDPT was dissolved in 1-phenyloctane and deposited on a highly ordered pyrolytic graphite (HOPG) substrate. Dilute (10−4 to 10−3 M) and more concentrated (10−2 M) solutions were used for the deposition of “nanowaves” or “nanowires”. STM images of the samples at the liquid/solid interface were acquired on a SPM Nanoscope IV (Veeco, Bruker) scanning tunneling microscope. Cut Pt/Ir tips (Goodfellow) were used to obtain constant-current images at room temperature with a bias voltage applied to the sample. STM images were processed and analyzed by means of the FabViewer application [43].

3. Results

TDPT is an achiral molecule which does not contain any chiral elements. Its adsorption on a surface results in two opposite enantiomerically oriented motifs indicating that the molecule is prochiral [15] and that the surface plane serves as a 2D mirror (orange dashed line in Figure 1b). The two enantiomeric motifs are defined by treating the surface as a substituent to a tetrahedral center and then applying the Cahn–Ingold–Prelog priority rules [44]. The surface is considered to have the highest priority (3), and the hydrogen (of the CH bridging the two rings) is considered to have the lowest (0). The barbituric ring has priority (2), and the phenyl ring has priority (1). As shown in Scheme S1 (see Supplementary Materials), the configuration on the left can then be defined as “Sp” and that on the right as “Rp” where p means planar (Sp and Rp in red and green, respectively; Figure 2b).
Different building blocks can be formed from the two enantiomeric motifs through intermolecular hydrogen bonding. Building block I, composed of two identical enantiomers (SpSp or RpRp, Figure 1c, left), is homochiral. On the contrary, building block II, made up of two opposite enantiomers (SpRp, Figure 1c, right), is heterochiral. In the case of SpSp or RpRp building blocks, the same enantiomeric motifs interact with one strong N-H∙∙∙O and one weak C-H∙∙∙O hydrogen bond. In the SpRp blocks, the opposite motifs interact with two strong N-H∙∙∙O and one weak C-H∙∙∙O hydrogen bonds. The supramolecular organization from these different building blocks I and II can result in different nanoarchitectures and will be investigated under high-resolution STM at the solid/liquid interface.
The large-scale STM image reveals that TDPT molecules self-assemble into highly ordered 2D nanoarchitectures at the 1-phenyloctane/HOPG interface at room temperature. Dilute solutions with concentrations between 10−4 and 10−3 M are used for deposition. Very complex supramolecular wave-shaped nanopatterns are clearly visible (Figure 2a). Within the “nanowaves”, adjacent arcs curve in the same way. Two facing arcs from neighboring waves show almost perfect 2D mirror symmetry with respect to the surface plane. In the center of symmetry of two facing arcs (curving oppositely), nanosized cavities are formed where two guest TDPT molecules are nested in a guest–host system. The green circle (Figure 2b), superimposed on the STM image in Figure 2b, highlights the position of paired guest molecules in the host structure. The “head” of each TDPT molecule appears as two bright spots in the high-resolution STM image (Figure 2b) due to a higher electron density of states (DOS) whereas the alkyl chain appears darker. To form the double-line “nanowaves”, TDPT molecules are oriented in head-to-head fashion, perpendicular to the aligned axes of the wave patterns. The network unit cell (Figure 2b, red dashed lines) is a parallelogram with cell constants of 4.0 nm (a), 6.3 nm (b) and an angle of around 110°. The unit cell is surprisingly large, composed of 14 molecules, i.e., 12 molecules forming the host structure and 2 guest molecules. The alkyl chains of the TDPT molecules (guest and host molecules) are aligned in parallel to each other, which maximizes van de Waals (VDW) interactions.
A supramolecular model of the TDPT “nanowave” structure is proposed in Figure 2c. The host unit cell is constructed from 8 Sp and 4 Rp molecules (the unit cell can also contain 4 Sp and 8 Rp molecules in other domains to be enantiomerically equal), where TDPT-Sp and TDPT-Rp are the two motifs enantiomerically opposite (Figure 1b). Two guest molecules occupy each cavity, and they can be either SpSp or RpRp blocks. Moreover, the homochiral building block I (SpSp/RpRp, Figure 1c) and heterochiral building block II (SpRp, Figure 1c) both contribute with an enantiomeric Sp/Rp (or Rp/Sp) ratio of 2:1 inside one host unit cell (without considering the guest molecules). The “nanowave”, with its more symmetrical structure, is a 2D heterochiral nanoarchitecture. The STM images show that the smallest separation between the two neighboring “nanowaves” is around 2δ = 3.3 nm. This value is twice the length of the C14H29 alkyl chains (1.7 nm), which reveals that there is no chain intercrossing (Figure 2c) from neighboring “nanowaves”. The detailed symmetry analysis has been published elsewhere [14]. In this structure, molecules interact head-to-head by N-H∙∙∙O hydrogen bonds and side-to-side by O-H∙∙∙O and N-H∙∙∙O hydrogen bonds (model in Figure 2c). Intermolecular H-bonding and molecule/substrate VDW interactions stabilize the “nanowaves”.
The TDPT “nanowaves” are formed immediately after deposition, as previously demonstrated (Figure 2). A few hours after deposition, short nano-length wires (green dashed circles in Figure 3a,b) appear between the “nanowaves” at the initial guest positions. The nanowires are doubled (Figure 3) with molecules organized in head-to-head fashion. The short molecular wires continue to grow. A transition area is recorded (pink arrow in Figure 3a,b) where the host “nanowaves” are pushed apart and their original curved arc symmetry is lost.
A molecular model is proposed to interpret the initial formation of the nanowires (Figure 3c). In contrast to the “nanowaves” which are based on the alternation of Sp and Rp enantiomeric motifs with a ratio of 2:1 (composed of building blocks I and II), the short wire is formed from mono-enantiomeric SpSp or RpRp (building blocks I only). In a transition area between the “nanowave” and the “nanowire” (yellow dashed square in Figure 3c), Sp and Rp motifs alternate with a ratio of 1:1 (building block II only). The dynamic evolvement of the short nanowires causes local symmetry breaking within the initial “nanowave” organizations and is considered as an early stage of a structural as well as chiral phase transition, where the “nanowave” is heterochiral, and the nanowire is homochiral.
When the initial “nanowave” structure is left to evolve on the surface long enough (tens of hours), the short nanowires continue to deposit on the surface and turn into longer wires replacing the “nanowaves”. Eventually, the surface organization is overwhelmed by almost linear long wires. The high-resolution STM image (Figure 4b) reveals that molecules are arranged in head-to-head fashion, making the wires double, whereas the alkyl chains are in parallel perpendicular to the main axes of the wires. The unit cell is marked by the red rectangle with cell constants of around 5.6 (a) nm and 1.6 (b) nm and an angle close to 90°. The unit cell of the network is composed of 4 TDPT molecules. The same long nanowire structure can be directly obtained by deposition of a more concentrated TDPT solution (10−2 M). This indicates that wires are preferred at high molecular coverage.
The proposed model (Figure 4c) suggests that the double-wire structure is constructed mainly from mono-enantiomeric building block I (either SpSp or RpRp) making the structure homochiral. The same enantiomeric motifs are stabilized head-to-head and side-to-side through N-H∙∙∙O and C-H∙∙∙O hydrogen bonds. Occasionally, an enantiomer-mixed building block II is inserted into the homochiral wires as a defect which generates local bending of the wire (Figure 4c). It is possible that the different long-wire domains are built up from either SpSp or RpRp chiral building block I so that the numbers of the two enantiomeric motifs are the same on the whole surface. However, the long-wire structure, with an enantiomer-pure unit cell, is reasonably considered as homochiral. Different domains of the long wires coexist with orientation angles around 60° or 120° induced by the three-fold symmetry of the substrate (Figure 4a). The separation between the two neighboring TDPT wire is δ = 1.7 nm which corresponds to the length of the C14H29 alkyl chain. The alkyl chains from adjacent wires therefore intercalate, making the structure close-packed and closing any cavities.
Molecular dynamics on surfaces is an essential point in nano-engineering to control self-assemblies. The investigation or recording of such dynamics remain very challenging, depending on the time-scale match between the STM imaging and the on-surface evolvement. In our system, dynamic adsorption of the TDPT molecules can be visualized through a series of scans. A real-time movie (from continuous STM images) is recorded (Supplementary Materials file) providing an overview of homochiral TDPT wire deposition on a time-scale of around 20 min; Figure 5a–c show stills from the movie. We demonstrate that the TDPT wires continue to grow locally with time (indicated by the pink, blue and green circles in Figure 5a–c) at the liquid/HOPG interface until the surface is completely covered, and finally, there are no more cavities. As we have understood from the previous discussion that the alkyl chains of the TDPT molecules are aligned in parallel all along the HOPG [100] direction, the molecular wires therefore grow along the HOPG [1110] direction. Each TDPT molecule adsorbed is attached to its neighboring molecules side-to-side and head-to-head through hydrogen bonds and VDW interactions. Both intermolecular and molecule-substrate interactions obviously play important roles in the continuous molecular adsorption on the surface. Different domains are observed to have angles around 60° with respect to each other, which is induced by the HOPG substrate. Neither the STM scan nor the applied bias voltage influences or disturbs molecular wire growth. This phenomenon indicates that the close-packed, long wires are highly favored structures on-surface, and that the “nanowave” to “wire” structural phase transition is irreversible.

4. Discussion and Conclusions

Once adsorbed on a surface, both enantiomeric motifs are created from the prochiral TDPT molecules. The Sp and Rp enantiomeric motifs form hetero- and homochiral building blocks I and II, respectively. The homochiral and heterochiral interactions compete in the formation of the nanoarchitectures. “Nanowaves” and “nanowires” are generated, and a transition area between the two structures is attributed to different organizations of the two building blocks. Briefly, the “nanowave” (Figure 6a) is composed of both I and II building blocks; it is a metastable host structure (with nanocavities). The unit cell is enantiomer-mixed with an Sp/Rp (or Rp/Sp) ratio of 1:2. The “transition area” (Figure 6b) is made up of building block II (Sp/Rp ratio 1:1); it is unstable and appears only when the nanowave structural symmetry breaks due to growth of the initial nanowires. The nanowire (Figure 6c) consists of building block I only, and the unit cell, composed of purely Sp or Rp motifs, is an enantiomer-separated structure. The nanowave’s unit cell is heterochiral whereas the wire’s is homochiral.
The “nanowaves” and the “nanowires” are both stabilized through similar hydrogen bonds as well as molecule/substrate interactions. Moreover, in the nanowire structures, side-chains intercalate in parallel (Figure 6c) [45,46,47,48], and an extra VDW force is involved. This extra VDW between parallel chains is chiro-selective and exists only in homochiral organizations. The competition between hetero- and homochirality, as well as the different chiro-selective intermolecular forces, plays an essential role in the structural transitions. The homochiral “nanowires” are the more favored on-surface structures. Eventually, the two TDPT enantiomers should appear in equal numbers for the surface organizations to remain globally achiral.
To summarize, we observe a dynamic multiphase transition of TDPT molecules: first, a structural phase transition occurs where “nanowaves” are transformed into “nanowires”; second, this involves a hetero- to homochiral transition; third, during the transition process, chiro-selective intermolecular interactions play essential roles. The “nanowave” is stabilized by both molecular/substrate interactions and intermolecular hydrogen bonds. The “nanowires” are stabilized by similar forces, but an extra intermolecular VDW force from intercalated alkyl chains is involved, and this favors the homochiral structure. The irreversible dynamic deposition of wires is another proof of their higher stability. Moreover, the “nanowave” is a guest–host system, and the “nanowires” are more closely packed and have a higher surface coverage. Our work provides novel insights into the control and the origin of chiral assemblies and chiral transitions at the molecular scale, revealing the various roles of enantiomeric selection and chiral competition, driving forces, stability and molecular coverage.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano13162304/s1, Section S1: Synthesis chemistry; Section S2: Chirality definition; Scheme S1: Reagents and conditions: (a) NaH, in dry DMF, 0 °C; (b) nC14H29Br; (c) Barburic acid, cat, piperidinium benzoate, toluene, reflux; Section S3: STM results—Figure S1: STM images of (a) TDPT “nanowaves” and “nanowires” coexist. 180 × 180 nm2, (b) TDPT “nanowaves” 68 × 66 nm2, (c) TDPT “nanowaves” host-guest nanoarchitecture, 50 × 27 nm2 with four guests, two guest or no guest molecules. It = 20–60 pA, Us = −0.4 V~−0.5 V; Figure S2: STM images of TDPT “nanowave” and its heterochirality is indicated where the surface plane is considered as the 2D mirror; Figure S3: Continuous real-time STM images of TDPT wires deposition on HOPG surface in a 9 min time window. The red, blue and the green arrows indicate where the wires fill the surface cavities. The pink dashed circles reveal the movement of the wires. 150 × 75 nm2, It = 30 pA, Us = −0.4 V.

Author Contributions

Conceptualization, X.S.; design and performance of experiments, F.S. and X.S.; TDPT molecule synthesis, C.D.; STM data analysis, X.S.; writing original draft preparation, X.S.; writing—review and editing, X.S., F.S. and F.M.; molecular modeling, F.M.; funding acquisition, X.S. All authors have read and agreed to the published version of the manuscript.

Funding

The Agence Nationale de la Recherche (France ANR-20-CE09-0004, ANR-19-CE09 APMJ, ANR-10-LABX-0096).

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank J.S. Lomas for English editing and scientific discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barlow, S.M.; Raval, R. Complex organic molecules at metal surfaces: Bonding, organisation and chirality. Surf. Sci. Rep. 2003, 50, 201–341. [Google Scholar] [CrossRef]
  2. Suarez, M.; Branda, N.; Lehn, J.M. Supramolecular chirality: Chiral hydrogen-bonded supermolecules from achiral molecular Components. Helv. Chem. Acta 1998, 8, 1–13. [Google Scholar] [CrossRef]
  3. Elemans, J.A.A.W.; De Cat, I.; Xu, H.; De Feyter, S. Two-dimensional chirality at liquid–solid interfaces. Chem. Soc. Rev. 2009, 38, 722–736. [Google Scholar] [CrossRef] [PubMed]
  4. Ernst, K.H. Molecular chirality at surfaces. Phys. Status Solidi B 2012, 249, 2057–2088. [Google Scholar] [CrossRef]
  5. Chen, T.; Wang, D.; Wan, L.J. Two-dimensional chiral molecular assembly on solid surfaces: Formation and regulation. Natl. Sci. Rev. 2015, 2, 205–216. [Google Scholar] [CrossRef]
  6. Xu, H.; Wolffs, M.; Tomovic, Z.; Meijer, E.W.; Schenning, A.P.H.J.; De Feyter, S. A multivalent hexapod having stereogenic centers: Chirality and conformational dynamics in homochiral and heterochiral systems. Cryst. Eng. Comm. 2011, 13, 5584–5590. [Google Scholar] [CrossRef]
  7. Haq, S.; Liu, N.; Humblot, V.; Jansen, A.P.J.; Raval, R. Drastic symmetry breaking in supramolecular organization of enantiomerically unbalanced monolayers at surfaces. Nat. Chem. 2009, 1, 409–414. [Google Scholar] [CrossRef]
  8. Gutzler, R.; Ivasenko, O.; Fu, C.; Brusso, J.L.; Rosei, F.; Perepichka, D.F. Halogen bonds as stabilizing interactions in a chiral self-assembled molecular monolayer. Chem. Commun. 2011, 47, 9453–9455. [Google Scholar] [CrossRef]
  9. Weckesser, J.; De Vita, A.; Barth, J.V.; Cai, C.; Kern, K. Mesoscopic correlation of supramolecular chirality in one-dimensional hydrogen-bonded assemblies. Phys. Rev. Lett. 2001, 87, 096101. [Google Scholar] [CrossRef] [Green Version]
  10. Sleczkowski, P.; Katsonis, N.; Kapitanchuk, O.; Marchenko, A.; Mathevet, F.; Croset, B.; Lacaze, E. Emergence of chirality in hexagonally packed monolayers of hexapentyloxytriphenylene on Au(111): A joint experimental and theoretical study. Langmuir 2014, 30, 13275–13282. [Google Scholar] [CrossRef]
  11. Silly, F.; Ausset, S.; Sun, X. Coexisting chiral two-dimensional self-assembled structures of 1,2,3,4-tetrahydronaphthalene molecules: Porous pinwheel nanoarchitecture and close-packed herringbone arrangement. J. Phys. Chem. C 2017, 121, 15288–15293. [Google Scholar] [CrossRef] [Green Version]
  12. Sun, X.; Mura, M.; Jonkman, H.T.; Kantorovich, L.N.; Silly, F. Fabrication of a complex two-dimensional adenine–perylene-3, 4, 9, 10-tetracarboxylic dianhydride chiral nanoarchitecture through molecular self-assembly. J. Phys. Chem. C 2012, 116, 2493–2499. [Google Scholar] [CrossRef] [Green Version]
  13. Rajwar, D.; Sun, X.; Cho, S.J.; Grimsdale, A.C.; Fichou, D. Synthesis and 2D self-assembly at the liquid-solid interface of end-substituted star-shaped oligophenylenes. Cryst. Eng. Comm. 2012, 14, 5182–5187. [Google Scholar] [CrossRef]
  14. Sun, X.; Silly, F.; Maurel, F.; Dong, C. Supramolecular chiral host–guest nanoarchitecture induced by the selective assembly of barbituric acid derivative enantiomers. Nanotechnology 2016, 27, 42LT01. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Karan, S.; Wang, Y.; Robles, R.; Lorente, N.; Berndt, R. Surface-supported supramolecular pentamers. J. Am. Chem. Soc. 2013, 135, 14004–14007. [Google Scholar] [CrossRef] [PubMed]
  16. Fasel, R.; Parschau1, M.; Ernst, K.H. Amplification of chirality in two-dimensional enantiomorphous lattices. Nature 2006, 439, 449–452. [Google Scholar] [CrossRef]
  17. Karageorgakia, C.; Ernst, K.H. A metal surface with chiral memory. Chem. Commun. 2014, 50, 1814–1816. [Google Scholar] [CrossRef]
  18. Barth, J.V. Molecular Architectonic on Metal Surfaces. Annu. Rev. Phys. Chem. 2007, 58, 375–407. [Google Scholar] [CrossRef] [Green Version]
  19. Raval, R. Chiral expression from molecular assemblies at metal surfaces: Insights from surface science techniques. Chem. Soc. Rev. 2009, 38, 707–721. [Google Scholar] [CrossRef]
  20. Zhang, H.; Liu, H.; Shen, C.; Gan, F.; Su, X.; Qiu, H.; Yang, B.; Yu, P. Chiral recognition of hexahelicene on a surface via the forming of asymmetric heterochiral trimers. Int. J. Mol. Sci. 2019, 20, 2018. [Google Scholar] [CrossRef] [Green Version]
  21. Guo, Z.; De Cat, I.; Van Averbeke, B.; Lin, J.; Wang, G.; Xu, H.; Lazzaroni, R.; Beljonne, D.; Meijer, E.W.; Schenning, A.P.H.J.; et al. Nucleoside-assisted self-assembly of oligo(p-phenylenevinylene)s at liquid/solid interface: Chirality and nanostructures. J. Am. Chem. Soc. 2011, 133, 17764–17771. [Google Scholar] [CrossRef] [PubMed]
  22. Tahara, K.; Yamaga, H.; Ghijsens, E.; Inukai, K.; Adisoejoso, J.; Blunt, M.O.; De Feyter, S.; Tobe, Y. Control and induction of surface-confined homochiral porous molecular networks. Nat. Chem. 2011, 3, 714. [Google Scholar] [CrossRef]
  23. Chen, T.; Yang, W.H.; Wang, D.; Wan, L.J. Globally homochiral assembly of two-dimensional molecular networks triggered by co-absorbers. Nat. Chem. 2013, 4, 1389. [Google Scholar] [CrossRef] [Green Version]
  24. Sun, X.; Yao, X.; Lafolet, F.; Lemercier, G.; Lacroix, J.C. One-Dimensional Double Wires and Two-Dimensional Mobile Grids: Cobalt/Bipyridine Coordination Networks at the Solid/Liquid Interface. J. Phys. Chem. Lett. 2019, 10, 4164–4169. [Google Scholar] [CrossRef] [PubMed]
  25. Zheng, Y.; Yu, L.; Zou, Y.; Yang, Y.; Wang, C. Steric dependence of chirality effect in surface-mediated peptide assemblies identified with scanning tunneling microscopy. Nano Lett. 2019, 19, 5403–5409. [Google Scholar] [CrossRef] [PubMed]
  26. Rivera, J.; Craig, S.L.; Martin, T.; Rebek, J., Jr. Chiral guests and their ghosts in reversibly assembled hosts. Angew. Chem. Int. Ed. 2000, 39, 2130–2132. [Google Scholar] [CrossRef]
  27. Sun, X.; Frath, D.; Lafolet, F.; Lacroix, J.C. Supramolecular Networks and Wires Dominated by Intermolecular BiEDOT Interactions. J. Phys. Chem. C 2018, 122, 22760–22766. [Google Scholar] [CrossRef]
  28. Yang, B.; Cao, N.; Ju, H.; Lin, H.; Li, Y.; Ding, H.; Ding, J.; Zhang, J.; Peng, C.; Zhang, H.; et al. Intermediate States Directed Chiral Transfer on a Silver Surface. J. Am. Chem. Soc. 2019, 141, 168–174. [Google Scholar] [CrossRef]
  29. Vidal, F.; Delvigne, E.; Stepanow, S.; Lin, N.; Barth, J.V.; Kern, K. Chiral phase transition in two-dimensional supramolecular assemblies of prochiral molecules. J. Am. Chem. Soc. 2005, 127, 10101–10106. [Google Scholar] [CrossRef]
  30. Mugarza, A.; Lorente, N.; Ordejon, P.; Krull, C.; Stepanow, S.; Bocquet, M.L.; Fraxedas, J.; Ceballos, G.; Gambardella, P. Orbital specific chirality and homochiral self-assembly of achiral molecules induced by charge transfer and spontaneous symmetry breaking. Phys. Rev. Lett. 2010, 105, 115702. [Google Scholar] [CrossRef] [Green Version]
  31. Yang, B.; Wang, Y.; Cun, H.; Du, S.; Xu, M.; Wang, Y.; Ernst, K.H.; Gao, H.J. Direct observation of enantiospecific substitution in a two-dimensional chiral phase transition. J. Am. Chem. Soc. 2010, 132, 10440–10444. [Google Scholar] [CrossRef] [PubMed]
  32. Xiang, F.; Schneider, M.A. Coverage-induced chiral transition of Co(II)-5,15-diphenylporphyrin self-assemblies on Cu(111). J. Phys. Chem. C 2022, 126, 6745–6752. [Google Scholar] [CrossRef]
  33. Han, D.; Wang, T.; Huang, J.; Li, X.; Zeng, Z.; Zhu, J. Chiral nanoporous networks featuring various chiral vertices from an achiral molecule on Ag(100). Nano Res. 2022, 15, 3753–3762. [Google Scholar] [CrossRef]
  34. Zhang, H.; Gong, Z.; Sun, K.; Duan, R.; Ji, P.; Li, L.; Li, C.; Müllen, K.; Chi, L. Chirality transfer via on-surface reaction. J. Am. Chem. Soc. 2016, 138, 11743–11748. [Google Scholar] [CrossRef]
  35. Stolz, S.; Gröning, O.; Prinz, J.; Brune, H.; Widmer, R. Molecular motor crossing the frontier of classical to quantum tunneling motion. Proc. Natl. Acad. Sci. USA 2020, 117, 14838–14842. [Google Scholar] [CrossRef]
  36. Schied, M.; Prezzi, D.; Liu, D.; Kowarik, S.; Jacobson, P.A.; Corni, S.; Tour, J.M.; Grill, L. Chirality-specific unidirectional rotation of molecular motors on Cu(111). ACS Nano 2023, 17, 3958–3965. [Google Scholar] [CrossRef]
  37. De Cat, I.; Guo, Z.; George, S.J.; Meijer, E.W.; Schenning, A.P.H.J.; De Feyter, S. Induction of chirality in an achiral monolayer at the liquid/solid interface by a supramolecular chiral auxiliary. J. Am. Chem. Soc. 2012, 134, 3171–3177. [Google Scholar] [CrossRef]
  38. Cucinotta, A.; Kahlfuss, C.; Minoia, A.; Eyley, S.; Zwaenepoel, K.; Velpula, G.; Thielemans, W.; Lazzaroni, R.; Bulach, V.; Hosseini, M.W.; et al. Metal ion and guest-mediated spontaneous resolution and solvent-induced chiral Symmetry breaking in guanine-based metallosupramolecular networks. J. Am. Chem. Soc. 2023, 145, 1194–1205. [Google Scholar] [CrossRef] [PubMed]
  39. Yamagata, K.; Maeda, M.; Tessari, Z.; Mali, K.S.; Tobe, Y.; De Feyter, S.; Tahara, K. Solvent mediated nanoscale quasi-periodic chirality reversal in self-assembled molecular networks featuring mirror twin boundaries. Small 2023, 19, 2207209. [Google Scholar] [CrossRef]
  40. Prins, L.J.; De Jong, F.; Timmerman, P.; Reinhoudt, D.N. An enantiomerically pure hydrogen-bonded assembly. Nature 2000, 408, 181–184. [Google Scholar] [CrossRef] [PubMed]
  41. Iski, E.V.; Tierney, H.L.; Jewell, A.D.; Sykes, E.C.H. Spontaneous transmission of chirality through multiple length scales. Chem. Eur. J. 2011, 17, 7205–7212. [Google Scholar] [CrossRef]
  42. Suarez, M.; De Armas, M.; Ramirez, O.; Alvarez, A.; Martınez-Alvarez, R.; Molero, D.; Seoane, C.; Liz, R.; De Armas, H.N.; Blaton, N.M.; et al. Synthesis and structural study of new highly lipophilic 1,4-dihydropyridines. New J. Chem. 2005, 29, 1567–1576. [Google Scholar] [CrossRef]
  43. Silly, F. A robust method for processing scanning probe microscopy images and determining nanoobject position and dimensions. J. Microsc. 2009, 236, 211–218. [Google Scholar] [CrossRef] [PubMed]
  44. Moss, G.P. Basic Terminology of Stereochemistry. Pure Appl. Chem. 1996, 68, 2193–2222. [Google Scholar] [CrossRef]
  45. Dong, M.; Miao, K.; Hu, Y.; Wu, J.; Li, J.; Pang, P.; Miao, X.; Deng, W. Cooperating dipole–dipole and van der Waals interactions driven 2D self-assembly of fluorenone derivatives: Ester chain length effect. Phys. Chem. Chem. Phys. 2017, 19, 31113–31120. [Google Scholar] [CrossRef] [PubMed]
  46. Colle, R.; Grosso, G.; Ronzani, A.; Zicovich-Wilson, C.M. Structure and X-ray spectrum of crystalline poly(3-hexylthiophene) from DFT-van der Waals calculations. Phys. Status Solidi B 2011, 248, 1360–1368. [Google Scholar] [CrossRef]
  47. Ciesielski, A.; Haar, S.; Bényei, A.; Paragi, G.; Guerra, C.F.; Bickelhaupt, F.M.; Masiero, S.; Szolomájer, J.; Samorì, P.; Spada, G.P.; et al. Self-Assembly of N3-Substituted Xanthines in the Solid State and at the Solid–Liquid Interface. Langmuir 2013, 29, 7283–7290. [Google Scholar] [CrossRef]
  48. Pierro, A.D.; Saracco, G.; Fina, A. Molecular junctions for thermal transport between graphene nanoribbons: Covalent bonding vs. interdigitated chains. Comput. Mater. Sci. 2018, 142, 255–260. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Molecular models: (a) TDPT molecular structure where C, N, O and H elements are grey, dark blue, red and white balls, respectively. (b) Surface-supported TDPT-Sp and Rp enantiomeric motifs showing 2D mirror symmetry. (c) Building blocks I (homochiral) and II (heterochiral) which are made from the same (SpSp/RpRp) and opposite (SpRp) enantiomers, respectively. The green dashed lines indicate the possible formation of hydrogen bonds.
Figure 1. Molecular models: (a) TDPT molecular structure where C, N, O and H elements are grey, dark blue, red and white balls, respectively. (b) Surface-supported TDPT-Sp and Rp enantiomeric motifs showing 2D mirror symmetry. (c) Building blocks I (homochiral) and II (heterochiral) which are made from the same (SpSp/RpRp) and opposite (SpRp) enantiomers, respectively. The green dashed lines indicate the possible formation of hydrogen bonds.
Nanomaterials 13 02304 g001
Figure 2. (a,b) Large-scale (44 × 40 nm2) and high-resolution (20 × 14 nm2) STM images of double-line “nanowaves” which are also guest–host systems. It = 30~60 pA, Us = 0.3~0.5 V. (c) Molecular modeling shows the structural organization. Red and green circles indicate the different Sp or Rp enantiomeric motifs. Inset model shows that neighboring molecules are organized in mixed SpSp/RpRp and SpRp modes.
Figure 2. (a,b) Large-scale (44 × 40 nm2) and high-resolution (20 × 14 nm2) STM images of double-line “nanowaves” which are also guest–host systems. It = 30~60 pA, Us = 0.3~0.5 V. (c) Molecular modeling shows the structural organization. Red and green circles indicate the different Sp or Rp enantiomeric motifs. Inset model shows that neighboring molecules are organized in mixed SpSp/RpRp and SpRp modes.
Nanomaterials 13 02304 g002
Figure 3. (a,b) Large-scale (55 × 55 nm2) and high-resolution (21 × 16 nm2) STM images. “Nanowires” start to grow between the “nanowaves”. It = 30~60 pA, Us = −0.3~−0.5 V. (c) Molecular modeling shows the transition between the two structures. Red and green circles indicate the different S or R enantiomers. Yellow rectangle highlights the transition area corresponding to the pink arrow in image (b).
Figure 3. (a,b) Large-scale (55 × 55 nm2) and high-resolution (21 × 16 nm2) STM images. “Nanowires” start to grow between the “nanowaves”. It = 30~60 pA, Us = −0.3~−0.5 V. (c) Molecular modeling shows the transition between the two structures. Red and green circles indicate the different S or R enantiomers. Yellow rectangle highlights the transition area corresponding to the pink arrow in image (b).
Nanomaterials 13 02304 g003
Figure 4. STM images of TDPT molecular wires on HOPG. (a) 75 × 75 nm2, It = 30 pA, Us = −0.4 V. (b) 14 × 7 nm2, It = 40 pA, Us = −0.3 V. The red rectangle highlights the unit cell of the structure. (c) Molecular model of the double wires. The green and red circles indicate the Rp and Sp enantiomeric motifs, respectively. The purple dashed arrow shows the separation between the two wires (δ~1.7 nm).
Figure 4. STM images of TDPT molecular wires on HOPG. (a) 75 × 75 nm2, It = 30 pA, Us = −0.4 V. (b) 14 × 7 nm2, It = 40 pA, Us = −0.3 V. The red rectangle highlights the unit cell of the structure. (c) Molecular model of the double wires. The green and red circles indicate the Rp and Sp enantiomeric motifs, respectively. The purple dashed arrow shows the separation between the two wires (δ~1.7 nm).
Nanomaterials 13 02304 g004
Figure 5. Real-time STM stills of TDPT wires being continuously deposited on HOPG surface in a 21 min time window. The different pink, green and blue circles indicate where the wires fill the surface cavities: 150 × 150 nm2, It = 30 pA, Us = −0.4 V.
Figure 5. Real-time STM stills of TDPT wires being continuously deposited on HOPG surface in a 21 min time window. The different pink, green and blue circles indicate where the wires fill the surface cavities: 150 × 150 nm2, It = 30 pA, Us = −0.4 V.
Nanomaterials 13 02304 g005
Figure 6. Schematic illustration of: (a) Heterochiral nanowave, (b) Transition area, and (c) Homochiral wires.
Figure 6. Schematic illustration of: (a) Heterochiral nanowave, (b) Transition area, and (c) Homochiral wires.
Nanomaterials 13 02304 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Silly, F.; Dong, C.; Maurel, F.; Sun, X. Two-Dimensional Hetero- to Homochiral Phase Transition from Dynamic Adsorption of Barbituric Acid Derivatives. Nanomaterials 2023, 13, 2304. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13162304

AMA Style

Silly F, Dong C, Maurel F, Sun X. Two-Dimensional Hetero- to Homochiral Phase Transition from Dynamic Adsorption of Barbituric Acid Derivatives. Nanomaterials. 2023; 13(16):2304. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13162304

Chicago/Turabian Style

Silly, Fabien, Changzhi Dong, François Maurel, and Xiaonan Sun. 2023. "Two-Dimensional Hetero- to Homochiral Phase Transition from Dynamic Adsorption of Barbituric Acid Derivatives" Nanomaterials 13, no. 16: 2304. https://0-doi-org.brum.beds.ac.uk/10.3390/nano13162304

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