Next Article in Journal
A Novel Effect of CO2 Laser Induced Piezoelectricity in Ag2Ga2SiS6 Chalcogenide Crystals
Next Article in Special Issue
A High Laser Damage Threshold and a Good Second-Harmonic Generation Response in a New Infrared NLO Material: LiSm3SiS7
Previous Article in Journal
A Review of Transmission Electron Microscopy of Quasicrystals—How Are Atoms Arranged?
Previous Article in Special Issue
Properties of LiGa0.5In0.5Se2: A Quaternary Chalcogenide Crystal for Nonlinear Optical Applications in the Mid-IR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Packing of Helices: Is Chirality the Highest Crystallographic Symmetry?

by
Romain Gautier
1,2,* and
Kenneth R. Poeppelmeier
2
1
Institut des Matériaux Jean Rouxel (IMN), Université de Nantes, CNRS, 2 rue de la Houssinière, BP 32229, 44322 Nantes cedex 3, France
2
Department of Chemistry, Northwestern University, 2145 Sheridan Road, Evanston, IL 60201-3113, USA
*
Author to whom correspondence should be addressed.
Submission received: 31 July 2016 / Revised: 16 August 2016 / Accepted: 22 August 2016 / Published: 30 August 2016
(This article belongs to the Special Issue Nonlinear Optical Crystals)

Abstract

:
Chiral structures resulting from the packing of helices are common in biological and synthetic materials. Herein, we analyze the noncentrosymmetry (NCS) in such systems using crystallographic considerations. A comparison of the chiral structures built from helices shows that the chirality can be expected for specific building units such as 31/32 or 61/65 helices which, in hexagonal arrangement, will more likely lead to a chiral resolution. In these two systems, we show that the highest crystallographic symmetry (i.e., the symmetry which can describe the crystal structure from the smallest assymetric unit) is chiral. As an illustration, we present the synthesis of two materials ([Zn(2,2’-bpy)3](NbF6)2 and [Zn(2,2’-bpy)3](TaF6)2) in which the 3n helices pack into a chiral structure.

Graphical Abstract

1. Introduction

Acentric basic building units (BBUs) provide interesting blocks to target noncentrosymmetric (NCS) structures exhibiting potential physical properties, for example piezoelectricity, optical activity, or ferroelectricity. However, one must be able to predict the packing of these BBUs into the 3D acentric structure. Two packings are usually described in the literature: (i) the packing of chiral units of one-handedness into chiral materials and (ii) the alignment of polar units into polar materials [1,2,3,4,5,6,7,8]. As a next step towards the engineering of NCS, it is important to explore the general concepts which drive the BBUs to pack in one of these structural arrangements.
In the solid state, the components tend to crystallize in the highest possible crystallographic symmetry. Thus, about 92% of crystal structures contain only one crystallographically independent molecule or formula unit [9,10]. In most cases, the presence of an inversion center is required to describe the highest crystallographic symmetry. However, systems where the highest symmetry is reached only without an inversion center can also be investigated to understand and predict the noncentrosymmetric arrangements. In this context, the helical systems are of interest. Chiral structures built from helices are common in either biological or synthetic materials [11,12,13]. To describe the packing of helical proteins in biology, Chothia et al. proposed the “ridges and grooves” model in which the proteins tend to pack in the way that the ridges and grooves on the surface intercalate [14]. Using a similar model, Boncheva et al. described the self-assembly of millimeter-sized helical drill bits and bolts into 3D aggregates via capillary forces [15]. These authors showed that, in the case of a high degree of shape complementarity, the components tend to aggregate into chiral structures with enantiomerically pure helices. The shape compatibility depends on the periodicity, handedness and arrangement of the helices and is also a general concept which can be extended to analyze the NCS in the solid state. In organic and inorganic solid state, the interactions between helices are also symmetry-dependent. Thus, Meille et al. previously reported that helical polymers with a hexagonal arrangement have a high probability to crystallize into a chiral space group [16]. They reported that it is not possible to pack two helices of opposite chirality in the hexagonal arrangement and keep them symmetrically equivalent. In parallel to this work, we reported a similar result in hybrid materials. We previously showed that 31 helices in the P3121 space group (i.e., in hexagonal arrangement) tend to pack with helices of the same handedness because they share identical spatial and translational orientation. From the investigation of these compounds, we emphasized that the packing of helices with opposite handedness does not allow equivalent bridging of the helices. Thus, the packing of helices is also dependent on their symmetries but independent of the system (millimeter-sized helical drill bits, proteins, helical polymers, organic or inorganic structure) or the nature of interactions (capillary forces [15], close-packing of proteins [14,17], covalent [18], or entropic [16]).

2. Results and Discussion

In helical systems, one can determine the arrangements in which every helicate (the smallest repetitive unit of a helix) can crystallize with a similar environment and which are also symmetrically equivalent. In such conditions, the helices can share identical spatial and translational orientations. Symmetry patterns can be drawn by considering different configurations for 3n helices in hexagonal arrangement (Figure 1a), 6n helices in hexagonal arrangement (Figure 1b), and 4n helices in tetragonal arrangement (Figure 1c,d).
In a hexagonal arrangement of 3n or 6n helices, the helicates can have the same environment only if the helices are of the same-handednesses (i.e., in a chiral structure) (Figure 1a,b). For every other case in which one would consider the packing of helices of both-handednesses, the helicates would have different environments. Conversely, 4n helices in a tetragonal arrangement can either crystallize in chiral (Figure 1c) or non-chiral structure (Figure 1d) in order for the helicate to have the same environment.
This observation from the symmetry patterns could also be generalized from group theory. To reach the highest possible crystallographic symmetry in the solid state, the smallest repetitive unit of the helices should also be (if possible) the repetitive unit of the overall arrangement. In other words, the 3D arrangement should be described only from the helicates together with symmetry operations.
Thus, one can identify the space groups in which two conditions are satisfied: (i) the presence of screw axes along which the helix could be described in one direction from a unique repetitive unit, and (ii) the presence of symmetry operations to relate the screw axes to each other along the other two directions. If these conditions are satisfied, the helix can pack in this space group and reach the highest symmetry.
Figure 2 lists the different space groups which include the screw axes related by symmetry [19]. For example, 3n screw axes are only included in chiral space groups (Figure 2a). It results that the 3n helices in hexagonal arrangement reach the highest symmetry only when crystallizing in a chiral space group. In the case of the hexagonal arrangement of 6n helices (Figure 2b), the highest symmetry can also be only reached when crystallizing in a chiral space group. However, for the tetragonal arrangement of 4n helices (Figure 2c), the resulting structure can either be chiral or non-chiral to reach the highest crystallographic symmetry because 41 and 43 can be related by symmetry in this arrangement.
This analysis shows that the hexagonal arrangement of helices in the solid state differs from the tetragonal one. The highest symmetry of a hexagonal arrangement of either 3n or 6n helices is only reached when the structure is chiral (i.e., when a chiral resolution occurs), whereas the tetragonal arrangement of 4n helices can reach the highest symmetry when the structure is either centrosymmetric or noncentrosymmetric.
Two new compounds, [Zn(2,2’-bpy)3](MF6)2 (M = Nb5+ (I) and Ta5+ (II)), isostructural to the previously reported [A(2,2’-bpy)3](PF6)2 (A = Mn, Ru, Ni, Cu, Zn and Cd), allow us to validate this investigation (Experimental Details in Supplementary Information) [20,21,22]. Chiral metal trischelates [A(2,2’-bpy)3]2+ are described as a pair of enantiomers Δ and Λ-[M(ligand)3]n+ with the symmetry D3 (Figure 3). In the structures of I–II and previously reported isostructural compounds, these units are the helicates of helices propagating along the c axis. Thus, π-stacking occurs between the 2,2’-bipyridyne ligands of the helicates. Such interactions are common in organometallic or DNA helices [23,24,25]. The handedness of the helices is induced by the intrinsic character of helicates: the Δ enantiomer leads to the 31 helices while the Λ enantiomer leads to the 32 helices. [26]. Thus, two enantiomorphic crystal forms (P31 and P32) are induced by the chiral resolution of Δ and Λ-[Zn(2,2’-bipy)3]2+, respectively. Thus, in order to reach the highest symmetry, the 31/32 helices in hexagonal arrangement crystallize in the chiral space group P31/P32. The structure determination carried out on several crystals showed the presence of both configurations in equal amount.

3. Conclusions

In the solid state, most of the materials crystallize with the highest possible symmetry with a single chemical unit per asymmetric unit. The identification of systems in which the highest symmetry is polar or chiral could also be of great interest for the engineering of NCS materials. Thus, one could determine which characteristics a basic building unit must exhibit to target the NCS in the solid state. Our investigation of helical systems shows that 3n or 6n helices in hexagonal arrangement likely crystallize into a NCS agglomerate because the packing reaches the highest symmetry when the inversion center is excluded. Conversely, 4n helices can reach the highest symmetry in either a centro- or noncentrosymmetric tetragonal packing. This new insight on the relationships between NCS and highest symmetry could be extended to other helical systems and could improve our understanding and control of the engineering of chirality in the solid state.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4352/6/9/106/s1, X-Ray crystallographic file in CIF format for compounds III (CCDC 893952–893955), Experimental Details, and Table S1: Crystal data and summary of data collections, structure solutions and refinements for compounds [Zn(2,2’-bpy)3] (NbF6)2 (space-group P31) (I) and [Zn(2,2’-bpy)3](TaF6)2 (space-group P31) (II); Figure S1: View along c and b of the structure with the relative orientation of the distortion represented by the shortest Nb-F bond. Along a and b axes, the individual dipole moments are canceled by the 3n screw axis along c. The alignment of the individual polarizations is collinear to the polar axis of the structure.

Acknowledgments

This work was supported by funding from the National Science Foundation (Solid State Chemistry Award Nos. DMR-1307698 and DMR-1608218). The single crystal X-ray data were acquired at Northwestern University’s Integrated Molecular Structure Education and Research Center (IMSERC); we would like to thank Charlotte Stern and Amy Sarjeant for acquiring some data.

Author Contributions

Romain Gautier performed the experiments and wrote the paper; Kenneth R. Poeppelmeier supervised the work.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Pasteur, L. Sur les relations qui peuvent exister entre la forme criastalline, la composition chimique, et le sens de la polarization rotatoire. Ann. Chim. Phys. 1848, 24, 442–459. (In French) [Google Scholar]
  2. Pasteur, L. Recherches sur les propriétés spécifiques des deux acides qui composent l’acide racémique. impr. Bachelier. Ann. Chim. Phys. 1850, 28, 56–99. (In French) [Google Scholar]
  3. Flack, H.D. Chiral and Achiral Crystal Structures. Helv. Chim. Acta 2003, 86, 905–921. [Google Scholar]
  4. Donakowski, M.D.; Gautier, R.; Yeon, J.; Moore, D.T.; Nino, J.C.; Halasyamani, P.S.; Poeppelmeier, K.R. The Role of Polar, Lamdba (Λ)-Shaped Building Units in Noncentrosymmetric Inorganic Structures. J. Am. Chem. Soc. 2012, 134, 7679–7689. [Google Scholar]
  5. Gautier, R.; Poeppelmeier, K.R. Preservation of Chirality and Polarity between Chiral and Polar Building Units in the Solid State. Inorg. Chem. 2012, 51, 10613–10618. [Google Scholar]
  6. Maggard, P.A.; Kopf, A.L.; Stern, C.L.; Poeppelmeier, K.R. Probing helix formation in chains of vertex-linked octahedra. CrystEngComm 2004, 6, 452–457. [Google Scholar] [CrossRef]
  7. Gautier, R.; Norquist, A.J.; Poeppelmeier, K.R. From Racemic Units to Polar Materials. Cryst. Growth Des. 2012, 12, 6267–6271. [Google Scholar] [CrossRef]
  8. Gautier, R.; Klingsporn, J.M.; Van Duyne, R.P.; Poeppelmeier, K.R. Optical activity from racemates. Nat. Mater. 2016, 15, 591–592. [Google Scholar] [CrossRef] [PubMed]
  9. Padmaja, N.; Ramakumar, S.; Viswamitra, M.A. Space-group frequencies of proteins and of organic compounds with more than one formula unit in the asymmetric unit. Acta Crystallogr. A 1990, 46, 725–730. [Google Scholar] [CrossRef]
  10. Steed, J.W. Should solid-state molecular packing have to obey the rules of crystallographic symmetry? CrystEngComm 2003, 5, 169–179. [Google Scholar] [CrossRef]
  11. Sharma, V.; Crne, M.; Park, J.O.; Srinivasarao, M. Structural Origin of Circularly Polarized Iridescence in Jeweled Beetles. Science 2009, 325, 449–451. [Google Scholar] [CrossRef] [PubMed]
  12. Chung, W.-J.; Oh, J.-W.; Kwak, K.; Lee, B.Y.; Meyer, J.; Wang, E.; Hexemer, A.; Lee, S.-W. Biomimetic self-templating supramolecular structures. Nature 2011, 478, 364–368. [Google Scholar] [CrossRef] [PubMed]
  13. Meister, R.; Hallé, M.A.; Dumoulin, H.; Pieranski, P. Structure of the cholesteric focal conic domains at the free surface. Phys. Rev. E 1996, 54, 3771–3782. [Google Scholar] [CrossRef]
  14. Chothia, C.; Levitt, M.; Richardson, D. Helix to helix packing in proteins. J. Mol. Biol. 1981, 145, 215–250. [Google Scholar] [CrossRef]
  15. Boncheva, M.; Bruzewicz, D.A.; Whitesides, G.M. Formation of chiral, three-dimensional aggregates by self-assembly of helical components. Langmuir 2003, 19, 6066–6071. [Google Scholar] [CrossRef]
  16. Meille, S.V.; Allegra, G. Chiral crystallization of helical polymers. Macromolecules 1995, 28, 7764–7769. [Google Scholar] [CrossRef]
  17. Yin, P.; Zhang, Z.; Lv, H.; Li, T.; Haso, F.; Hu, L.; Zhang, B.; Bacsa, J.; Wei, Y.; Gao, Y.; et al. Chiral recognition and selection during the self-assembly process of protein-mimic macroanions. Nat. Commun. 2015, 6. [Google Scholar] [CrossRef] [PubMed]
  18. Maggard, P.A.; Stern, C.L.; Poeppelmeier, K.R. Understanding the Role of Helical Chains in the Formation of Noncentrosymmetric Solids. J. Am. Chem. Soc. 2001, 123, 7742–7743. [Google Scholar] [CrossRef] [PubMed]
  19. Hahn, T. Int. Tables Crystallogr; D. Reidel: Dordrecht, The Netherlands, 1983. [Google Scholar]
  20. Breu, J.; Domel, H.; Stoll, A. Racemic Compound Formation versus Conglomerate Formation with [M (bpy) 3](PF6) 2 (M = Ni, Zn, Ru); Molecular and Crystal Structures. Eur. J. Inorg. Chem. 2000, 2000, 2401–2408. [Google Scholar] [CrossRef]
  21. Wang, L.; Yang, X.-Y.; Huang, W. Tris (2, 2′-bipyridyl-κ2N, N′) copper (II) hexafluoridophosphate. Acta Crystallogr. Sect. E 2007, 63, m835–m836. [Google Scholar] [CrossRef]
  22. Kundu, N.; Mandal, D.; Chaudhury, M.; Tiekink, E.R.T. Luminescence characteristics and X-ray crystal structure of [Cd (bipy) 3][PF6] 2 (bipy = 2, 2′-bipyridine). Appl. Organomet. Chem. 2005, 19, 1268–1270. [Google Scholar] [CrossRef]
  23. He, C.; Zhao, Y.; Guo, D.; Lin, Z.; Duan, C. Chirality transfer through helical motifs in coordination compounds. Eur. J. Inorg. Chem. 2007, 2007, 3451–3463. [Google Scholar] [CrossRef]
  24. Ghazzali, M.; Langer, V.; Öhrström, L. The role of intermolecular interactions in the assemblies of FeII and CoII tetrakis-isothiocyanatometalates with tris(1,10-phenanthroline)-RuII: Crystal structures of two dual-metal assemblies featuring octahedral cationic and tetrahedral anionic modules. J. Solid State Chem. 2008, 181, 2191–2198. [Google Scholar] [CrossRef]
  25. Matta, C.F.; Castillo, N.; Boyd, R.J. Extended Weak Bonding Interactions in DNA:  π-Stacking (Base−Base), Base−Backbone, and Backbone−Backbone Interactions. J. Phys. Chem. B 2005, 110, 563–578. [Google Scholar] [CrossRef] [PubMed]
  26. Zheng, X.-D.; Jiang, L.; Feng, X.-L.; Lu, T.-B. Constructions of 1D helical chains with left-handed/right-handed helicity: A correlation between the helicity of 1D chains and the chirality of building blocks. Dalton Trans. 2009, 6802–6808. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Representation of examples of packings for (a) a chiral assembly of 3n helices in hexagonal arrangement, (b) a chiral assembly of 6n helices in hexagonal arrangement, (c) a chiral assembly of 4n helices in tetragonal arrangement, (d) a non-chiral assembly of 4n helices in tetragonal arrangement. The red arrows and black triangles represent the handedness of helices and the helicates, respectively. In order to be symmetrically equivalent, the helicates must have the same environment.
Figure 1. Representation of examples of packings for (a) a chiral assembly of 3n helices in hexagonal arrangement, (b) a chiral assembly of 6n helices in hexagonal arrangement, (c) a chiral assembly of 4n helices in tetragonal arrangement, (d) a non-chiral assembly of 4n helices in tetragonal arrangement. The red arrows and black triangles represent the handedness of helices and the helicates, respectively. In order to be symmetrically equivalent, the helicates must have the same environment.
Crystals 06 00106 g001
Figure 2. List of space groups exhibiting (a) 3n screw axes in hexagonal arrangement, (b) 6n screw axes in hexagonal arrangement and, (c) 4n screw axes in tetragonal arrangement. Space groups in red color are chiral. Space groups in bold are polar. No space groups combining 31 and 32 screw axes or 61 and 65 screw axes exist.
Figure 2. List of space groups exhibiting (a) 3n screw axes in hexagonal arrangement, (b) 6n screw axes in hexagonal arrangement and, (c) 4n screw axes in tetragonal arrangement. Space groups in red color are chiral. Space groups in bold are polar. No space groups combining 31 and 32 screw axes or 61 and 65 screw axes exist.
Crystals 06 00106 g002
Figure 3. View of (a) the two enantiomers Δ and Λ-[Zn(2,2’-bipy)3]2+, along the three-fold rotation axes, and (b) the two enantiomorphic 31 and 32 helices built from Δ and Λ-[Zn(2,2’-bipy)3]2+, respectively; π-stacking occurs between [Zn(2,2’-bipy)3]2+ helicates.
Figure 3. View of (a) the two enantiomers Δ and Λ-[Zn(2,2’-bipy)3]2+, along the three-fold rotation axes, and (b) the two enantiomorphic 31 and 32 helices built from Δ and Λ-[Zn(2,2’-bipy)3]2+, respectively; π-stacking occurs between [Zn(2,2’-bipy)3]2+ helicates.
Crystals 06 00106 g003

Share and Cite

MDPI and ACS Style

Gautier, R.; Poeppelmeier, K.R. Packing of Helices: Is Chirality the Highest Crystallographic Symmetry? Crystals 2016, 6, 106. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst6090106

AMA Style

Gautier R, Poeppelmeier KR. Packing of Helices: Is Chirality the Highest Crystallographic Symmetry? Crystals. 2016; 6(9):106. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst6090106

Chicago/Turabian Style

Gautier, Romain, and Kenneth R. Poeppelmeier. 2016. "Packing of Helices: Is Chirality the Highest Crystallographic Symmetry?" Crystals 6, no. 9: 106. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst6090106

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