Next Article in Journal
Improved Sensitivity of Surface-Enhanced Raman Scattering with Gold Nanoparticles-Insulator-Metal Sandwich Layers on Flat Sapphire Substrate
Next Article in Special Issue
Low Resistance Asymmetric III-Nitride Tunnel Junctions Designed by Machine Learning
Previous Article in Journal
A Review on the Production Methods and Applications of Graphene-Based Materials
Previous Article in Special Issue
Investigation on Plastic Flow Behaviors of FCC Polycrystalline Aluminum under Pre-Cyclic Tension-Compression Loading: Experiments and Crystal Plasticity Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modeling of Self-Assembled Peptide Nanotubes and Determination of Their Chirality Sign Based on Dipole Moment Calculations

1
Institute of Mathematical Problems of Biology, The Branch of Keldysh Institute of Applied Mathematics, RAS, 142290 Pushchino, Russia
2
Faculty of Physics, Lomonosov Moscow State University, 119991 Moscow, Russia
3
School of Natural Sciences and Mathematics, Ural Federal University, 620000 Ekaterinburg, Russia
4
Department of Physics & CICECO-Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
5
Physical Materials Science and Composite Materials Centre, Research School of Chemistry & Applied Biomedical Sciences, National Research Tomsk Polytechnic University, 634050 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(9), 2415; https://doi.org/10.3390/nano11092415
Submission received: 31 July 2021 / Revised: 3 September 2021 / Accepted: 9 September 2021 / Published: 16 September 2021
(This article belongs to the Special Issue Simulation and Modeling of Nanomaterials)

Abstract

:
The chirality quantification is of great importance in structural biology, where the differences in proteins twisting can provide essentially different physiological effects. However, this aspect of the chirality is still poorly studied for helix-like supramolecular structures. In this work, a method for chirality quantification based on the calculation of scalar triple products of dipole moments is suggested. As a model structure, self-assembled nanotubes of diphenylalanine (FF) made of L- and D-enantiomers were considered. The dipole moments of FF molecules were calculated using semi-empirical quantum-chemical method PM3 and the Amber force field method. The obtained results do not depend on the used simulation and calculation method, and show that the D-FF nanotubes are twisted tighter than L-FF. Moreover, the type of chirality of the helix-like nanotube is opposite to that of the initial individual molecule that is in line with the chirality alternation rule general for different levels of hierarchical organization of molecular systems. The proposed method can be applied to study other helix-like supramolecular structures.

1. Introduction

Self-assembly of biomolecules such as amino acids, nucleotide bases, phospholipids, and oligo- and polypeptides is the basis for the formation of DNA, molecular motors, viruses, and many other biological systems [1,2,3,4]. Biomimetic self-assembly is also a promising bottom-up approach for nanomaterials design in nanobiotechnology [5,6]. Helical self-organizing structures of different levels of hierarchical organization is an often result of such self-assembly [1,4,5,7,8]. Among them, α-helices, a common type of regular secondary structure of many proteins, are the simplest and most energetically favorable structures [1,6]. For natural proteins consisting of L-amino acids, a chirality sign alternation from the “left-handed” type to the “right-handed” is observed at different levels of hierarchical organization [9,10,11,12]. In the case of DNA, the sequence of the chirality sign is “right-handed deoxyribose carbohydrate”–“left-handed nucleotides”–“right-handed DNA double helix”–“left-handed superhelix” with the increasing complexity of their level of organization. This feature of chirality is one of the key points in the hierarchy and self-organization of any biological system [9,10,11,12].
Despite the concept of chirality, in its qualitative sense, being widespread in natural sciences, its quantitative aspects (the magnitude and sign of chirality) are still poorly studied [13,14,15,16,17,18,19,20,21,22,23,24]. In structural biology, it is of great importance to obtain the quantitative estimates of the magnitude of chirality and the chirality sign to compare both molecular constructs with the same symmetry type and those with different types of the symmetry. The problem still lacks a solution, although many studies have been performed in the field. The detailed analysis of these studies can be found in [23].
Recently, a new method for the chirality quantification based on the scalar triple product of three consecutive vectors connecting Cα carbon atoms of neighboring amino acid residues in the polypeptide helical or superhelical structures was proposed [21,22,23,24]. Despite this approach being successfully applied to a variety of proteins [24] taken from the Protein Data Bank [25], until now it has not been used for helix-like supramolecular structures such as peptide nanotubes.
Peptide nanotubes (PNTs) is an important example of helix-like self-organizing supramolecular systems [5,26,27]. Peptides and their derivatives were recognized recently as biological and bio-inspired building blocks for the construction of various advanced functional materials for nanotechnology and biomedicine [28,29]. Short, linear peptides containing aromatic amino acid residues such as phenylalanine (F, H-Phe-OH) attract special attention due to their ability to mimic the self-organizing behavior of more complex proteins [26].
The simplest aromatic dipeptide capable to form helix-like nanotubes is diphenylalanine (FF, H-Phe-Phe-OH) [30,31,32,33,34,35,36,37,38,39,40]. Each turn of such helix PNT consists of six FF molecules (Figure 1). Fast self-assembly of such PNTs occurs in aqueous media, and the variation of external conditions allows tuning the PNT’s growth rate, length, and their physical properties [33,38,39,41,42,43,44,45]. FF PNTs possess a wide range of useful functional properties [41,46,47,48] that make them promising material for various applications in nanotechnology [33,41,46,47,48,49,50], nanoelectronics [28,35,48,51,52], and biomedicine [34,35,36,47,50,53,54,55].
The effect of chirality of FF molecules on the self-assembly and properties of PNTs has recently been studied in detail both experimentally and by computer simulation [37,38,42,56,57,58,59]. Lattice cell parameters for PNT made of “right-handed” FF molecules (H-D-Phe-D-Phe-OH, abbreviated hereafter D-FF) are close to those for PNTs made of “left-handed” FF molecules (H-L-Phe-L-Phe-OH, abbreviated hereafter L-FF), but their space groups are different [38]. Due to the difference of FF monomers chirality, the L-FF PNTs belong to the P61 space group, whereas D-FF PNTs belong to the P65. This P61–P65 pair is one of 11 pairs of enantiomorphic space groups [60] that are distinguished by the twisting direction of their 6-fold screw axis [61]. It was also shown that “left-handed” L-FF molecules form “right-handed” helix-like PNTs, whereas “right-handed” D-FF molecules form “left-handed” PNTs distinguished with their intermolecular interaction energies, self-assembly kinetics, and characteristic lengths [30,38].
In this work, we demonstrated that the method for chirality quantification proposed recently for protein helical structures [23,24] can be adopted for analysis of helix-like self-assembled PNTs. For the calculation of magnitude and sign of the chirality of L-FF and D-FF PNTs a set of sequential vectors of individual dipole moments of FF molecules comprising the turn of each helix of PNTs was used. The dipole moments were calculated using the HyperChem software [62].

2. Models Details and Computational Methods

2.1. Main Models and Used Software

Recently, we studied the structure and properties of empty L-FF and D-FF PNTs, as well as those with inner cavity filled with water molecules [56,57]. As water molecules do not affect the chirality of the PNTs, in this work, we considered empty (anhydrous) PNTs to simplify the calculations. The initial models of the PNTs were constructed using the same approach as in [56,57] based on X-ray crystallographic data for L-FF PNT (CCDC 16337, work [31]) and for D-FF PNT (CCDC 1853771, work [38]) taken from Cambridge Crystallographic Data Center (CCDC) [63]. The structural optimization and calculations were carried out using the density functional theory (DFT) methods (in Vienna Ab initio Simulation Package (VASP) program [64]), taking into account the Van der Waals interactions by “PBE + D3” method. The resulted molecular structures visualized by CCDC Mercury [65] are presented in Figure 1, whereas their main crystallographic parameters are summarized in Table 1.

2.2. Models of FF Nanotubes

The molecular structures of both L-FF and D-FF PNTs were converted using OpenBabel software [66] from *.cif to *.hin format for their further analysis and calculations of their polar properties with various molecular mechanical and quantum-mechanical semi-empirical methods in HyperChem package [62] (Figure 2). These structures contain two coils of the helix arranged along with the c axis. Each coil consists of 6 FF molecules (258 atoms) and coils are separated with a lattice constant c around 5.45 Å [56]. The repetition of the coils along the c-axis leads to the formation of PNT with the corresponding chirality: right-handed helix for L-FF and left-handed helix for D-FF (Figure 3).
It is known that water molecules in PNT nanochannels can affect both the structure and properties of the PNTs [41,44,56,57]. Therefore, in this work we considered in more detail the empty (anhydrous) nanotubes (Figure 1) to better understand how the dipole moments of FF molecules form the helix-like PNT with different chiralities and quantify the chirality.
The dipole moments, Di, of L-FF and D-FF molecules and corresponding helix-like structures were calculated using the semi-empirical quantum-mechanical method PM3 in the restricted Hartree-Fock approximation (RHF) and molecular mechanical force field method Amber from the HyperChem package [62]. Previous studies [37,38,42] have shown that for the dipole moments and energy calculated with other methods AM1 and BIOCHARM are similar to those obtained by PM3 and Amber. Therefore, in this work we used only PM3 and Amber methods. The calculated values of the dipole moments for individual L-FF and D-FF molecules are presented in Table 2. They are similar to the results obtained and analyzed earlier [30,37,38,42], and correspond well to the molecules’ orientation in experimentally observed helix-like structures.

3. Results and Discussions

In contrast to α-helix proteins, supramolecular PNTs are comprised of individual FF molecules held by relatively weak hydrogen bonds [31,32,33]. Therefore, the chirality quantification method developed earlier for protein structures [21,22,23,24] cannot be directly applied for PNTs and requires some adaptations.
Briefly, the original method considered a helical polypeptide chain consisting of n amino acid residues, and a set of (n − 1) vectors v i was built between each two adjacent Cα atoms in amino acid residues (Figure 4). For each three consecutive vectors, their scalar triple product was calculated:
( [ v 1 , v 2 ] , v 3 ) = ( y 1 z 2 y 2 z 1 ) x 3 + ( z 1 x 2 z 2 x 1 ) y 3 + ( x 1 y 2 x 2 y 1 ) z 3 ,
The sum of all these scalar triple products (see Equation (2)) allowed us to estimate the chirality sign. If χtotal is positive, the structure is right-handed; for left-handed structures, χtotal is negative [23,24]. This method has been validated for almost 1000 proteins [23,24] taken from the Protein Data Bank [25].
χ t o t a l = i = 1 n 3 ( [ v i , v i + 1 ] , v i + 2 )
In this work, the abovementioned approach was adopted for analysis of supramolecular FF PNTs. Instead of vectors between adjacent Cα atoms, a scalar triple product of dipole moments Di of the successive individual FF molecules constituting a turn of the PNT helix-like nanotube was used. The origin of Di vectors is taken relative to the center of mass of the corresponding molecules. The absolute value of each dipole moment Di is
D i = | D i | = D x , i 2 + D y , i 2 + D z , i 2 ,
where Dx,i, Dy,i, and Dz,i are the components of the i-th vector Di in the Cartesian coordinates. Similar to Equation (2), the sum of the scalar triple products of the dipole moments related to the PNT’s chirality can be written as:
c t o t a l = i = 1 n 2 ( [ D i , D i + 1 ] , D i + 2 ) .
It is necessary to note that the summation here is taken over i in the range from 1 to (n − 2), whereas in Equation (2), the i range is from 1 to (n − 3). This is because in supramolecular helixes i numerates the individual molecules instead of the Cα atoms in proteins. The ctotal can be normalized over the average value of the total dipole momentum of the PNT’s coil, D a v = 1 6 i = 1 6 D i , to find a universal measure of the chirality:
c n o r m = c t o t a l D a v 3 .
Individual dipole moments of FF molecules in one coil of helix-like PNTs were obtained using semiempirical PM3 method in restricted Hartree–Fock (RHF) approximation and molecular mechanic Amber method (after PM3) implemented in HyperChem software [62]. The results are presented below in Table 3 and Table 4 for L-FF and for D-FF, respectively. Schematic representation of the spatial arrangement of FF individual dipole moments Di in two coils of PNTs are shown in Figure 5a,b for L-FF and in Figure 5c,d for D-FF PNTs.
It is important to note that, due to the helix-like structure of PNT, the dipole moment Di of each next FF molecule in the coil is rotated by ~60° in the XOY plane. Therefore, at a full vector rotation at 360 degrees in a coil, the components Di,x and Di,y almost compensate for one another. Thus, the x and y components of the total dipole moment of the coil, Dcoil, are much smaller than Dcoil,z, which is always oriented along the OZ axis and increases Dcoil (Table 3 and Table 4). As a result, the total dipole moment of a coil Dcoil is directed mainly along OZ axis with slight deviations (Figure 5), which corresponds to the previously obtained data [30,37,38,56,57].
It is worth noting that, in contrast to the original chirality quantification method developed for proteins [21,22,23,24], where the vectors were built between the carbon atoms Cα of each subsequent amino acid, in the current modification of the method, the vector of the dipole moment of each FF molecule in the PNT is taken relative to the center of mass of the corresponding molecule. The origin of Di is a point in space defined by the vector rDi with components {xDi; yDi; zDi} calculated as follows:
x D i = { j = 1 N m j · x j } / { j = 1 N m j } ,
y D i = { j = 1 N m j · y j } / { j = 1 N m j } ,
z D i = { j = 1 N m j · z j } / { j = 1 N m j } .
Here, mj, xj, yj, and zj are the mass and coordinates, respectively, of the j-th atom in the i-th FF molecule in the PNT, and N = 43 is the number of atoms in one FF molecule. For example, the coordinates of the center of mass for the first FF molecule (i = 1) in a coil of L-FF PNT (Figure 6) are:
xD1 = 2.35 Å; yD1 = −7.80 Å; zD1 = 1.02 Å.
This point is the origin for the D1 dipole moment vector. Similarly, the origins for other vectors Di can be calculated. As a result, dipole moments form a helix with a pitch equal to c = 5.456 Å (for L-FF, c = 5.441 Å for D-FF PNTs), and the helix radius R is about 8.15 Å for L-FF PNT.
The obtained values of the dipole moments for L-FF and D-FF PNTs allow us to quantify their chiralities following the Equations (4) and (5). The calculated magnitudes of the PNTs chirality, ctotal, and the normalized chirality, cnorm, are presented in Table 5. For each type of PNT, both PM3 and Amber methods ctotal demonstrate close values with the divergence about 15%, whereas for cnorm the divergence is less than 5%. Therefore, cnorm can be considered as a universal value for chirality quantification that does not depend on the calculation method.
At the same time, the absolute value of cnorm for L-FF PNT is about 10% higher than that for D-FF PNT. This difference exceeds the calculation error and thus shows the difference in PNTs chiralities. Lower cnorm value observed for D-FF PNTs indicates that this PNT is twisted tighter than L-FF. This is also confirmed by the lower volume of the D-FF PNT unit cell (Table 1).
Following the original method [23,24], the sign of the cnorm corresponds to the PNT’s chirality type. For L-FF PNT, cnorm is positive thus this PNT should be right-handed, whereas negative cnorm value for D-FF PNT indicates its left-handed twisting. This result is confirmed by the previous crystallographic studies [31,38] and the individual dipole moments completely follow this arrangement (Figure 5). It is worth noting that the chirality alternating observed earlier for natural proteins and DNA [9,10,11,12] preserves in the supramolecular PNTs as well. The type of chirality of the helix-like PNT is opposite to that of the individual dipeptide. This fact also can be a confirmation of the adequacy of the proposed method for supramolecular PNTs chirality quantification.

4. Conclusions

A method for quantification of the chirality of self-assembled helix-like FF nanotubes based on the scalar triple products of the individual FF molecules dipole moments is described. The dipole moments were calculated for nanotubes comprised of L-FF and D-FF molecules by quantum-chemical and molecular mechanics methods, and the independence of the magnitude and the sign of the chirality on the calculation method is demonstrated. The obtained magnitudes of the chirality for L-FF nanotubes are about 10% higher than those of L-FF, which indicates that D-FF nanotubes are twisted tighter than L-FF. The alternating of the chirality type observed earlier for natural proteins and DNA also preserves in the supramolecular PNTs. The type of chirality of the helix-like PNT is opposite to that of the individual dipeptide. This effect is in line with the chirality alternation rule, general for different levels of hierarchical organization of molecular systems, and additionally corroborates the validity of the proposed method.
The extension of the chirality quantification method to supramolecular helix-like nanostructures opens new facilities for comparing both molecular constructs of the same chirality type and those with different constructs. Moreover, it provides an opportunity to reveal the physical basis for the chirality sign formation, which is associated with the electrostatic dipole–dipole interaction of individual molecules. This approach can be applied to study other helical and helix-like supramolecular structures.

Author Contributions

V.B., A.S. and P.Z. wrote the manuscript. A.K. and E.M. supervised and supported this study. E.M. also developed the concept of chirality. A.L. and D.S. performed the calculations. S.K., A.N. and P.Z. participated in the experimental data processing and data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Foundation for Basic Research, grant n. 19-01-00519-a.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors are grateful to the Russian Foundation for Basic Researches (RFBR): grants №№ 19–01–00519_A and 20-51-53014_GFEN_A. Part of this work (A.K.) was supported by the Ministry of Science and Higher Education of the Russian Federation (grant No. 075-15-2021-588). P.Z., S.K. and A.K. are grateful to the FCT project “BioPiezo”—PTDC/CTM−CTM/31679/2017 (CENTRO-01-0145-FEDER-031679) and to the project CICECO-Aveiro Institute of Materials, refs. UIDB/50011/2020 and UIDP/50011/2020, financed by national funds through the Portuguese Foundation for Science and Technology/MCTES. Part of this work was funded by national funds (OE), through FCT–Fundação para a Ciência e a Tecnologia, I.P., in the scope of the framework contract foreseen in the numbers 4, 5, and 6 of the article 23, of the Decree-Law 57/2016, of August 29, changed by Law 57/2017, of July 19. A.S. and E.M. are grateful to the the Interdisciplinary Scientific and Educational School of Moscow University “Fundamental and Applied Space Research”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lehninger, A.L. Biochemistry—The Molecular Basis of Cell Structure and Function; Worth Publishers, Inc.: New York, NY, USA, 1972. [Google Scholar]
  2. Stock, D.; Leslie, A.G.; Walker, J.E. Molecular architecture of the rotary motor in ATP synthase. Science 1999, 286, 1700–1705. [Google Scholar] [CrossRef] [Green Version]
  3. Ban, N.; McPherson, A. The structure of satellite panicum mosaic virus at 1.9 Å resolution. Nat. Struct. Biol. 1995, 2, 882–890. [Google Scholar] [CrossRef]
  4. Zhang, S. Emerging biological materials through molecular self-assembly. Biotechn. Adv. 2002, 20, 321–339. [Google Scholar] [CrossRef]
  5. Levin, A.; Hakala, T.A.; Schnaider, L.; Bernardes, G.J.L.; Gazit, E.; Knowles, T.P.J. Biomimetic peptide self-assembly for functional materials. Nat. Rev. Chem. 2020, 4, 615–634. [Google Scholar] [CrossRef]
  6. Mendes, A.C.; Baran, E.T.; Reis, R.L.; Azevedo, H.S. Self-assembly in nature: Using the principles of nature to create complex nanobiomaterials. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2013, 5, 582–612. [Google Scholar] [CrossRef] [PubMed]
  7. Whitesides, G.M.; Mathias, J.P.; Seto, C.T. Molecular self-assembly and nanochemistry: A chemical strategy for the synthesis of nanostructures. Science 1991, 254, 1312–1319. [Google Scholar] [CrossRef] [PubMed]
  8. Mandal, D.; Shirazi, A.N.; Parang, K. Self-assembly of peptides to nanostructures. Org. Biomol. Chem. 2014, 12, 3544–3561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Tverdislov, V.A. Chirality as a primary switch of hierarchical levels in molecular biological systems. Biophysics 2013, 58, 128–132. [Google Scholar] [CrossRef]
  10. Malyshko, E.V.; Tverdislov, V.A. Chirality as a physical aspect of structure formation in biological macromolecular systems. IOP J. Phys. Conf. Ser. 2016, 741, 012065. [Google Scholar] [CrossRef] [Green Version]
  11. Tverdislov, V.A.; Malyshko, E.V. Chiral Dualism as an Instrument of Hierarchical Structure Formation in Molecular Biology. Symmetry 2020, 12, 587. [Google Scholar] [CrossRef] [Green Version]
  12. Tverdislov, V.A.; Malyshko, E.V. On regularities in the spontaneous formation of structural hierarchies in chiral systems of nonliving and living matter. Phys.-Uspekhi 2019, 62, 354–363. [Google Scholar] [CrossRef]
  13. Buda, A.B.; der Heyde, T.A.; Mislow, K. On Quantifying Chirality. Angew. Chem. Int. Ed. Engl. 1992, 31, 989–1007. [Google Scholar] [CrossRef]
  14. Weinberg, N.; Mislow, K. On chirality measures and chirality properties. Can. J. Chem. 2000, 78, 41–45. [Google Scholar] [CrossRef]
  15. Randic, M. Shape recognition and chirality measure: Reestablishing the link between similarity and dissimilarity in discrete space. Chemom. Intell. Lab. Syst. 1991, 10, 213–225. [Google Scholar] [CrossRef]
  16. Zabrodsky, H.; Peleg, S.; Avnir, D. Continuous symmetry measures. J. Am. Chem. Soc. 1992, 114, 7843–7851. [Google Scholar] [CrossRef]
  17. Mezey, P.G. Chirality Measures and Graph Representations. Comput. Math. Appl. 1997, 34, 105–112. [Google Scholar] [CrossRef] [Green Version]
  18. Pinsky, M.; Dryzun, C.; Casanova, D.; Alemany, P.; Avnir, D. Analytical methods for calculating Continuous Symmetry Measures and the Chirality Measure. J. Comput. Chem. 2008, 29, 2712–2721. [Google Scholar] [CrossRef] [PubMed]
  19. Petitjean, M. Chirality and Symmetry Measures: A Transdisciplinary Review. Entropy 2003, 5, 271–312. [Google Scholar] [CrossRef]
  20. Zhao, T.; Zhang, Q.; Long, H.; Xu, L. Graph Theoretical Representation of Atomic Asymmetry and Molecular Chirality of Benzenoids in Two-Dimensional Space. PLoS ONE 2014, 9, e102043. [Google Scholar] [CrossRef]
  21. Sidorova, A.E.; Malyshko, E.V.; Kotov, A.R.; Tverdislov, V.A.; Ustinin, M.N. Quantitative Criteria of Chirality in Hierarchical Protein Structures. Biophysics 2019, 64, 155–166. [Google Scholar] [CrossRef]
  22. Sidorova, A.E.; Levashova, N.T.; Malyshko, E.V.; Tverdislov, V.A. Autowave Self-Organization in the Folding of Proteins. Mosc. Univ. Phys. Bull. 2019, 74, 213–226. [Google Scholar] [CrossRef]
  23. Sidorova, A.E.; Lutsenko, A.O.; Shpigun, D.K.; Malyshko, E.V.; Tverdislov, V.A. Methods for Determining the Chirality sign of the Helical and Superhelical Protein Structures. Biophysics 2021, 66, 421. [Google Scholar] [CrossRef]
  24. Sidorova, A.E.; Malyshko, E.V.; Lutsenko, A.O.; Shpigun, D.K.; Bagrova, O.E. Protein Helical Structures: Defining Handedness and Localization Features. Symmetry 2021, 13, 879. [Google Scholar] [CrossRef]
  25. RCSB Protein Data Bank. Available online: http://www.rcsb.org (accessed on 30 August 2021).
  26. Yuan, C.; Ji, W.; Xing, R.; Li, J.; Gazit, E.; Yan, X. Hierarchically oriented organization in supramolecular peptide crystals. Nat. Rev. Chem. 2019, 3, 567. [Google Scholar] [CrossRef]
  27. Hamley, I.W. Peptide Nanotubes. Angew. Chem. Int. Ed. 2014, 53, 6866–6881. [Google Scholar] [CrossRef]
  28. Tao, K.; Makam, P.; Aizen, R.; Gazit, E. Self-assembling peptide semiconductors. Science 2017, 358, eaam9756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Ghadiri, M.R.; Granja, J.R.; Milligan, R.A.; McRee, D.E.; Khazanovich, N. Self-assembling organic nanotubes based on cyclic peptide architecture. Nature 1993, 366, 324–327. [Google Scholar] [CrossRef]
  30. Bystrov, V.S.; Zelenovskiy, P.S.; Nuraeva, A.S.; Kopyl, S.; Zhulyabina, O.A.; Tverdislov, V.A. Chiral peculiar properties of self-organization of diphenylalanine peptide nanotubes: Modeling of structure and properties. Math. Biol. Bioinform. 2019, 14, 94–124. [Google Scholar] [CrossRef] [Green Version]
  31. Görbitz, C.H. Nanotube formation by hydrophobic dipeptides. Chem. Eur. J. 2001, 7, 5153–5159. [Google Scholar] [CrossRef]
  32. Görbitz, C.H. The structure of nanotubes formed by diphenylalanine, the core recognition motif of Alzheimer′s b-amyloid polypeptide. Chem. Commun. 2006, 1, 2332–2334. [Google Scholar] [CrossRef]
  33. Kim, J.; Han, T.E.; Kim, Y.; Park, J.S.; Choi, J.; Churchill, D.G.; Kim, S.O.; Ihee, H. Role of water in directing diphenylalanine assembly into nanotubes and nanowires. Adv. Mater. 2010, 22, 583–587. [Google Scholar] [CrossRef]
  34. Bystrov, V.S.; Bdikin, I.; Heredia, A.; Pullar, R.C.; Mishina, E.; Sigov, A.; Kholkin, A.L. Piezoelectricity and Ferroelectricity in biomaterials: From proteins to self-assembled peptide nanotubes. In Piezoelectric Nanomaterials for Biomedical Applications; Ciofani, G., Menciassi, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 187–211. [Google Scholar]
  35. Bystrov, V.S. Computer Simulation Nanostructures: Bioferroelectric Peptide Nanotubes. Bioferroelectricity: Peptide Nanotubes; LAP Lambert Academic Publishing: Saarbruecken, Germay, 2016. [Google Scholar]
  36. Bystrov, V.S.; Paramonova, E.V.; Bdikin, I.K.; Kopyl, S.; Heredia, A.; Pullar, R.C.; Kholkin, A.L. Bioferroelectricity: Diphenylalanine peptide nanotubes computational modeling and ferroelectric properties at the nanoscale. Ferroelectrics 2012, 440, 3–24. [Google Scholar] [CrossRef]
  37. Bystrov, V.S.; Zelenovskiy, P.S.; Nuraeva, A.S.; Kopyl, S.; Zhulyabina, O.A.; Tverdislov, V.A. Molecular modeling and computational study of the chiral-dependent structures and properties of the self-assembling diphenylalanine peptide nanotubes. J. Mol. Modeling 2019, 25, 199. [Google Scholar] [CrossRef]
  38. Zelenovskiy, P.S.; Nuraeva, A.S.; Kopyl, S.; Arkhipov, S.G.; Vasilev, S.G.; Bystrov, V.S.; Gruzdev, D.A.; Waliszek, M.; Svitlyk, V.; Shur, V.Y.; et al. Chirality-Dependent Growth of Self-Assembled Diphenylalanine Microtubes. Cryst. Growth Des. 2019, 19, 6414–6421. [Google Scholar] [CrossRef]
  39. Nuraeva, A.; Vasilev, S.; Vasileva, D.; Zelenovskiy, P.; Chezganov, D.; Esin, A.; Kopyl, S.; Romanyuk, K.; Shur, V.Y.; Kholkin, A.L. Evaporation-Driven Crystallization of Diphenylalanine Microtubes for Microelectronic Applications. Cryst. Growth Des. 2016, 16, 1472–1479. [Google Scholar] [CrossRef]
  40. Yuan, T.; Xu, Y.; Fei, J.; Xue, H.; Li, X.; Wang, C.; Fytas, G.; Li, J. The Ultrafast Assembly of a Dipeptide Supramolecular Organogel and its Phase Transition from Gel to Crystal. Angew. Chem. Int. Ed. 2019, 58, 11072–11077. [Google Scholar] [CrossRef] [PubMed]
  41. Andrade-Filho, T.; Martins, T.C.; Ferreira, F.F.; Alves, W.A.; Rocha, A.R. Water-driven stabilization of diphenylalanine nanotube structures. Theor. Chem. Acc. 2016, 135, 185. [Google Scholar] [CrossRef] [Green Version]
  42. Bystrov, V.S.; Kopyl, S.A.; Zelenovskiy, P.; Zhulyabina, O.A.; Tverdislov, V.A.; Salehli, F.; Ghermani, N.E.; Shur, V.Y.; Kholkin, A.L. Investigation of physical properties of diphenylalanine peptide nanotubes having different chiralities and embedded water molecules. Ferroelectrics 2018, 525, 168–177. [Google Scholar] [CrossRef]
  43. Dayarian, S.; Kopyl, S.; Bystrov, V.; Correia, M.R.; Ivanov, M.S.; Pelegova, E.; Kholkin, A. Effect of the Chloride Anions on the Formation of Self-Assembled Diphenylalanine Peptide Nanotubes. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2018, 65, 1563–1570. [Google Scholar] [CrossRef]
  44. Zelenovskiy, P.; Yuzhakov, V.; Nuraeva, A.; Kornev, M.; Shur, V.Y.; Kopyl, S.; Kholkin, A.; Vasilev, S.; Tofail, S.A.M. The Effect of Water Molecules on Elastic and Piezoelectric Properties of Diphenylalanine Microtubes. IEEE Trans. Dielectr. Electr. Insul. 2020, 27, 1474–1477. [Google Scholar] [CrossRef]
  45. Guo, C.; Luo, Y.; Zhou, R.; Wei, G. Probing the Self-Assembly Mechanism of Diphenylalanine-Based Peptide Nanovesicles and Nanotubes. ACS Nano 2012, 6, 3907–3918. [Google Scholar] [CrossRef]
  46. Kholkin, A.; Amdursky, N.; Bdikin, I.; Gazit, E.; Rosenman, G. Strong piezoelectricity in bioinspired peptide nanotubes. ACS Nano 2010, 4, 610–614. [Google Scholar] [CrossRef] [PubMed]
  47. Hereida, A.; Bdikin, I.; Kopyl, S.; Mishina, E.; Semin, S.; Sigov, A.; German, K.; Bystrov, V.; Gracio, J.; Kholkin, A.L. Temperature-driven phase transformation in self-assembled diphenylalanine peptide nanotubes. J. Phys. D Appl. Phys. 2010, 43, 462001. [Google Scholar] [CrossRef] [Green Version]
  48. Bdikin, I.; Bystrov, V.S.; Delgadillo, I.; Gracio, J.; Kopyl, S.; Wojtas, M.; Mishina, E.; Sigov, A.; Kholkin, A.L. Polarization switching and patterning in self-assembled peptide tubular structures. J. Appl. Phys. 2012, 111, 074104. [Google Scholar] [CrossRef] [Green Version]
  49. Reches, M.; Gazit, E. Controlled patterning of aligned self-assembled peptide nanotubes. Nat. Nanotechnol. 2006, 1, 195–200. [Google Scholar] [CrossRef] [PubMed]
  50. Silva, R.F.; Araújo, D.R.; Silva, E.R.; Ando, R.A.; Alves, W.A. L-diphenylalanine microtubes as a potencial drug-delivery system: Characterization, release kinetics, and cytotoxicity. Langmuir 2013, 29, 10205–10212. [Google Scholar] [CrossRef]
  51. Reches, M.; Gazit, E. Casting metal nanowires within discrete self-assembled peptide nanotubes. Science 2003, 300, 625–627. [Google Scholar] [CrossRef] [Green Version]
  52. Nguyen, V.; Zhu, R.; Jenkins, K.; Yang, R. Self-assembly of diphenylalanine peptide with controlled polarization for power generation. Nat. Commun. 2016, 7, 13566. [Google Scholar] [CrossRef] [Green Version]
  53. Emtiazi, G.; Zohrabi, T.; Lee, L.Y.; Habibi, N.; Zarrabi, A. Covalent diphenylalanine peptide nanotube conjugated to folic acid/magnetic nanoparticles for anti-cancer drug delivery. J. Drug Deliv. Sci. Technol. 2017, 41, 90–98. [Google Scholar] [CrossRef]
  54. Gupta, S.; Singh, I.; Sharma, A.K.; Kumar, P. Ultrashort Peptide Self-Assembly: Front-Runners to Transport Drug and Gene Cargos. Front. Bioeng. Biotechnol. 2020, 8, 504. [Google Scholar] [CrossRef]
  55. Marchesan, S.; Vargiu, A.V.; Styan, K.E. The Phe-Phe Motif for Peptide Self-Assembly in Nanomedicine. Molecules 2015, 20, 19775–19788. [Google Scholar] [CrossRef] [Green Version]
  56. Bystrov, V.S.; Coutinho, J.; Zelenovskiy, P.; Nuraeva, A.; Kopyl, S.; Zhulyabina, O.; Tverdislov, V. Structures and properties of the self-assembling diphenylalanine peptide nanotubes containing water molecules: Modeling and data analysis. Nanomaterials 2020, 10, 1999. [Google Scholar] [CrossRef]
  57. Bystrov, V.S.; Coutinho, J.; Zhulyabina, J.A.; Kopyl, S.A.; Zelenovskiy, P.S.; Nuraeva, A.S.; Tverdislov, V.A.; Filippov, S.V.; Kholkin, A.L.; Shur, V.Y. Modeling and physical properties of diphenylalanine peptide nanotubes containing water molecules. Ferroelectrics 2021, 574, 78–91. [Google Scholar] [CrossRef]
  58. Tamamis, P.; Adler-Abramovich, L.; Reches, M.; Marshall, K.; Sikorski, P.; Serpell, L.; Gazit, E.; Archontis, G. Self-assembly of phenylalanine oligopeptides: Insights from experiments and simulations. Biophys. J. 2009, 96, 5020–5029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Zheng, Y.; Mao, K.; Chen, S.; Zhu, H. Chirality Effects in Peptide Assembly Structures. Front. Bioeng. Biotechnol. 2021, 9, 703004. [Google Scholar] [CrossRef] [PubMed]
  60. Müller, U. Symmetry Relationships between Crystal Structures. Applications of Crystallographic Group Theory in Crystal Chemistry; University Press: Oxford, UK, 2013. [Google Scholar]
  61. Hahn, T. (Ed.) International Tables for Crystallography. Volume A: Space-Group Symmetry; Springer: Berlin/Heidelberg, Germany, 2005. [Google Scholar]
  62. HyperChem 8. Tools for Molecular Modeling. Professional Edition for Windows AC Release 8.0 USB (on CD); Hypercube. Inc.: Gainesville, FL, USA, 2011. [Google Scholar]
  63. The Cambridge Crystallographic Data Centre (CCDC). Available online: https://www.ccdc.cam.ac.uk/ (accessed on 30 August 2021).
  64. VASP (Vienna Ab initio Simulation Package). Available online: https://www.vasp.at/ (accessed on 30 August 2021).
  65. CCDC Mercury. Available online: https://www.ccdc.cam.ac.uk/solutions/csd-system/components/mercury/ (accessed on 30 August 2021).
  66. Open Babel. Available online: https://openbabel.org/docs/dev/Installation/install.html (accessed on 30 August 2021).
Figure 1. Images of molecular crystals composed of: (a) L-FF PNTs (space group P61), and (b) D-FF PNTs (space group P65). Hexagonal unit cells are marked with green, red, and white lines. The individual PNTs in crystal are highlighted by yellow circles. Atom colors: oxygen—red, nitrogen—blue, carbon—grey, and hydrogen—white.
Figure 1. Images of molecular crystals composed of: (a) L-FF PNTs (space group P61), and (b) D-FF PNTs (space group P65). Hexagonal unit cells are marked with green, red, and white lines. The individual PNTs in crystal are highlighted by yellow circles. Atom colors: oxygen—red, nitrogen—blue, carbon—grey, and hydrogen—white.
Nanomaterials 11 02415 g001
Figure 2. Two coils of FF PNT in HyperChem workspace based on experimental crystallographic data [31,38]: (a) L-FF in the Z-plane, (b) L-FF in the Y-plane, (c) D-FF in the Z-plane, and (d) D-FF in the Y-plane.
Figure 2. Two coils of FF PNT in HyperChem workspace based on experimental crystallographic data [31,38]: (a) L-FF in the Z-plane, (b) L-FF in the Y-plane, (c) D-FF in the Z-plane, and (d) D-FF in the Y-plane.
Nanomaterials 11 02415 g002
Figure 3. Images of helix-like FF PNTs based on experimental crystallographic data [31,38]: (a) L-FF and (b) D-FF. L-FF PNT shows right-handed twist, whereas D-FF shows left-handed.
Figure 3. Images of helix-like FF PNTs based on experimental crystallographic data [31,38]: (a) L-FF and (b) D-FF. L-FF PNT shows right-handed twist, whereas D-FF shows left-handed.
Nanomaterials 11 02415 g003
Figure 4. Graphical representation of vectors between neighboring Cα atoms in a helical protein used for calculating a scalar triple product.
Figure 4. Graphical representation of vectors between neighboring Cα atoms in a helical protein used for calculating a scalar triple product.
Nanomaterials 11 02415 g004
Figure 5. Schematic presentation of dipole moments Di in two coils of (a,b) L-FF and (c,d) D-FF PNTs: (a,c) Z-plane projection, (b,d) Y-plane projection. For L-FF PNT dipole moments form a right-hand helix, whereas for D-FF PNT form a left-hand helix. Red arrows show the directions of the total dipole moments of the coil Dcoil.
Figure 5. Schematic presentation of dipole moments Di in two coils of (a,b) L-FF and (c,d) D-FF PNTs: (a,c) Z-plane projection, (b,d) Y-plane projection. For L-FF PNT dipole moments form a right-hand helix, whereas for D-FF PNT form a left-hand helix. Red arrows show the directions of the total dipole moments of the coil Dcoil.
Nanomaterials 11 02415 g005
Figure 6. A coil of L-FF PNT with vectors rD1 and rD4 pointing the centers of mass of individual FF molecules with dipole moments D1 and D4, correspondingly.
Figure 6. A coil of L-FF PNT with vectors rD1 and rD4 pointing the centers of mass of individual FF molecules with dipole moments D1 and D4, correspondingly.
Nanomaterials 11 02415 g006
Table 1. Experimental lattice cell parameters for L-FF [31] and D-FF [38] PNTs and inner cavity sizes R1, R2 [56].
Table 1. Experimental lattice cell parameters for L-FF [31] and D-FF [38] PNTs and inner cavity sizes R1, R2 [56].
L-FFD-FF
Space groupP61P65
a, Å24.070923.9468
b, Å24.070923.9468
c, Å5.45605.4411
V, Å32737.72702.2
R1, Å15.315.2
R2, Å12.212.1
Table 2. The dipole moments and some other structural parameters of individual L-FF and D-FF molecules calculated by PM3 method.
Table 2. The dipole moments and some other structural parameters of individual L-FF and D-FF molecules calculated by PM3 method.
MoleculeDx,
Debye
Dy,
Debye
Dz,
Debye
Dtot,
Debye
Van der Waals Volume, Å3Polarization, C/m2Total Energy, a.u.RMS
Gradient, a.u./Å
L-FF11.6451.1150.89911.733291.9190.134−133.959~0.06
D-FF−11.6301.0521.11311.730291.9770.134−133.959~0.07
Table 3. Values of dipole moments for a coil of the helix-like L-FF PNT computed using PM3 (RHF) and Amber (after PM3) methods. All values of dipole moments are given in Debye units.
Table 3. Values of dipole moments for a coil of the helix-like L-FF PNT computed using PM3 (RHF) and Amber (after PM3) methods. All values of dipole moments are given in Debye units.
iPM3 RHFAmber
DiDxDyDzDiDxDyDz
124.02214.576−15.421−11.26123.45814.901−15.250−9.781
222.549−6.313−18.923−10.51321.734−6.280−18.879−8.748
322.389−18.646−3.636−11.84921.545−18.698−3.629−10.070
422.381−11.56414.461−12.57321.530−11.69514.495−10.801
522.4417.55517.308−12.12321.5787.39717.408−10.384
622.58718.7672.568−12.30321.63818.5812.745−10.742
Dcoil70.8514.376−3.643−70.62260.7524.206−3.109−60.526
Dav22.7280.729−0.607−11.77021.9140.701−0.518−10.088
Table 4. Values of dipole moments for a coil of the helix-like D-FF PNT computed using PM3 (RHF) and Amber (after PM3) methods. All values of dipole moments are given in Debye units.
Table 4. Values of dipole moments for a coil of the helix-like D-FF PNT computed using PM3 (RHF) and Amber (after PM3) methods. All values of dipole moments are given in Debye units.
iPM3 RHFAmber
DiDxDyDzDiDxDyDz
122.523−12.228−15.267−11.16721.707−12.299−15.170−9.475
222.3407.302−18.014−11.07221.5277.210−17.995−9.360
322.37219.234−2.597−11.12521.52019.168−2.656−9.416
422.47511.90515.905−11.29021.62511.91415.274−9.612
522.629−6.61317.478−12.76121.703−6.48717.386−11.256
623.855−19.8204.382−12.53123.271−19.8934.727−11.112
Dcoil69.971−0.2181.888−69.94560.253−0.3871.565−60.231
Dav22.704−0.0360.315−11.65821.892−0.0640.261−10.038
Table 5. Magnitudes and signs of the chirality obtained for L-FF and D-FF PNTs for various calculating methods.
Table 5. Magnitudes and signs of the chirality obtained for L-FF and D-FF PNTs for various calculating methods.
Type of PNTL-FFD-FF
Calculating MethodPM3AmberPM3Amber
ctotal, Debye316,034.8213,870.71−14,497.03−12,161.23
cnorm1.371.32−1.23−1.16
Chirality signpositivepositivenegativenegative
Chirality symbolDDLL
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bystrov, V.; Sidorova, A.; Lutsenko, A.; Shpigun, D.; Malyshko, E.; Nuraeva, A.; Zelenovskiy, P.; Kopyl, S.; Kholkin, A. Modeling of Self-Assembled Peptide Nanotubes and Determination of Their Chirality Sign Based on Dipole Moment Calculations. Nanomaterials 2021, 11, 2415. https://doi.org/10.3390/nano11092415

AMA Style

Bystrov V, Sidorova A, Lutsenko A, Shpigun D, Malyshko E, Nuraeva A, Zelenovskiy P, Kopyl S, Kholkin A. Modeling of Self-Assembled Peptide Nanotubes and Determination of Their Chirality Sign Based on Dipole Moment Calculations. Nanomaterials. 2021; 11(9):2415. https://doi.org/10.3390/nano11092415

Chicago/Turabian Style

Bystrov, Vladimir, Alla Sidorova, Aleksey Lutsenko, Denis Shpigun, Ekaterina Malyshko, Alla Nuraeva, Pavel Zelenovskiy, Svitlana Kopyl, and Andrei Kholkin. 2021. "Modeling of Self-Assembled Peptide Nanotubes and Determination of Their Chirality Sign Based on Dipole Moment Calculations" Nanomaterials 11, no. 9: 2415. https://doi.org/10.3390/nano11092415

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