Next Article in Journal
Environmentally Friendly Compatibilizers from Soybean Oil for Ternary Blends of Poly(lactic acid)-PLA, Poly(ε-caprolactone)-PCL and Poly(3-hydroxybutyrate)-PHB
Next Article in Special Issue
Finite Element Analysis of Tunable Composite Tubes Reinforced with Auxetic Structures
Previous Article in Journal
Time-Resolved Photoluminescence Microscopy for the Analysis of Semiconductor-Based Paint Layers
Previous Article in Special Issue
The Isotropic and Cubic Material Designs. Recovery of the Underlying Microstructures Appearing in the Least Compliant Continuum Bodies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Auxeticity of Yukawa Systems with Nanolayers in the (111) Crystallographic Plane

by
Paweł M. Pigłowski
1,
Jakub W. Narojczyk
1,*,
Artur A. Poźniak
2,
Krzysztof W. Wojciechowski
1 and
Konstantin V. Tretiakov
1
1
Institute of Molecular Physics, Polish Academy of Sciences, Smoluchowskiego 17/19, 60-179 Poznan, Poland
2
Department of Technical Physics, Poznan University of Technology, Piotrowo 3, 60-695 Poznan, Poland
*
Author to whom correspondence should be addressed.
Materials 2017, 10(11), 1338; https://doi.org/10.3390/ma10111338
Submission received: 31 October 2017 / Revised: 16 November 2017 / Accepted: 17 November 2017 / Published: 22 November 2017
(This article belongs to the Special Issue Auxetic Materials 2017-2018)

Abstract

:
Elastic properties of model crystalline systems, in which the particles interact via the hard potential (infinite when any particles overlap and zero otherwise) and the hard-core repulsive Yukawa interaction, were determined by Monte Carlo simulations. The influence of structural modifications, in the form of periodic nanolayers being perpendicular to the crystallographic axis [111], on auxetic properties of the crystal was investigated. It has been shown that the hard sphere nanolayers introduced into Yukawa crystals allow one to control the elastic properties of the system. It has been also found that the introduction of the Yukawa monolayers to the hard sphere crystal induces auxeticity in the [ 11 1 ¯ ] [ 112 ] -direction, while maintaining the negative Poisson’s ratio in the [ 110 ] [ 1 1 ¯ 0 ] -direction, thus expanding the partial auxeticity of the system to an additional important crystallographic direction.

1. Introduction

The colloidal crystals [1] have been a subject of intense studies in recent years [2,3,4,5,6,7,8,9]. This is mainly due to their potential applications [10,11,12]. Some physical properties of charge-stabilized colloids can be well described using the hard-core repulsive Yukawa potential [13,14,15,16,17,18]. Recently, it has been shown that systems with such kind of interaction exhibit interesting elastic properties, namely these systems reveal a negative Poisson’s ratio [19]. The Poisson’s ratio describes the lateral mechanical response under the longitudinal uniaxial strain of infinitesimal magnitude [20]. Materials with negative Poisson’s ratio (so-called auxetics) [21] increase their dimensions in directions perpendicular to the applied stretching and shrink in directions perpendicular to the applied compression. In general, Poisson’s ratio depends on both the direction of applied deformation and the direction of measurement of response to external stress. Materials in which the Poisson’s ratio is negative only in some crystallographic directions are called partial auxetics [22].
Auxetics are a relatively new class of materials. A few decades ago, the first man-made material with a negative Poisson’s ratio [23] and the model system [24] exhibiting auxetic properties were presented. Since then, one can observe quickly growing interest in this class of materials because of their unusual, counter-intuitive properties [25,26,27,28], which originate from various mechanisms involving both micro- and macroscopic structures [23,29,30,31,32,33,34,35,36,37,38], inter-particle interactions [24,39,40,41] as well as special conditions applied to the system [42,43]. At present, research is being conducted extensively in the pursuit for auxetic properties in already existing and new materials [39,44,45,46,47,48,49,50,51,52], the fabrication of auxetic composites [53,54,55], the construction of models for explaining auxetic properties [56,57,58,59,60,61,62,63,64], and the search for new mechanisms leading to the appearance of auxetic properties in various systems. The studies on the influence of structural modifications on the auxetic properties of Yukawa crystals are an example of the latter [65,66]. In the referenced studies, it has been shown that the introduction of nanochannel filled by hard spheres into Yukawa crystal enables enhancing the auxetic properties [65]. On the other hand, the introduction of the monolayer of hard spheres in the Yukawa crystal oriented in the (010) crystallographic plane leads to the appearance of a new auxetic direction to emerge [66]. Gathering the facts, the following question arises: what effect on the auxetic properties will the introduction of nanolayer into the Yukawa crystal oriented in a different crystallographic direction have?
This work concerns the study of the effects of structural modification in the form of insertion into crystal of nanolayers that are oriented parallel to the (111) crystallographic plane. The aim of this work is to determine the elastic properties of such structurally modified crystals, in which the particles interact via hard and hard-core repulsive Yukawa potentials. In particular, we examine the effect of this structural modification on the auxetic properties of the system under study. The purpose of this paper is to present data that show that the discussed structural modifications have a significant effect on auxeticity of the studied model system. We believe that in a not far future it will be possible to use this knowledge in designing real materials.

2. Model and Computational Details

In the considered model, the initial face-centered cubic structure forms by particles interacting through the hard-core repulsive Yukawa potential (HCRYP) [13,14,15]
β u i j = , r i j < σ , β ε exp [ κ σ ( r i j / σ 1 ) ] r i j / σ , r i j σ ,
where β = 1 / k B T , k B is the Boltzmann constant, T is the temperature, σ is the diameter of the particles’ hard core, ϵ is the contact potential, and κ is the inverse of the Debye screening length. In order to form a system with nanolayers, Yukawa particles that belong to periodic arrays of crystallographic planes (111) are replaced by particles interacting via hard potential
β u i j = , r i j < σ , 0 , r i j σ .
The result is the system of Yukawa particles with nanolayers formed by hard spheres. Figure 1 illustrates a few examples of studied systems. From Figure 1a, one can see that the same system can be considered as the system with nanolayers consisting of monolayers of hard spheres or as the system with nanolayers consisting of multilayers of particles interacting through Yukawa potential. On the other hand, the system with nanolayers consisting of monolayers of particles interacting via Yukawa potential can be viewed as the system with nanolayers consisting of multilayers of hard spheres (see Figure 1c). Table 1 contains the details of the studied systems. The use of periodic boundary conditions in the directions x , y , z leads to the periodic structure with parallel nanolayers. Then, the systems described in the table can be considered as single super-cells. In order to analyse obtained results, we introduce concentration as a ratio of nanolayer particles ( N HS ) to all particles in the system (N) [66]
c = N HS N × 100 % .
In order to determine the elastic properties of the studied systems, extensive Monte Carlo simulations in isobaric-isothermal ensemble ( N p T ) were performed using the Parrinello-Rahman method [40,67,68]. As the result of using that method, one obtains an elastic compliance 4th order tensor ( S i j k l ) that describes elastic properties of studied systems entirely. Knowing each component of the compliance tensor, the Poisson’s ratio in arbitrary crystallographic direction and for any symmetry of the crystal can be determined using the formula [69]
ν n m = m i m j S i j k l n k n l n p n r S p r s t n s n t ,
where n ^ is a versor in the direction of the applied load and m ^ is the versor in the direction in which the Poisson’s ratio is measured, and the following relation has to be fulfilled: n ^ · m ^ = 0 . More details on calculating the Poisson’s ratio can be found in References [65,66].
Elastic properties of studied systems were determined for the following parameters of the Yukawa potential: κ σ = 10 , β ε = 20 , and pressure p * β P σ 3 = 100 expressed in dimensionless units. The choice of these parameters was made on the basis of previous works [16,19,65]. The cut-off radius used in the simulations was equal to 2.5 σ . The acceptance ratio of box moves and particle moves in the Monte Carlo method was 30%. In order to improve the accuracy of the obtained results, all physical values (elastic compliances and the Poisson’s ratios) were averaged over 10 independent runs for each of the studied systems. In other words, each system was simulated at least ten times. Single runs lasted ( 2 ÷ 4 ) × 10 6 Monte Carlo (MC) cycles (depending on system sizes). The simulations of hard sphere system took 12 × 10 6 MC cycles. First, 10 6 cycles were treated as a period in which the system reaches the state of thermodynamic equilibrium.

3. Results and Discussion

Introducing a nanolayer into a perfect crystal with the face-centred cubic structure leads to change in the symmetry of the studied system. These changes were noticed in recent computer simulations [66]. A transition from fcc structure to the tetragonal one was observed. This is manifested by the change of the shape of simulation box from a cuboid to a rhombohedron. The matrix describing the simulation box of the system with nanolayers in the (111) plane has the form
h = h x x h x y h x y h x y h x x h x y h x y h x y h x x .
In this case, using Monte Carlo simulations utilizing the Parrinello–Rahman method, one obtains the following form the elastic compliance tensor (given in the matrix form in the Voigt notation [70]):
S = S 11 S 12 S 12 S 14 S 15 S 15 · S 11 S 12 S 15 S 14 S 15 · · S 11 S 15 S 15 S 14 · · · S 44 0 0 · · · · S 44 0 · · · · · S 44 .
The above form is the consequence of the symmetry of the S i j k l tensor. To determine the symmetry of the studied system uniquely, we bring the box matrix to its principal axes h i j = R i p R j r h p r using the following rotation matrix
R = 1 6 3 3 3 3 2 3 3 3 3 3 2 3 2 3 2 3 2 3 .
In the rotated coordinate system, the h and S matrices can be written, respectively, as
h = h x x 0 0 0 h x x 0 0 0 h z z ,
S = S 11 S 12 S 13 S 14 S 15 0 · S 11 S 13 S 14 S 15 0 · · S 33 0 0 0 · · · S 44 0 2 S 15 · · · · S 44 2 S 14 · · · · · 2 ( S 11 S 12 ) ,
where
S 11 = 1 2 ( S 11 + S 12 4 S 15 ) + S 44 ,
S 12 = 1 6 ( S 11 + 5 S 12 2 ( 4 S 14 + 2 S 15 + S 44 ) ) ,
S 13 = 1 3 ( S 11 + 2 S 12 + S 14 + 2 S 15 2 S 44 ) ,
S 14 = 1 3 ( S 11 S 12 + S 14 S 15 2 S 44 ) ,
S 44 = 4 3 ( S 11 S 12 2 S 14 + 2 S 15 + S 44 ) .
The above form of the S matrix corresponds to the system of trigonal symmetry [70]. This clearly indicates that the systems with nanolayers in the (111) plane have trigonal symmetry.
Figure 2 shows the elements of elastic compliance matrix as a function of concentration for all studied systems. It can be seen that the values of elastic compliances ( S 11 * , S 12 * , S 44 * ) of systems in which nanolayers contain Yukawa multilayers (2Y–5Y) are, respectively, similar and the remaining elastic compliances ( S 14 * , S 15 * ) present the same (continuous) trend as functions of concentration. However, at least in the concentration dependence of S 14 * , a discontinuity at the transition to the systems with Yukawa monolayers (1Y) is observed. For the latter systems, some components of elastic compliances ( S 11 * , S 14 * , S 15 * ) differ notably when compared with corresponding values of the same elements of double-layer Yukawa systems (2Y). This difference may be due to the lack of strong repulsion between the adjacent Yukawa layers inside the nanolayer in 1Y systems, which presents in multilayer systems (2Y–5Y). This has a significant effect on the auxetic properties of the 1Y systems.
In Figure 3, the Poisson’s ratio in the main crystallographic directions has been shown. It can be observed here that the dependencies of Poisson’s ratio on c for Yukawa multilayer systems are very similar between each other. However, for certain crystallographic directions ( [ 110 ] [ 001 ] , [ 111 ] [ 1 1 ¯ 0 ] , [ 111 ] [ 11 2 ¯ ] , [ 11 1 ¯ ] [ 1 1 ¯ 0 ] , [ 11 1 ¯ ] [ 112 ] ), in the case of 1Y systems, we observe a qualitatively different character of the dependence of Poisson’s ratio from the concentration (c). Moreover, Poisson’s ratio in the [ 11 1 ¯ ] [ 112 ] crystallographic direction exhibits negative values (see Figure 3d). This means that this direction is an auxetic one. It is worth noting that this happens while maintaining the auxeticity in the crystallographic direction of [ 110 ] [ 1 1 ¯ 0 ] (see Figure 3b).
The dependencies of the minimum and maximum Poisson’s ratio values of the studied systems on the concentration are presented in Figure 4. Here, one can see a qualitatively different dependence of the Poisson’s ratio on the concentration in systems with Yukawa monolayer and systems with Yukawa multilayers. For multilayered Yukawa systems, the concentration weakly affects the extreme values of the Poisson’s ratio. Moreover, the extreme values of the Poisson’s ratio are barely dependent on the number of Yukawa layers in the nanolayers (see Figure 4). This is reflected in the results obtained for multilayered Yukawa systems, which are presented in Figure 4 in the range of concentrations from 0 % to 80 % . However, in the case of monolayers of Yukawa particles in the system, one can find in Figure 4 that the dependence of the Poisson’s ratio on the concentration is qualitatively different from those obtained for multilayered Yukawa systems. It is important to note that minimum value of Poisson’s ratio in the monolayer Yukawa system (1Y) at c = 66.67 % reaches the value −0.39(3) in the [ 2 1 ¯ 2 ] [ 141 ] crystallographic direction.
In Figure 4 (see inserts), some structures studied for a few concentrations are presented. They illustrate extreme values of Poisson’s ratio in all crystallographic directions in the three-dimensional plot. If we assume that, in the plot, a vector from the origin of the coordinate system points to some point of three-dimensional surface, then a direction of the vector represents the direction of the applied stress ( n ^ ), and its modulus has the extreme value of Poisson’s ratio in that direction. Thus, each point on the surface of the plot corresponds to the maximal value of Poisson’s ratio (top row of inserts in Figure 4) or absolute value of minimal negative Poisson’s ratio (bottom row of inserts in Figure 4) for the direction in which the stress is applied. This gives a qualitative picture of changes in the extreme values of Poisson’s ratio in all crystallographic directions due to the presence of nanolayers in the studied system. In the case of minimal Poisson’s ratio (Figure 4 bottom row), the increase in volume of the figure signals the enhancement of the auxetic properties in the system—this is observed for the Yukawa monolayers (shaded area in Figure 4). On the other hand, it can be seen that, for Yukawa multilayer systems, there are crystallographic directions in which the value of Poisson’s ratio changes with concentration (see Figure 3), although the extreme values of Poisson’s of the entire system, obtained from extremal values in all crystallographic directions, are almost constant.
Besides all this, it follows from Figure 4 that not only does the concentration of particles of nanolayer affect the elastic properties of the system, but also the distance between the layers in the nanolayer. In particular, an absence of inter-layers Yukawa interactions leads to sharp changes of the elastic properties of the system (see Figure 4). In order to quantitatively describe the change of the auxeticity in studied systems, recently, the degree of axeticity has been defined as [71]
χ = 3 A 4 π 3 ,
where A = 0 2 π d ϕ 0 π sin θ d θ 0 R ( θ , ϕ ) r 2 d r is the integral of absolute values of mean negative Poisson’s ratios and R ( θ , ϕ ) = 1 2 π 0 π ν n ( θ , ϕ ) ( α ) ν n ( θ , ϕ ) ( α ) d α is the average over negative values of Poisson’s ratio in the direction of n ^ . Here, the versor n ^ is determined in spherical coordinates. For more details, see Reference [71].
In Figure 5, the degree of auxeticity of studied systems as a function of concentration is presented. A qualitative difference in elastic properties of monolayered Yukawa systems and multilayered Yukawa systems is observed. In the case of monolayered Yukawa systems, an essential enhancement of auxeticity was found. However, the increase of concentration of nanolayer particles is causing a weakening of auxetic properties of the system.

4. Conclusions

This work is a continuation of the study on the effect of structural modifications on elastic properties, and particularly on the auxetic properties of the model of nanocomposites. Comprehensive Monte Carlo simulations revealed that the structural modification of the Yukawa system in the form of nanolayers in the (111) crystallographic plane can lead both to strengthening as well as weakening the auxetic properties of the system. Reduction of auxetic properties is achieved by increasing the concentration of nanolayer particles in multilayer Yukawa systems, whereas monolayered Yukawa systems exhibit the enhancement of auxetic properties. The latter is the appearance of a new auxetic direction ( [ 11 1 ¯ ] [ 112 ] ) while preserving other, already existing, auxetic directions in the system.
The extreme values of the Poisson’s ratio of the studied systems were also determined. The presentation of them in the form of a three-dimensional plot is suitable for a quick qualitative assessment of changes in the auxetic properties of the whole system caused by the modification of the structure of the studied system. On the other hand, a use of degree of auxeticity gives a quantitative description of auxetic properties of the system under study.
The results presented in this paper demonstrate another (simple and efficient) way of controlling the elastic properties of materials and show the consequences of introducing the nanolayers in the (111) crystallographic plane of given crystal. This information may be useful for the construction of nanocomposites, and may indicate directions for further research on materials with desired elastic properties.

Acknowledgments

The calculations were partially performed at Poznań Supercomputing and Networking Center (PCSS). This work was partially supported by the (Polish) National Centre for Science under the grant NCN 2012/05/N/ST5/01476.

Author Contributions

Konstantin V. Tretiakov conceived and designed the model for simulations. Paweł M. Pigłowski preformed the computer simulations. Konstantin V. Tretiakov, Paweł M. Pigłowski, Artur A. Poźniak, Jakub W. Narojczyk and Krzysztof W. Wojciechowski analyzed the data. Konstantin V. Tretiakov, Jakub W. Narojczyk, and Krzysztof W. Wojciechowski wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
HCRYPhard-core repulsive Yukawa potential
HShard spheres potential

References

  1. Van Blaaderen, A.; Ruel, R.; Wiltzius, P. Template-directed colloidal crystallization. Nature 1997, 385, 321–324. [Google Scholar] [CrossRef]
  2. Van Blaaderen, A. Colloids under external control. MRS Bull. 2004, 29, 85–90. [Google Scholar] [CrossRef]
  3. Lee, S.H.; Liddell, C.M. Anisotropic Magnetic Colloids: A Strategy to Form Complex Structures Using Nonspherical Building Blocks. Small 2009, 5, 1957–1962. [Google Scholar] [CrossRef] [PubMed]
  4. Demirors, A.F.; Johnson, P.M.; van Kats, C.M.; van Blaaderen, A.; Imhof, A. Directed Self-Assembly of Colloidal Dumbbells with an Electric Field. Langmuir 2010, 26, 14466–14471. [Google Scholar] [CrossRef] [PubMed]
  5. Sacanna, S.; Rossi, L.; Pine, D.J. Magnetic Click Colloidal Assembly. J. Am. Chem. Soc. 2012, 134, 6112–6115. [Google Scholar] [CrossRef] [PubMed]
  6. Smallenburg, F.; Vutukuri, H.R.; Imhof, A.; van Blaaderen, A.; Dijkstra, M. Self-assembly of colloidal particles into strings in a homogeneous external electric or magnetic field. J. Phys.-Condes. Matter 2012, 24, 464113. [Google Scholar] [CrossRef] [PubMed]
  7. Demirors, A.F.; Pillai, P.P.; Kowalczyk, B.; Grzybowski, B.A. Colloidal assembly directed by virtual magnetic moulds. Nature 2013, 503, 99–103. [Google Scholar] [CrossRef] [PubMed]
  8. Peng, B.; Smallenburg, F.; Imhof, A.; Dijkstra, M.; van Blaaderen, A. Colloidal Clusters by Using Emulsions and Dumbbell-Shaped Particles: Experiments and Simulations. Angew. Chem.-Int. Ed. 2013, 52, 6709–6712. [Google Scholar] [CrossRef] [PubMed]
  9. Bakker, H.E.; Dussi, S.; Droste, B.L.; Besseling, T.H.; Kennedy, C.L.; Wiegant, E.I.; Liu, B.; Imhof, A.; Dijkstra, M.; van Blaaderen, A. Phase diagram of binary colloidal rod-sphere mixtures from a 3D real-space analysis of sedimentation-diffusion equilibria. Soft Matter 2016, 12, 9238–9245. [Google Scholar] [CrossRef] [PubMed]
  10. Alexeev, V.L.; Finegold, D.N.; Asher, S.A. Photonic Crystal Glucose-Sensing Material for Noninvasive Monitoring of Glucose in Tear Fluid. Clin. Chem. 2004, 50, 2353–2360. [Google Scholar] [CrossRef] [PubMed]
  11. Talapin, D.V.; Lee, J.S.; Kovalenko, M.V.; Shevchenko, E.V. Prospects of colloidal nanocrystals for electronic and optoelectronic applications. Chem. Rev. 2010, 110, 389–458. [Google Scholar] [CrossRef] [PubMed]
  12. Galisteo-López, J.; Ibisate, M.; Sapienza, R.; Froufe-Peérez, L.; Blanco, A.; López, C. Self-Assembled Photonic Structures. Adv. Mater. 2011, 23, 30–69. [Google Scholar] [CrossRef] [PubMed]
  13. Azhar, F.E.; Baus, M.; Ryckaert, J.P. Line of triple points for the hard-core Yukawa model: A computer simulation study. J. Chem. Phys. 2000, 112, 5121–5126. [Google Scholar] [CrossRef]
  14. Auer, S.; Frenkel, D. Crystallization of weakly charged colloidal spheres: A numerical study. J. Phys.-Condes. Matter 2002, 14, 7667–7680. [Google Scholar] [CrossRef]
  15. Hynninen, A.P.; Dijkstra, M. Phase diagrams of hard-core repulsive Yukawa particles. Phys. Rev. E 2003, 68, 021407. [Google Scholar] [CrossRef] [PubMed]
  16. Colombo, J.; Dijkstra, M. Effect of quenched size polydispersity on the fluid-solid transition in charged colloidal suspensions. J. Chem. Phys. 2011, 134, 154504. [Google Scholar] [CrossRef] [PubMed]
  17. Heinen, M.; Holmqvist, P.; Banchio, A.J.; Nagele, G. Pair structure of the hard-sphere Yukawa fluid: An improved analytic method versus simulations, Rogers-Young scheme, and experiment. J. Chem. Phys. 2011, 134, 044532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Van der Linden, M.N.; van Blaaderen, A.; Dijkstra, M. Effect of size polydispersity on the crystal-fluid and crystal-glass transition in hard-core repulsive Yukawa systems. J. Chem. Phys. 2013, 138, 114903. [Google Scholar] [CrossRef] [PubMed]
  19. Tretiakov, K.V.; Wojciechowski, K.W. Partially auxetic behavior in fcc crystals of hard-core repulsive Yukawa particles. Phys. Status Solidi B 2014, 251, 383–387. [Google Scholar] [CrossRef]
  20. Landau, L.D.; Lifshitz, E.M. Theory of Elasticity, 3rd ed.; Pergamon Press: London, UK, 1986. [Google Scholar]
  21. Evans, K.E. Auxetic polymers: A new range of materials. Endeavour 1991, 15, 170–174. [Google Scholar] [CrossRef]
  22. Brańka, A.C.; Heyes, D.M.; Wojciechowski, K.W. Auxeticity of cubic materials under pressure. Phys. Status Solidi B 2011, 248, 96–104. [Google Scholar] [CrossRef]
  23. Lakes, R.S. Foam structures with a negative Poisson’s ratio. Science 1987, 235, 1038–1040. [Google Scholar] [CrossRef] [PubMed]
  24. Wojciechowski, K.W. Two-dimensional isotropic model with a negative Poisson ratio. Phys. Lett. A 1989, 137, 60–64. [Google Scholar] [CrossRef]
  25. Evans, K.E.; Alderson, A. Auxetic Materials: Functional Materials and Structures from Lateral Thinking! Adv. Mater. 2000, 12, 617–628. [Google Scholar] [CrossRef]
  26. Alderson, A.; Alderson, K.L. Auxetic materials. Proc. Inst. Mech. Eng. Part G-J. Aerosp. Eng. 2007, 221, 565–575. [Google Scholar] [CrossRef]
  27. Prawoto, Y. Seeing auxetic materials from the mechanics point of view: A structural review on the negative Poisson’s ratio. Comput. Mater. Sci. 2012, 58, 140–153. [Google Scholar] [CrossRef]
  28. Lim, T.C. Auxetic Materials and Structures; Springer: Singapore, 2015. [Google Scholar]
  29. Almgren, R.F. An isotropic three-dimensional structure with Poisson’s ratio = −1. J. Elast. 1985, 15, 427–430. [Google Scholar]
  30. Lakes, R.S. Advances in negative Poisson’s ratio materials. Adv. Mater. 1993, 5, 293–296. [Google Scholar] [CrossRef]
  31. Ishibashi, Y.; Iwata, M. A microscopic model of a negative Poisson’s ratio in some crystals. J. Phys. Soc. Jpn. 2000, 69, 2702–2703. [Google Scholar] [CrossRef]
  32. Wojciechowski, K.W. Non-chiral, molecular model of negative Poisson’s ratio in two dimensions. J. Phys. A: Math. Gen. 2003, 36, 11765–11778. [Google Scholar] [CrossRef]
  33. Wojciechowski, K.W. Poisson’s ratio of anisotropic systems. Comp. Meth. Sci. Technol. 2005, 11, 73–79. [Google Scholar] [CrossRef]
  34. Narojczyk, J.W.; Alderson, A.; Imre, A.R.; Scarpa, F.; Wojciechowski, K.W. Negative Poisson’s ratio behavior in the planar model of asymmetric trimers at zero temperature. J. Non-Cryst. Solids 2008, 354, 4242–4248. [Google Scholar] [CrossRef]
  35. Pasternak, E.; Shufrin, I.; Dyskin, A.V. Planar isotropic structures with negative Poisson’s ratio. Int. J. Solid Struct. 2012, 49, 2239–2253. [Google Scholar]
  36. Pasternak, E.; Shufrin, I.; Dyskin, A.V. Negative Poisson’s ratio in hollow sphere materials. Int. J. Solid Struct. 2015, 54, 192–214. [Google Scholar]
  37. Pasternak, E.; Shufrin, I.; Dyskin, A.V. Thermal stresses in hybrid materials with auxetic inclusions. Comp. Struct. 2016, 138, 313–321. [Google Scholar] [CrossRef]
  38. Schwerdtfeger, J.; Wein, F.; Leugering, G.; Singer, R.F.; Körner, C.; Stingl, M.; Schury, F. Design of auxetic structures via mathematical optimization. Adv. Mater. 2011, 23, 2650–2654. [Google Scholar] [CrossRef] [PubMed]
  39. Baughman, R.H.; Shacklette, J.M.; Zakhidov, A.A.; Stafstrom, S. Negative Poisson’s ratios as a common feature of cubic metals. Nature 1998, 392, 362–365. [Google Scholar] [CrossRef]
  40. Wojciechowski, K.W.; Tretiakov, K.V.; Kowalik, M. Elastic properties of dense solid phases of hard cyclic pentamers and heptamers in two dimensions. Phys. Rev. E 2003, 67, 036121. [Google Scholar] [CrossRef] [PubMed]
  41. Tretiakov, K.V. Negative Poisson’s ratio of two-dimensional hard cyclic tetramers. J. Non-Cryst. Solids 2009, 355, 1435–1438. [Google Scholar] [CrossRef]
  42. Wojciechowski, K.W. Negative Poisson ratios at negative pressures. Mol. Phys. Rep. 1995, 10, 129–136. [Google Scholar]
  43. Rechtsman, M.C.; Stillinger, F.H.; Torquato, S. Negative Poisson’s Ratio Materials via Isotropic Interactions. Phys. Rev. Lett. 2008, 101, 085501. [Google Scholar] [CrossRef] [PubMed]
  44. Yeganeh-Haeri, A.; Weidner, D.J.; Parise, J.B. Elasticity of α-Cristobalite: A Silicon Dioxide with a Negative Poisson’s Ratio. Science 1992, 257, 650–652. [Google Scholar] [CrossRef] [PubMed]
  45. Bezazi, A.; Scarpa, F. Tensile fatigue of conventional and negative Poisson’s ratio open cell PU foams. Int. J. Fatigue 2009, 31, 488–494. [Google Scholar] [CrossRef]
  46. Gatt, R.; Caruana-Gauci, R.; Attard, D.; Casha, A.R.; Wolak, W.; Dudek, K.; Mizzi, L.; Grima, J.N. On the properties of real finite-sized planar and tubular stent-like auxetic structures. Phys. Status Solidi B 2014, 251, 321–327. [Google Scholar] [CrossRef]
  47. Gatt, R.; Mizzi, L.; Azzopardi, J.I.; Azzopardi, K.M.; Attard, D.; Casha, A.; Briffa, J.; Grima, J.N. Hierarchical Auxetic Mechanical Metamaterials. Sci. Rep. 2015, 5, 8395. [Google Scholar] [CrossRef] [PubMed]
  48. Shufrin, I.; Pasternak, E.; Dyskin, A.V. Hybrid materials with negative Poisson’s ratio inclusions. Int. J. Eng. Sci. 2015, 89, 100–120. [Google Scholar] [CrossRef]
  49. Slann, A.; White, W.; Scarpa, F.; Boba, K.; Farrow, I. Cellular plates with auxetic rectangular perforations. Phys. Status Solidi B 2015, 252, 1533–1539. [Google Scholar] [CrossRef]
  50. Verma, P.; Shofner, M.L.; Lin, A.; Wagner, K.B.; Griffin, A.C. Inducing out-of-plane auxetic behavior in needle-punched nonwovens. Phys. Status Solidi B 2015, 252, 1455–1464. [Google Scholar] [CrossRef]
  51. Pasternak, E.; Dyskin, A.V. Materials and structures with macroscopic negative Poisson’s ratio. Int. J. Eng. Sci. 2012, 52, 103–114. [Google Scholar] [CrossRef]
  52. Zaitsev, V.; Radostin, A.; Pasternak, E.; Dyskin, A.V. Extracting real-crack properties from non-linear elastic behaviour of rocks: Abundance of cracks with dominating normal compliance and rocks with negative Poisson ratios. Nonlinear Process. Geophys. 2017, 24, 543–551. [Google Scholar] [CrossRef]
  53. Alderson, K.L.; Simkins, V.R.; Coenen, V.L.; Davies, P.J.; Alderson, A.; Evans, K.E. How to make auxetic fibre reinforced composites. Phys. Status Solidi B 2005, 252, 509–518. [Google Scholar] [CrossRef]
  54. Airoldi, A.; Bettini, P.; Panichelli, P.; Oktem, M.F.; Sala, G. Chiral topologies for composite morphing structures? Part I: Development of a chiral rib for deformable airfoils. Phys. Status Solidi B 2015, 252, 1435–1445. [Google Scholar] [CrossRef]
  55. Hou, X.; Hu, H. A novel 3D composite structure with tunable Poisson’s ratio and stiffness. Phys. Status Solidi B 2015, 252, 1565–1574. [Google Scholar] [CrossRef]
  56. Hoover, W.G.; Hoover, C.G. Searching for auxetics with DYNA3D and ParaDyn. Phys. Status Solidi B 2005, 242, 585–594. [Google Scholar] [CrossRef]
  57. Grima, J.N.; Zammit, V.; Gatt, R.; Alderson, A.; Evans, K.E. Auxetic behaviour from rotating semi-rigid units. Phys. Status Solidi B 2007, 244, 866–882. [Google Scholar] [CrossRef]
  58. Ravirala, N.; Alderson, A.; Alderson, K.L. Interlocking hexagons model for auxetic behaviour. J. Mater. Sci. 2007, 42, 7433–7445. [Google Scholar] [CrossRef]
  59. Attard, D.; Grima, J.N. Auxetic behaviour from rotating rhombi. Phys. Status Solidi B 2008, 245, 2395–2404. [Google Scholar] [CrossRef]
  60. Grima, J.N.; Gatt, R.; Farrugia, P.S. On the properties of auxetic meta-tetrachiral structures. Phys. Status Solidi B 2008, 245, 511–520. [Google Scholar] [CrossRef]
  61. Strek, T.; Jopek, H.; Maruszewski, B.T.; Nienartowicz, M. Computational analysis of sandwich-structured composites with an auxetic phase. Phys. Status Solidi B 2014, 251, 354–366. [Google Scholar] [CrossRef]
  62. Strek, T.; Jopek, H.; Nienartowicz, M. Dynamic response of sandwich panels with auxetic cores. Phys. Status Solidi B 2015, 252, 1540–1550. [Google Scholar] [CrossRef]
  63. Shufrin, I.; Pasternak, E.; Dyskin, A.V. Deformation analysis of reinforced-core auxetic assemblies by close-range photogrammetry. Phys. Status Solidi B 2016, 253, 1342–1358. [Google Scholar] [CrossRef]
  64. Dudek, K.K.; Gatt, R.; Mizzi, L.; Dudek, M.R.; Attard, D.; Evans, K.E.; Grima, J.N. On the dynamics and control of mechanical properties of hierarchical rotating rigid unit auxetics. Sci. Rep. 2017, 7, 46529. [Google Scholar] [CrossRef] [PubMed]
  65. Tretiakov, K.V.; Piglowski, P.M.; Hyzorek, K.; Wojciechowski, K.W. Enhanced auxeticity in Yukawa systems due to introduction of nanochannels in [001]-direction. Smart Mater. Struct. 2016, 25, 054007. [Google Scholar] [CrossRef]
  66. Piglowski, P.M.; Wojciechowski, K.W.; Tretiakov, K.V. Partial auxeticity induced by nanoslits in the Yukawa crystal. Phys. Status Solidi-Rapid Res. Lett. 2016, 10, 566–569. [Google Scholar] [CrossRef]
  67. Parrinello, M.; Rahman, A. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52, 7182–7190. [Google Scholar] [CrossRef]
  68. Parrinello, M.; Rahman, A. Strain fluctuations and elastic constants. J. Chem. Phys. 1982, 76, 2662–2666. [Google Scholar] [CrossRef]
  69. Tokmakova, S.P. Stereographic projections of Poisson’s ratio in auxetic crystals. Phys. Status Solidi B 2005, 242, 721–729. [Google Scholar] [CrossRef]
  70. Nye, J.F. Physical Properties of Crystalls, Their Representation by Tensors and Matrices; Clarendon Press: Oxford, UK, 1957. [Google Scholar]
  71. Piglowski, P.M.; Narojczyk, J.W.; Wojciechowski, K.W.; Tretiakov, K.V. Auxeticity enhancement due to size polydispersity in fcc crystals of hard-core repulsive Yukawa particles. Soft Matter 2017, 13, 7916–7921. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Typical structure of a studied crystal with nanolayers parallel to the (111) crystallographic plane. Yukawa’s particles are denoted with a green color and the hard spheres are marked with a red color. (a) a system with nanolayers consisting of a monolayer of hard spheres can be also seen as multilayer Yukawa system. This system is denoted in the text as 1HS or 6Y; (b) a system with nanolayers consisting of double layer of hard spheres; (c) a system with nanolayers consisting of four layers of hard spheres; (d) a system with nanolayers consisting of a monolayer of particles that interact via Yukawa potential. For more details, see Table 1.
Figure 1. Typical structure of a studied crystal with nanolayers parallel to the (111) crystallographic plane. Yukawa’s particles are denoted with a green color and the hard spheres are marked with a red color. (a) a system with nanolayers consisting of a monolayer of hard spheres can be also seen as multilayer Yukawa system. This system is denoted in the text as 1HS or 6Y; (b) a system with nanolayers consisting of double layer of hard spheres; (c) a system with nanolayers consisting of four layers of hard spheres; (d) a system with nanolayers consisting of a monolayer of particles that interact via Yukawa potential. For more details, see Table 1.
Materials 10 01338 g001
Figure 2. Elastic compliances versus concentration involved in particles of nanolayers in studied systems. Colors indicate the type of the system depending on the number of layers of appropriate particles in a nanolayer. 1HS (black) is the system with hard spheres monolayer. 1Y (red) is the system with monolayers of Yukawa particles. 2Y (light brown) is the system with double layers of Yukawa particles. 3Y (blue) is the system with triple layers of Yukawa particles. 4Y (magenta) is the system with four-layers of Yukawa particles. 5Y (green) is the system with five-layers of Yukawa particles. In addition, miniature images of studied systems are placed in the graph so that the centres of the structures coincided with the corresponding concentrations. The systems in the figure are grouped according to the numbers of layers in nanolayers. Miniatures of structures represent the studied systems seen from the [ 1 ¯ 10 ] -direction. The color convention of data presentation from this figure is maintained in all following figures. Lines are drawn to guide the eyes.
Figure 2. Elastic compliances versus concentration involved in particles of nanolayers in studied systems. Colors indicate the type of the system depending on the number of layers of appropriate particles in a nanolayer. 1HS (black) is the system with hard spheres monolayer. 1Y (red) is the system with monolayers of Yukawa particles. 2Y (light brown) is the system with double layers of Yukawa particles. 3Y (blue) is the system with triple layers of Yukawa particles. 4Y (magenta) is the system with four-layers of Yukawa particles. 5Y (green) is the system with five-layers of Yukawa particles. In addition, miniature images of studied systems are placed in the graph so that the centres of the structures coincided with the corresponding concentrations. The systems in the figure are grouped according to the numbers of layers in nanolayers. Miniatures of structures represent the studied systems seen from the [ 1 ¯ 10 ] -direction. The color convention of data presentation from this figure is maintained in all following figures. Lines are drawn to guide the eyes.
Materials 10 01338 g002
Figure 3. Poisson’s ratio in the main crystallographic directions versus concentration of particles of nanolayers in the studied system. The deformation of the system is applied respectively in the direction: (a) [ 100 ] , (b) [ 110 ] , (c) [ 111 ] , (d) [ 11 1 ¯ ] , and (e) [ 1 1 ¯ 0 ] . The meaning of colors is the same as in Figure 2.
Figure 3. Poisson’s ratio in the main crystallographic directions versus concentration of particles of nanolayers in the studied system. The deformation of the system is applied respectively in the direction: (a) [ 100 ] , (b) [ 110 ] , (c) [ 111 ] , (d) [ 11 1 ¯ ] , and (e) [ 1 1 ¯ 0 ] . The meaning of colors is the same as in Figure 2.
Materials 10 01338 g003aMaterials 10 01338 g003b
Figure 4. Maximal (squares) and minimal (circles) Poisson’s ratio taken from all crystallographic directions as a function of concentration of particles of nanoslit in studied system. All miniature images are placed in the graph so that the center of it coincided with the corresponding concentration (except pure Yukawa and HS systems). Upper row of miniature images correspond to the maximal values of Poisson’s ratio in all crystallographic directions. Lower row of miniature images describe the absolute value of minimal negative Poisson’s ratio in all crystallographic directions of studied systems. The studied structures of trigonal symmetry and crystallographic axes of three-dimensional plots have the same orientation as in Figure 1. The shaded area denotes the data on the graph concerned with the system with the nanolayer consisting of a monolayer of particles that interact via Yukawa potential. Lines are drawn to guide the eyes.
Figure 4. Maximal (squares) and minimal (circles) Poisson’s ratio taken from all crystallographic directions as a function of concentration of particles of nanoslit in studied system. All miniature images are placed in the graph so that the center of it coincided with the corresponding concentration (except pure Yukawa and HS systems). Upper row of miniature images correspond to the maximal values of Poisson’s ratio in all crystallographic directions. Lower row of miniature images describe the absolute value of minimal negative Poisson’s ratio in all crystallographic directions of studied systems. The studied structures of trigonal symmetry and crystallographic axes of three-dimensional plots have the same orientation as in Figure 1. The shaded area denotes the data on the graph concerned with the system with the nanolayer consisting of a monolayer of particles that interact via Yukawa potential. Lines are drawn to guide the eyes.
Materials 10 01338 g004
Figure 5. Degree of auxeticity as a function of concentration c. χ = 0 corresponds to the case of non-auxetic system. Lines are drawn to guide the eyes.
Figure 5. Degree of auxeticity as a function of concentration c. χ = 0 corresponds to the case of non-auxetic system. Lines are drawn to guide the eyes.
Materials 10 01338 g005
Table 1. Examples of structural modifications of studied systems. Each structure was based on the face-centered cubic (fcc) lattice which unit cell contains four atoms. n is the number of unit cells in direction’s x , y , z . N = 4 n 3 is the number of particles in the system. N HS is the number of particles in nanolayers. c is the concentration of nanolayers particles in the system. ρ = N / V is the density of the system. n Y is the number of layers of Yukawa particles in a nanolayer. n HS is the number of layers of hard spheres in a nanolayer. The description column holds the abbreviations that indicate the number of layers in a nanolayer, and are used in figures and in the text to refer to those systems.
Table 1. Examples of structural modifications of studied systems. Each structure was based on the face-centered cubic (fcc) lattice which unit cell contains four atoms. n is the number of unit cells in direction’s x , y , z . N = 4 n 3 is the number of particles in the system. N HS is the number of particles in nanolayers. c is the concentration of nanolayers particles in the system. ρ = N / V is the density of the system. n Y is the number of layers of Yukawa particles in a nanolayer. n HS is the number of layers of hard spheres in a nanolayer. The description column holds the abbreviations that indicate the number of layers in a nanolayer, and are used in figures and in the text to refer to those systems.
nN N HS c [%] ρ n Y n HS Description
7137219614.29 0.9940 ( 4 ) 616Y,1HS
686414416.67 1.0042 ( 9 ) 515Y,1HS
7137239228.57 1.0358 ( 7 ) 525Y,2HS
550010020.00 1.0190 ( 7 ) 414Y,1HS
686428833.33 1.0554 ( 4 ) 424Y,2HS
7137258842.86 1.0854 ( 9 ) 434Y,3HS
42566425.00 1.0422 ( 2 ) 313Y,1HS
550020040.00 1.0848 ( 1 ) 323Y,2HS
686443250.00 1.1189 ( 3 ) 333Y,3HS
7137278457.14 1.1472 ( 8 ) 343Y,4HS
31083633.33 1.0840 ( 4 ) 212Y,1HS
425612850.00 1.1343 ( 2 ) 222Y,2HS
550030060.00 1.1729 ( 8 ) 232Y,3HS
686457666.67 1.2036 ( 3 ) 242Y,4HS
7137298071.43 1.2281 ( 8 ) 252Y,5HS
82048153675.00 1.2476 ( 1 ) 262Y,6HS
104000320080.00 1.2753 ( 5 ) 282Y,8HS
31087266.67 1.2406 ( 4 ) 121Y,2HS
425619275.00 1.2789 ( 1 ) 131Y,3HS
550040080.00 1.3021 ( 1 ) 141Y,4HS
686472083.33 1.3159 ( 9 ) 151Y,5HS
71372117685.71 1.3247 ( 4 ) 161Y,6HS

Share and Cite

MDPI and ACS Style

Pigłowski, P.M.; Narojczyk, J.W.; Poźniak, A.A.; Wojciechowski, K.W.; Tretiakov, K.V. Auxeticity of Yukawa Systems with Nanolayers in the (111) Crystallographic Plane. Materials 2017, 10, 1338. https://doi.org/10.3390/ma10111338

AMA Style

Pigłowski PM, Narojczyk JW, Poźniak AA, Wojciechowski KW, Tretiakov KV. Auxeticity of Yukawa Systems with Nanolayers in the (111) Crystallographic Plane. Materials. 2017; 10(11):1338. https://doi.org/10.3390/ma10111338

Chicago/Turabian Style

Pigłowski, Paweł M., Jakub W. Narojczyk, Artur A. Poźniak, Krzysztof W. Wojciechowski, and Konstantin V. Tretiakov. 2017. "Auxeticity of Yukawa Systems with Nanolayers in the (111) Crystallographic Plane" Materials 10, no. 11: 1338. https://doi.org/10.3390/ma10111338

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