Skip to main content
Log in

On the Ternary Ohta–Kawasaki Free Energy and Its One-dimensional Global Minimizers

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We study the ternary Ohta–Kawasaki free energy that has been used to model triblock copolymer systems. Its one-dimensional global minimizers are conjectured to have cyclic patterns. However, some physical experiments and computer simulations found triblock copolymers forming noncyclic lamellar patterns. In this work, by comparing the free energies of the cyclic pattern and some noncyclic candidates, we show that the conjecture does not hold for some choices of parameters. Our results suggest that even in one dimension, the global minimizers may take on very different patterns in different parameter regimes. To unify the existing choices of the long-range coefficient matrix, we present a reformulation of the long-range term using a generalized charge interpretation and thereby propose conditions on the matrix in order for the global minimizers to reproduce physically relevant nanostructures of block copolymers.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Data Availability

All data generated or analyzed during this study are included in the supplementary material.

References

  • Alama, S., et al.: Periodic minimizers of a ternary non-local isoperimetric problem. arXiv preprint arXiv:1912.08971. To appear in Indiana U. Math, Jour (2019)

  • Alama, S., et al.: Droplet phase in a nonlocal isoperimetric problem under confinement. Commun. Pure Appl. Anal. 19(1), 175 (2020)

    MathSciNet  MATH  Google Scholar 

  • Alberti, G., Choksi, R., Otto, F.: Uniform energy distribution for an isoperimetric problem with long-range interactions. J. Am. Math. Soc. 22(2), 569–605 (2009)

    MathSciNet  MATH  Google Scholar 

  • Bailey, T.S., et al.: A noncubic triply periodic network morphology in poly (isoprene-b-styrene-b-ethylene oxide) triblock copolymers. Macromolecules 35(18), 7007–7017 (2002)

    Google Scholar 

  • Bailey, T.S., Pham, H.D., Bates, F.S.: Morphological behavior bridging the symmetric AB and ABC states in the poly (styrene-b-isoprene-b-ethylene oxide) triblock copolymer system. Macromolecules 34(20), 6994–7008 (2001)

    Google Scholar 

  • Bates, F.S., Fredrickson, GH.: Block copolymers-designer soft materials. Phys. Today 52 (2000)

  • Bates, F.S., Fredrickson, G.H.: Block copolymer thermodynamics: theory and experiment. Annu. Rev. Phys. Chem. 41(1), 525–557 (1990)

    Google Scholar 

  • Birdi, K.S.: Introduction to surface and colloid chemistry, in Handbook of Surface and Colloid Chemistry (4th ed.). CRC Press (2015): 1–144

  • Burchard, A., Choksi, R., Topaloglu, I.: Nonlocal shape optimization via interactions of attractive and repulsive potentials. arXiv preprint arXiv:1512.07282 (2015)

  • Cahn, J.W., Hilliard, J.E.: Free energy of a nonuniform system. I. Interfacial free energy. J. Chem. Phys. 28(2), 258–267 (1958)

    MATH  Google Scholar 

  • Carazzato, D., Fusco, N., Pratelli, A.: Minimality of balls in the small volume regime for a general Gamow-type functional. Adv. Calc. Var. (2021)

  • Chan, H., Nejad, M.J., Wei, J.: Lamellar phase solutions for diblock copolymers with nonlocal diffusions. Phys. D 388, 22–32 (2019)

    MathSciNet  MATH  Google Scholar 

  • Chang, AB., Bates, FS.: The ABCs of block polymers. pp 2765–2768 (2020)

  • Choksi, R., Peletier, M.A.: Small volume fraction limit of the diblock copolymer problem: I. Sharp-interface functional. SIAM J. Math. Anal. 42(3), 1334–1370 (2010)

    MathSciNet  MATH  Google Scholar 

  • Choksi, R., Ren, X.: On the derivation of a density functional theory for microphase separation of diblock copolymers. J. Stat. Phys. 113(1), 151–176 (2003)

    MathSciNet  MATH  Google Scholar 

  • Choksi, R., Ren, X.: Diblock copolymer/homopolymer blends: derivation of a density functional theory. Phys. D 203(1–2), 100–119 (2005)

    MathSciNet  MATH  Google Scholar 

  • Choksi, R., Muratov, C.B., Topaloglu, I.: An old problem resurfaces nonlocally: Gamow’s liquid drops inspire today’s research and applications. Notices of the AMS 64(11), 1275–1283 (2017)

    MathSciNet  MATH  Google Scholar 

  • Cicalese, M., et al.: Ground states of a two phase model with cross and self attractive interactions. SIAM J. Math. Anal. 48(5), 3412–3443 (2016)

    MathSciNet  MATH  Google Scholar 

  • Daneri, S., Kerschbaum, A., Runa, E.: One-dimensionality of the minimizers for a diffuse interface generalized antiferromagnetic model in general dimension. arXiv preprint arXiv:1907.06419 (2019)

  • Daneri, S., Runa, E.: Exact periodic stripes for a local/nonlocal minimization problem with volume constraint. arXiv preprint arXiv:2106.08135 (2021)

  • Daneri, S., Runa, E.: Exact periodic stripes for minimizers of a local/nonlocal interaction functional in general dimension. Arch. Ration. Mech. Anal. 231(1), 519–589 (2019)

    MathSciNet  MATH  Google Scholar 

  • Daneri, S., Runa, E.: Pattern formation for a local/nonlocal interaction functional arising in colloidal systems. SIAM J. Math. Anal. 52(3), 2531–2560 (2020)

    MathSciNet  MATH  Google Scholar 

  • Daneri, S., Runa, E.: One-dimensionality of the minimizers in the large volume limit for a diffuse interface attractive/repulsive model in general dimension. Calc. Var. Partial. Differ. Equ. 61(1), 1–31 (2022)

    MathSciNet  MATH  Google Scholar 

  • Du, Q.: Nonlocal Modeling, Analysis, and Computation. Society for Industrial and Applied Mathematics, Philadelphia (2019)

    MATH  Google Scholar 

  • Du, Q., Gunzburger, M., Lehoucq, R., Zhou, K.: Analysis and approximation of nonlocal diffusion problems with volume constraints. SIAM Rev. 54, 667–696 (2012)

    MathSciNet  MATH  Google Scholar 

  • Feng, H., et al.: Block copolymers: synthesis, self-assembly, and applications. Polymers 9(10), 494 (2017)

    Google Scholar 

  • Frank, R.L.: Non-spherical equilibrium shapes in the liquid drop model. J. Math. Phys. 60(7), 071506 (2019)

    MathSciNet  MATH  Google Scholar 

  • Gamow, G.: Mass defect curve and nuclear constitution. Proc. R. Soc. Lond. Ser. A 126(803), 632–644 (1930)

    MATH  Google Scholar 

  • Gennip, V., Yves, P., Mark, A.: Copolymer- homopolymer blends: global energy minimisation and global energy bounds. Calc. Var. Partial. Differ. Equ. 33(1), 75–111 (2008)

    MathSciNet  MATH  Google Scholar 

  • Giuliani, Alessandro., Lebowitz, Joel L., Lieb, Elliott H..: “Pattern formation in systems with competing interactions.” AIP Conference Proceedings. Vol. 1091. No. 1. American Institute of Physics, 2009

  • Giuliani, A., Seiringer, R.: Periodic striped ground states in Ising models with competing interactions. Commun. Math. Phys. 347(3), 983–1007 (2016)

    MathSciNet  MATH  Google Scholar 

  • Goldman, M., Runa, E.: On the optimality of stripes in a variational model with non-local interactions. Calc. Var. Partial. Differ. Equ. 58(3), 1–26 (2019)

    MathSciNet  MATH  Google Scholar 

  • Goldman, D., Muratov, C.B., Serfaty, S.: The \(\Gamma \)-limit of the two-dimensional Ohta–Kawasaki energy. I. Droplet density. Arch. Ration. Mech. Anal. 210(2), 581–613 (2013)

    MathSciNet  MATH  Google Scholar 

  • Goldman, D., Muratov, C.B., Serfaty, S.: The \(\Gamma \)-limit of the two-dimensional Ohta–Kawasaki energy. Droplet arrangement via the renormalized energy. Arch. Ration. Mech. Anal. 212(2), 445–501 (2014)

    MathSciNet  MATH  Google Scholar 

  • Hardy, C.M., et al.: Model ABC triblock copolymers and blends near the order-disorder transition. Macromolecules 35(8), 3189–3197 (2002)

    Google Scholar 

  • Huang, H., Alexander-Katz, A.: Dissipative particle dynamics for directed self-assembly of block copolymers. J. Chem. Phys. 151(15), 154905 (2019)

    Google Scholar 

  • Ito, A.: Domain patterns in copolymer-homopolymer mixtures. Phys. Rev. E 58(5), 6158 (1998)

    Google Scholar 

  • Jiang, Ying, et al.: Effect of polydispersity on the phase diagrams of linear ABC triblock copolymers in two dimensions. J. Phys. Chem. B 109(44), 21047–21055 (2005)

    Google Scholar 

  • Kang, X., Ren, X.: Ring pattern solutions of a free boundary problem in diblock copolymer morphology. Phys. D 238(6), 645–665 (2009)

    MathSciNet  MATH  Google Scholar 

  • Kang, X., Ren, X.: The pattern of multiple rings from morphogenesis in development. J. Nonlinear Sci. 20(6), 747–779 (2010)

    MathSciNet  MATH  Google Scholar 

  • Kerschbaum, A.: Striped patterns for generalized antiferromagnetic functionals with power law kernels of exponent smaller than \( d+ 2\). arXiv preprint arXiv:2101.02992 (2021)

  • Knüpfer, H., Muratov, C.B., Novaga, M.: Low density phases in a uniformly charged liquid. Commun. Math. Phys. 345(1), 141–183 (2016)

    MathSciNet  MATH  Google Scholar 

  • Lawlor, G.R.: Double bubbles for immiscible fluids in \(\mathbb{R}^n\). J. Geom. Anal. 24(1), 190–204 (2014)

    MathSciNet  MATH  Google Scholar 

  • Liu, M., et al.: Theoretical study of phase behavior of frustrated ABC linear triblock copolymers. Macromolecules 45(23), 9522–9530 (2012)

    Google Scholar 

  • Liu, Y.H., Chew, L.Y., Yu, M.Y.: Self-assembly of complex structures in a two-dimensional system with competing interaction forces. Phys. Rev. E 78(6), 066405 (2008)

    Google Scholar 

  • Luo, W., Zhao, Y.: Nonlocal effect on a generalized Ohta-Kawasaki model. arXiv preprint arXiv:2204.05394 (2022)

  • Lyubimov, I., Wessels, M.G., Jayaraman, A.: Molecular dynamics simulation and PRISM theory study of assembly in solutions of amphiphilic bottlebrush block copolymers. Macromolecules 51(19), 7586–7599 (2018)

    Google Scholar 

  • Mai, Y., Eisenberg, A.: Self-assembly of block copolymers. Chem. Soc. Rev. 41(18), 5969–5985 (2012)

    Google Scholar 

  • Matsen, M.W.: Gyroid versus double-diamond in ABC triblock copolymer melts. J. Chem. Phys. 108(2), 785–796 (1998)

    Google Scholar 

  • Mogi, Y., et al.: Preparation and morphology of triblock copolymers of the ABC type. Macromolecules 25(20), 5408–5411 (1992)

    Google Scholar 

  • Mogi, Y., et al.: Superlattice structures in morphologies of the ABC triblock copolymers. Macromolecules 27(23), 6755–6760 (1994)

    Google Scholar 

  • Morini, M., Sternberg, P.: Cascade of minimizers for a nonlocal isoperimetric problem in thin domains. SIAM J. Math. Anal. 46(3), 2033–2051 (2014)

    MathSciNet  MATH  Google Scholar 

  • Mossa, S., et al.: Ground-state clusters for short-range attractive and long-range repulsive potentials. Langmuir 20(24), 10756–10763 (2004)

    Google Scholar 

  • Muratov, CB.: Theory of domain patterns in systems with long-range interactions of Coulombic type. Ph.D thesis, Boston University, (1998)

  • Muratov, C.B.: Theory of domain patterns in systems with long-range interactions of Coulomb type. Phys. Rev. E 66(6), 066108 (2002)

    MathSciNet  Google Scholar 

  • Muratov, C.B.: Droplet phases in non-local Ginzburg-Landau models with Coulomb repulsion in two dimensions. Commun. Math. Phys. 299(1), 45–87 (2010)

    MathSciNet  MATH  Google Scholar 

  • Muratov, C.B., Simon, T.M.: A nonlocal isoperimetric problem with dipolar repulsion. Commun. Math. Phys. 372(3), 1059–1115 (2019)

    MathSciNet  MATH  Google Scholar 

  • Nakazawa, H., Ohta, T.: Microphase separation of ABC-type triblock copolymers. Macromolecules 26(20), 5503–5511 (1993)

    Google Scholar 

  • Ohta, T., Ito, A.: Dynamics of phase separation in copolymer-homopolymer mixtures. Phys. Rev. E 52(5), 5250 (1995)

    Google Scholar 

  • Ohta, T., Kawasaki, K.: Equilibrium morphology of block copolymer melts. Macromolecules 19(10), 2621–2632 (1986)

    Google Scholar 

  • Reddy, A., et al.: Block Copolymers beneath the Surface: measuring and modeling complex morphology at the subdomain scale. Macromolecules (2021)

  • Ren, X., Wang, C.: A stationary core-shell assembly in a ternary inhibitory system. Discret. Contin. Dyn. Syst. A 37(2), 983 (2017)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wang, C.: Stationary disk assemblies in a ternary system with long range interaction. Commun. Contemp. Math. 21(06), 1850046 (2019)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: On the multiplicity of solutions of two nonlocal variational problems. SIAM J. Math. Anal. 31(4), 909–924 (2000)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Triblock copolymer theory: ordered ABC lamellar phase. J. Nonlinear Sci. 13(2) (2003)

  • Ren, X., Wei, J.: Triblock copolymer theory: free energy, disordered phase and weak segregation. Phys. D 178(1–2), 103–117 (2003)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Many droplet pattern in the cylindrical phase of diblock copolymer morphology. Rev. Math. Phys. 19(08), 879–921 (2007)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Single droplet pattern in the cylindrical phase of diblock copolymer morphology. J. Nonlinear Sci. 17(5), 471–503 (2007)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Spherical solutions to a nonlocal free boundary problem from diblock copolymer morphology. SIAM J. Math. Anal. 39(5), 1497–1535 (2008)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Oval shaped droplet solutions in the saturation process of some pattern formation problems. SIAM J. Appl. Math. 70(4), 1120–1138 (2009)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: A toroidal tube solution to a problem involving mean curvature and Newtonian potential. Interfaces Free Bound. 13(1), 127–154 (2011)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: A double bubble in a ternary system with inhibitory long range interaction. Arch. Ration. Mech. Anal. 208(1), 201–253 (2013)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Asymmetric and symmetric double bubbles in a ternary inhibitory system. SIAM J. Math. Anal. 46(4), 2798–2852 (2014)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: Double tori solution to an equation of mean curvature and Newtonian potential. Calc. Var. Partial. Differ. Equ. 49(3), 987–1018 (2014)

    MathSciNet  MATH  Google Scholar 

  • Ren, X., Wei, J.: A double bubble assembly as a new phase of a ternary inhibitory system. Arch. Ration. Mech. Anal. 215(3), 967–1034 (2015)

    MathSciNet  MATH  Google Scholar 

  • Sides, S.W., Fredrickson, G.H.: Parallel algorithm for numerical self-consistent field theory simulations of block copolymer structure. Polymer 44(19), 5859–5866 (2003)

    Google Scholar 

  • Spadaro, N.: Uniform energy and density distribution: diblock copolymers’ functional. Interfaces Free Bound. 11(3), 447–474 (2009)

    MathSciNet  MATH  Google Scholar 

  • Sternberg, P., Topaloglu, I.: On the global minimizers of a nonlocal isoperimetric problem in two dimensions. Interfaces Free Bound. 13(1), 155–169 (2011)

    MathSciNet  MATH  Google Scholar 

  • Sun, M., et al.: Morphology and phase diagram of A B C linear triblock copolymers: parallel real-space self-consistent-field-theory simulation. Phys. Rev. E 77(1), 016701 (2008)

    Google Scholar 

  • Tang, P., et al.: Morphology and phase diagram of complex block copolymers: ABC linear triblock copolymers. Phys. Rev. E 69(3), 031803 (2004)

    Google Scholar 

  • Topaloglu, I.: On a nonlocal isoperimetric problem on the two-sphere. Commun. Pure Appl. Anal. 12(1), 597 (2013)

    MathSciNet  MATH  Google Scholar 

  • Van Gennip, Y., Peletier, M.A.: Stability of monolayers and bilayers in a copolymer-homopolymer blend model. Interfaces Free Bound. 11(3), 331–373 (2009)

    MathSciNet  MATH  Google Scholar 

  • Wang, C.: Analysis and modeling of self-organized systems with long range interaction. Ph.D thesis, The George Washington University, (2018)

  • Wang, C., Ren, X., Zhao, Y.: Bubble assemblies in ternary systems with long range interaction. Commun. Math. Sci. 17(8), 2309–2324 (2019)

    MathSciNet  MATH  Google Scholar 

  • Wickham, R.A., Shi, A.-C.: Noncentrosymmetric lamellar phase in blends of ABC triblock and ac diblock copolymers. Macromolecules 34(18), 6487–6494 (2001)

    Google Scholar 

  • Wong, B.: Points of view: color blindness. Nat. Methods 8, 441 (2011). https://doi.org/10.1038/nmeth.1618

    Article  Google Scholar 

  • Xia, J., et al.: Microphase ordering mechanisms in linear ABC triblock copolymers. A dynamic density functional study. Macromolecules 38(22), 9324–9332 (2005)

    Google Scholar 

  • Zheng, W., Wang, Z.-G.: Morphology of ABC triblock copolymers. Macromolecules 28(21), 7215–7223 (1995)

    Google Scholar 

Download references

Acknowledgements

The authors would like to thank Professors Chong Wang, Xiaofeng Ren, An-Chang Shi, Juncheng Wei and Yanxiang Zhao for helpful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zirui Xu.

Additional information

Communicated by Rustum Choksi.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This work is supported in part by the National Science Foundation DMS-2012562 and DMS-1937254.

Appendices

A Underlying Mechanism of Interactions Between Generalized Charges

The decomposition (5.10) of \([f_{ij}]\) suggests a possible way to interpret the interactions between the generalized charges:

  • Each charge consists of two sub-charges, as shown in Fig. 9. For example, a charge of type 1 consists of a sub-charge of type and a sub-charge of type .

  • For \(i\ne j\), the potential energy due to the interaction between the pair of sub-charges at \(\vec x\) and at \(\vec y\) is \(f_{ij}\,G(\vec x,\vec y)\), and that between at \(\vec x\) and at \(\vec y\) is \(-f_{ij}\,G(\vec x,\vec y)\). There is no interaction within other pairs.

Fig. 9
figure 9

Decomposition of charges into sub-charges. , and balls represent charges of types 1, 2 and 3, respectively. Each charge consists of two (out of six) types of sub-charges (Color figure online)

We further assume that there are van der Waals forces between charges, that the cohesive forces between charges of the same type are stronger than the adhesive forces between charges of different types, so that the charges behave like immiscible fluids in the thermodynamic limit at certain temperature, with the interfacial tensions being \(\{c_{ij}\}\). (For a general account of the interfacial tension, see, e.g., Birdi (2015).) To ensure the overall charge neutrality, there must be the same number of charges of each type (Fig. 9 shows four charges of each type). In accordance with the volume constraints, the volume ratio of charges of types 1, 2 and 3 is \(\omega _1:\omega _2:\omega _3\). On the continuum level, such a discrete particle system can be described by the ternary O–K free energy and thus serves as an intuitive analogy. This analogy naturally raises a question on how to arrange balls in 1-D in order to minimize the potential energy between charges, a question further discussed in Sect. 6.2.1 and Appendix B. The decomposition into simple interactions between sub-charges helps us answer this question for \(f_{12},f_{13},f_{23}\leqslant 0\) (see Proposition 7.6).

B Optimal Arrangement of Charged Balls in 1-D

1.1 B.1 Binary case

Given a positive integer n, consider a 1-D periodic cell [0, 1] packed with n balls of type \(A\) and n balls of type \(B\) , with unit amounts of positive and negative point charges at their centers, respectively. We assume that all the balls have the same radius \(\frac{1}{4n}\), and that their centers are located at \(\frac{k}{2n}\) for \(k = 1,2,\cdots ,2n\). Let \(u:\big \{\frac{k}{2n}\big \}_{k=1}^{2n}\rightarrow \pm 1\) represent the arrangement of the balls, with 1 and \(-1\) denoting \(A\) and \(B\) , respectively, then the total potential energy between charges can be written as

$$\begin{aligned} U(u)=\frac{1}{2}\sum _{k=1}^{2n}\sum _{m=1}^{2n}u\Big (\frac{k}{2n}\Big )u\Big (\frac{m}{2n}\Big )G\Big (\frac{k}{2n},\frac{m}{2n}\Big ), \end{aligned}$$
(7.1)

where G is given by (7.4). By Proposition 7.1, the alternating arrangement \(A\)\(B\)\(\cdots \;\) \(A\)\(B\) minimizes U, as shown in Fig. 10 (Throughout Appendix B,“minimize” refers to “globally minimize”.). This is a long-range variant of the 1-D antiferromagnetic Ising model without external fields, subject to the zero overall spin constraint.

Fig. 10
figure 10

An optimal arrangement for \(n=3\). and represent \(A\) and \(B\) balls with positive and negative point charges at their centers, respectively (Color figure online)

Proposition 7.1

The minimizer of (7.1) is (up to translation) \(u\big (\frac{k}{2n}\big )=(-1)^{k-1}\) for \(k = 1,2,\cdots ,2n\).

Proof

(Inspired by Ren and Wei (2000, Proof of Proposition 3.1) Within an optimal arrangement, let us prove that l must be 1 for any segment like the following

By translational invariance, we can assume that the above segment occupies the first \(l\!+\!2\) sites \(\big \{\frac{k}{2n}\big \}_{k=1}^{l+2}\). By assumption, U should not decrease if we swap the \(A\) and \(B\) at the first two sites. Let \(u^*\) represent this optimal arrangement, then we have

$$\begin{aligned} \sum _{m=3}^{2n}u^*\Big (\frac{m}{2n}\Big )G\Big (\frac{1}{2n},\frac{m}{2n}\Big )\leqslant \sum _{m=3}^{2n}u^*\Big (\frac{m}{2n}\Big )G\Big (\frac{2}{2n},\frac{m}{2n}\Big ). \end{aligned}$$
(7.2)

Define the electrostatic potential

$$\begin{aligned} V(x;u)=\sum _{m=1}^{2n}u\Big (\frac{m}{2n}\Big )G\Big (x,\frac{m}{2n}\Big ),\quad x\in [0,1]. \end{aligned}$$

Then from (7.2), we know

$$\begin{aligned} \begin{aligned} V\Big (\frac{1}{2n};u^*\Big )\!-\!V\Big (\frac{2}{2n};u^*\Big )&\leqslant G\Big (\frac{1}{2n},\frac{1}{2n}\Big )\!-\!2G\Big (\frac{1}{2n},\frac{2}{2n}\Big )\!+\!G\Big (\frac{2}{2n},\frac{2}{2n}\Big )\\&=\frac{1}{2n}-\frac{1}{4n^2}. \end{aligned} \end{aligned}$$

From (7.4) and \(\sum _{m=1}^{2n}u\big (\frac{m}{2n}\big )=0\), we can see that the potential V is piecewise quadratic in x with the coefficient of \(x^2\) being 0 and thus is linear on every subinterval \([\frac{k-1}{2n},\frac{k}{2n}]\) for \(k=1,\cdots ,2n\). Define the electrostatic field \(E(x;u)=\mathrm {d}V(x;u)/\mathrm {d}x\), then E is piecewise constant in x and

$$\begin{aligned} E(x;u^*)\big |_{x\in \big (\frac{1}{2n},\frac{2}{2n}\big )}\geqslant \frac{1}{2n}\!-\!1>-1. \end{aligned}$$

Analogously we have

$$\begin{aligned} E(x;u^*)\big |_{x\in \big (\frac{l+1}{2n},\frac{l+2}{2n}\big )}<1. \end{aligned}$$

We also know for \(k=1,\cdots ,2n\!-\!1\),

$$\begin{aligned} E(x;u)\big |_{x\in \big (\frac{k-1}{2n},\frac{k}{2n}\big )}-E(x;u)\big |_{x\in \big (\frac{k}{2n},\frac{k+1}{2n}\big )}=u\Big (\frac{k}{2n}\Big ). \end{aligned}$$

Consequently, we have \(-\,l>(-1)-1\), that is, \(l=1\). \(\square \)

Remark 7.2

The above proof can be generalized to the case where different types of balls have different sizes, that is, the positions of the point charges are no longer uniform (i.e., \(x_k=\frac{k}{2n}\) for \(k=1,2,\cdots ,2n\)). Instead, we have

$$\begin{aligned} x_k-x_{k-1}=\frac{1\!+\!u(x_k)\omega }{4n}+\frac{1\!+\!u(x_{k-1})\omega }{4n},\quad \text {for }k=1,2,\cdots ,2n, \end{aligned}$$

where \(x_0=0\) is identified with \(x_{2n}=1\) so that \(u(x_0)=u(x_{2n})\), and the radii of \(A\) and \(B\) balls are \(\frac{1+\omega }{4n}\) and \(\frac{1-\omega }{4n}\), respectively, for some \(\omega \in (-1,1)\backslash \{0\}\).

1.2 B.2 Ternary Case

Given a positive integer n, consider a 1-D periodic cell [0, 1] packed with n balls of type \(A\) , n balls of type \(B\) , and n balls of type \(C\) , labeled by 1, 2 and 3, respectively. Balls of type i are assumed to have the radius \(\omega _i/(2n)\). We also assume \(f_{ij}\,G(x,y)\) to be the potential energy between a ball of type i centered at x and a ball of type j centered at y. The total potential energy U is the sum of all the pairwise interactions:

$$\begin{aligned} U=\frac{1}{2}\sum _{k=1}^{3n}\sum _{m=1}^{3n}f_{\,i_k\,j_m}G(x_k,y_m), \end{aligned}$$
(7.3)

where G is given by (7.4). The k-th ball, which is of type \(i_k\), is centered at \(x_k\), and the m-th ball, which is of type \(j_m\), is centered at \(y_m\). Therefore, we have the following relation

$$\begin{aligned} x_k\!-\!x_{k-1}=\frac{\omega _{i_k}}{2n}\!+\!\frac{\omega _{i_{k-1}}}{2n},\;\;x_k=y_k,\;\;\text {and}\;\;i_k=j_k, \end{aligned}$$

for \(k=1,2,\cdots ,3n\), with \(x_0=y_0=0\) and \(i_0=j_0=i_{3n}=j_{3n}\).

Conjecture 7.3

For any admissible interaction strength matrix \([f_{ij}]\) and any positive \(\{\omega _i\}\) satisfying \(\sum _i\omega _i=1\), the arrangement \(A\)\(B\)\(C\)\(\cdots \)\(A\)\(B\)\(C\) (i.e., \(i_k\equiv k\mod 3\)) minimizes (7.3), as illustrated on the right side of Fig. 5.

We have numerically verified Conjecture 7.3 from \(n=2\) to 8 for the Cartesian product of 100 choices of \([f_{ij}]\) and 72 choices of \(\{\omega _i\}\) (up to permutations we can assume \(\omega _1\leqslant \omega _2\leqslant \omega _3\)), as shown in Figs.11 and 12. We also prove some special cases of Conjecture 7.3 in Propositions 7.4 and 7.6.

Fig. 11
figure 11

Choices of \([f_{ij}]\) in the numerical verification. Colorful cap: the entire range of admissible \([f_{ij}]\) subject to \(f_{12}^2+f_{13}^2+f_{23}^2=1\). Black and white dots: samples of \([f_{ij}]\) used in the numerical computation. Colors are only for visualization

Fig. 12
figure 12

Choices of \(\{\omega _i\}\) in the numerical verification. triangular plate: the entire range of \(\{\omega _i\}\). dots: samples of \(\{\omega _i\}\). Portion enclosed by dashed line segments: \(\omega _1\leqslant \omega _2\leqslant \omega _3\) (Color figure online)

Proposition 7.4

For \([f_{ij}]\) given by (5.6), the minimizers of (7.3) are (up to translation and reflection) the following with any \(l_k\geqslant 0\) satisfying \(\sum _{k=1}^nl_k=n\),

Proof

For (5.6), balls of type \(C\) do not engage in the interaction between charges. Similar to Proof of Proposition 7.1, we can prove that within any optimal arrangement, \(A\) and \(B\) must be alternating if \(C\) is ignored. Now let us exclude the following segment from optimal arrangements:

By translational invariance, we can assume that the above segment occupies the first 3 sites \(\{x_k\}_{k=1}^3\). Define the potential created by charges at all other sites as

$$\begin{aligned} V_2(x)=\sum _{m=4}^{3n}f_{\,i_2\,j_m}G(x,y_m),\quad x\in [0,1],\quad \text {where}\;i_2=1. \end{aligned}$$

Among the remaining balls centered at \(\{y_m\}_{m=4}^{3n}\), there are one more \(B\) than \(A\) , so \(V_2\) is piecewise quadratic in x with the coefficient of \(x^2\) being \(-\frac{1}{2}\), and thus is strictly concave on \([x_0,x_4]\). Therefore by swapping the \(A\) at \(x_2\) and the \(C\) at either \(x_1\) or \(x_3\), we can decrease U by the amount

$$\begin{aligned} V_2(x_2)-\min \Big \{V_2\Big (x_2\!-\!\frac{\omega _3}{n}\Big ),\;V_2\Big (x_2\!+\!\frac{\omega _3}{n}\Big )\Big \}. \end{aligned}$$

Analogously, we can rule out other segments like \(C\)\(A\)\(B\)\(A\)\(C\) , \(C\)\(A\)\(B\)\(A\)\(B\)\(A\)\(C\) , etc. Lastly, let us verify that different choices of \(\{l_k\}\) yield the same U. Consider two \(A\)\(B\) dipoles separated by \(C\)

between which the potential energy is

$$\begin{aligned} \begin{aligned}&G(x_1,x_{l+3})-G(x_1,x_{l+4})-G(x_2,x_{l+3})+G(x_2,x_{l+4})\\&\quad =-(x_2-x_1)\,(x_{l+4}-x_{l+3})\\&\quad =-\Big (\frac{\omega _1}{2n}+\frac{\omega _2}{2n}\Big )^2, \end{aligned} \end{aligned}$$

and is independent of l. \(\square \)

Remark 7.5

In Proposition 7.4, if we penalize \(A\)\(C\) and \(B\)\(C\) interfaces with equal weights, then the minimizer is unique, i.e., \(A\)\(B\)\(\cdots \;\) \(A\)\(B\)\(C\)\(\cdots \;\) \(C\) . This is reminiscent of the results in Van Gennip and Peletier (2008, Theorems 4 and 5) where \(C\) forms only one macrodomain.

Proposition 7.6

For \([f_{ij}]\) given by (5.10) with nonpositive \(f_{12}\), \(f_{13}\) and \(f_{23}\), the arrangement \(A\)\(B\)\(C\)\(\cdots \)\(A\)\(B\)\(C\) minimizes (7.3).

Proof

By (5.10), \([f_{ij}]\) can be decomposed into three components, which are permutations of (5.6) and represent the interactions between sub-charges shown in Fig. 9. With the assumption \(f_{12},f_{13},f_{23}\leqslant 0\), all three components are simultaneously minimized by the cyclic arrangement \(A\)\(B\)\(C\)\(\cdots \)\(A\)\(B\)\(C\) , according to Proposition 7.4.   \(\square \)

C Numerical Computation of Free Energy in a 1-D Periodic Cell

We now offer some details of the numerical computation to seek for the 1-D global minimizers of the free energy (1.2). For each pattern, we obtain the optimal layer widths numerically, with the initial guess (for optimization) having uniform layer widths. We use , a constrained local minimization function in MATLAB, with the constraints being the volume constraints and nonnegativity of layer widths. For the input argument , we set , and to be \(10^{-6}\), which should be sufficient for our purposes. Since patterns are defined modulo translation (Ren and Wei 2003b, Definition 4.1), we can avoid some redundant computation. For further acceleration, we use MATLAB’s parallel tool to work on multiple (e.g., 24) patterns simultaneously.

We adopt a simple algorithm based on (4.3) to compute the long-range term of (1.2). Noticing that the Green’s function on [0, 1] with periodic boundary conditions is given by Ren and Wei (2003b, Equation (4.19))

$$\begin{aligned} G(x,y)=\frac{|x\!-\!y|^2}{2}-\frac{|x\!-\!y|}{2}+\frac{1}{12},\quad \text {for}\;x,y\in [0,1], \end{aligned}$$
(7.4)

we have \(\int _{y_1}^{y_2}\int _{x_1}^{x_2}G(x,y)\mathrm {d}{x}\mathrm {d}{y}=F(x_2\!-\!y_1)-F(x_1\!-\!y_1)-F(x_2\!-\!y_2)+F(x_1\!-\!y_2)\), where

$$\begin{aligned} F(x)=\frac{\big (1\!-\!|x|\big )^2x^2}{24}. \end{aligned}$$

Now, given a pattern and the positions of interfaces, the following algorithm returns the free energy J.

figure aq

Although this algorithm is of complexity \(O\big (\) \(2\big )\), it is not a bottleneck compared to the exhaustive search (among all the patterns) of complexity \(O\big (2\) \(\big )\), since the total complexity is the product of the above two and that of (for constrained optimization in ).

D Analytic calculation of free energy of 1-D periodic patterns

For \(\Omega =[0,1]\) with periodic boundary conditions, the long-range term of (1.1) can be rewritten as

$$\begin{aligned} \sum _{i=1}^3\sum _{j=1}^3\gamma _{ij}\int _0^1\big (u_i(x)\!-\!\omega _i\big )v_j(x)\mathrm {d}{x}= \sum _{i=1}^3\sum _{j=1}^3\gamma _{ij}\int _0^1w_i(x)\,w_j(x)\mathrm {d}{x}, \end{aligned}$$

where \(v_j(x)=\int _0^1G(x,y)\,\big (u_j(y)-\omega _j\big )\mathrm {d}{y}\) (so that \(-v_j''=u_j-\omega _j\)) and \(w_j=-v_j'\,\). Denoting \({\vec {w}}=[w_1,w_2,w_3]^\mathrm{T}\), we can rewrite the above right-hand side as

$$\begin{aligned} \int _0^1{\vec {w}}(x)^\mathrm{T}\,[\gamma _{ij}]\,{\vec {w}}(x)\mathrm {d}{x}. \end{aligned}$$

Ren and Wei (2003b, Section 4), studied a local minimizer which is \(A\)\(B\)\(C\) identically repeating for n times. In that case, \({\vec {w}}\) is periodic with period 1/n, so one only needs to solve the following equation in one period in order to obtain the free energy

$$\begin{aligned} \frac{\mathrm {d}{\vec {w}}}{\mathrm {d}x}= \left\{ \begin{aligned}&\vec e_1-{\vec {\omega }},&\;0<x<\frac{\omega _1}{n},\\&\vec e_2-{\vec {\omega }},&\;\frac{\omega _1}{n}<x<\frac{\omega _1\!+\!\omega _2}{n},\\&\vec e_3-{\vec {\omega }},&\;\frac{\omega _1\!+\!\omega _2}{n}<x<\frac{1}{n}, \end{aligned} \right. \quad \text {with}\;\int _0^1{\vec {w}}(x)\mathrm {d}{x}={\vec {0}}, \end{aligned}$$

where \({\vec {\omega }}\) denotes \([\omega _1,\omega _2,\omega _3]^\mathrm{T}\), and \(\{\vec e_1,\vec e_2,\vec e_3\}\) forms the standard basis. In this way, we can obtain (3.1).

Analogously, for \(A\)\(B\)\(A\)\(C\) identically repeating for n times (with all the \(A\) layers having the same width), we can obtain (3.2) by solving the following equation

$$\begin{aligned} \frac{\mathrm {d}{\vec {w}}}{\mathrm {d}x}= \left\{ \begin{aligned}&\vec e_1-{\vec {\omega }},&\;0<x<\frac{\omega _1}{2n},\\&\vec e_2-{\vec {\omega }},&\;\frac{\omega _1}{2n}<x<\frac{\omega _1}{2n}\!+\!\frac{\omega _2}{n},\\&\vec e_1-{\vec {\omega }},&\;\frac{\omega _1}{2n}\!+\!\frac{\omega _2}{n}<x<\frac{\omega _1\!+\!\omega _2}{n},\\&\vec e_3-{\vec {\omega }},&\;\frac{\omega _1\!+\!\omega _2}{n}<x<\frac{1}{n}, \end{aligned} \right. \quad \text {with}\;\int _0^1{\vec {w}}(x)\mathrm {d}{x}={\vec {0}}. \end{aligned}$$

E Alternative Derivation of the Admissibility Conditions

As mentioned in Remark -(i), there is an alternative derivation of the conditions in Theorem 5.2 from the following three requirements:

  • For \(\vec u={\vec {\omega }}\), the long-range term (4.2) attains zero;

  • For any \(\vec u\) satisfying the incompressibility condition \(\vec u^\mathrm{T}{\vec {1}}=1\), the long-range term (4.2) is nonnegative;

  • \([\gamma _{ij}]\) is symmetric.

Under Neumann or periodic boundary conditions, we have \(\int _{\Omega }G(\vec x,\vec y)\mathrm {d}{\vec x}=0\) for any \(\vec y\in \Omega \), so the first requirement is automatically satisfied and therefore does not lead to the condition \({\vec {\omega }}^\mathrm{T}[\gamma _{ij}]{\vec {\omega }}=0\). Moreover, the second requirement does not lead to the condition \([\gamma _{ij}]\succcurlyeq 0\) because of the incompressibility condition. However, as explained in Proposition 7.7, there are many equivalent choices of \([\gamma _{ij}]\), and one of them satisfies \([\gamma _{ij}]{\vec {\omega }}={\vec {0}}\) and \([\gamma _{ij}]\succcurlyeq 0\) as desired.

Proposition 7.7

Under the incompressibility condition, among all the \([\gamma _{ij}]\) fulfilling the above three requirements and yielding the same long-range term (4.2), there is a unique one satisfying \([\gamma _{ij}]\,{\vec {\omega }}={\vec {0}}\). Meanwhile, it also satisfies \([\gamma _{ij}]\succcurlyeq 0\).

Proof

Under the asssumption \(u_1+u_2+u_3=1\), we have \(\vec u=[u_1,u_2,u_3]^\mathrm{T}={\mathcal {A}}[u_1,u_2]^\mathrm{T}+[0,0,1]^\mathrm{T}\), where \({\mathcal {A}}\) is given by (7.5), and the second summand is constant. Since \(\int _{\Omega }G(\vec x,\vec y)\mathrm {d}{\vec x}=0\) for any \(\vec y\in \Omega \), we can rewrite (4.2) as

$$\begin{aligned} \int _{\Omega }\int _{\Omega } \begin{bmatrix} u_1(\vec x)&u_2(\vec x) \end{bmatrix} [{\tilde{\gamma }}_{ij}] \begin{bmatrix} u_1(\vec y)\\ u_2(\vec y) \end{bmatrix} G(\vec x,\vec y)\mathrm {d}{\vec x}\mathrm {d}{\vec y}, \end{aligned}$$

where \([{\tilde{\gamma }}_{ij}]={\mathcal {A}}^\mathrm{T}[\gamma _{ij}]{\mathcal {A}}\). To ensure that the above integral is nonnegative, we need to impose the condition \([{\tilde{\gamma }}_{ij}]\succcurlyeq 0\). (In fact, we can diagonalize \([{\tilde{\gamma }}_{ij}]\) into \(Q^\mathrm{T}\mathrm{diag}(\lambda _1,\lambda _2)\,Q\), and rewrite the above integral as a quadratic form like (4.4), from which it would be clear that \(\lambda _1\) and \(\lambda _2\) should be both nonnegative.)

By Lemma 7.8, there is a class of equivalent choices of \([\gamma _{ij}]\), but only one of them satisfies \([\gamma _{ij}]{\vec {\omega }}={\vec {0}}\). Such \([\gamma _{ij}]\) is given by (5.7) and is positive semi-definite since we have \([{\tilde{\gamma }}_{ij}]\succcurlyeq 0\). \(\square \)

Lemma 7.8

Define \(T:S_3\rightarrow S_2\) to be \(T(H)={\mathcal {A}}^\mathrm{T}H{\mathcal {A}}\), where \(S_m\) is the set of \(m\times m\) symmetric real matrices, and

$$\begin{aligned} {\mathcal {A}}= \begin{bmatrix} 1 &{}0\\ 0 &{}1\\ -1 &{}-1 \end{bmatrix}, \end{aligned}$$
(7.5)

then T is surjective. The kernel of T is \(\big \{{\vec {1}}\vec p^\mathrm{T}+\vec p{\vec {1}}^\mathrm{T}\;\big |\;\vec p\in {\mathbb {R}}^3\big \}\). Given any \({\vec {w}}=[w_1,w_2,w_3]^\mathrm{T}\in {\mathbb {R}}^3\) with \({\vec {w}}^\mathrm{T}{\vec {1}}=1\), in the quotient space \(S_3/\mathrm{ker}(T)\), the equivalence class of any \(H\in S_3\) has a unique representative \({\tilde{H}}\) satisfying \({\tilde{H}}\,{\vec {w}}={\vec {0}}\). This representative is given by \({\tilde{H}}={\mathcal {B}}^\mathrm{T}T(H){\mathcal {B}}\), where

$$\begin{aligned} {\mathcal {B}}= \begin{bmatrix} 1-w_1 &{} -w_1 &{} -w_1 \\ -w_2 &{} 1-w_2 &{} -w_2 \end{bmatrix}. \end{aligned}$$
(7.6)

Proof

We can take \(H=\begin{bmatrix}H_1 &{} 0\\ 0 &{} 0\end{bmatrix}\), where \(H_1\in S_2\), then \(T(H)=H_1\), so T is surjective. Since \({\vec {1}}^\mathrm{T}{\mathcal {A}}={\vec {0}}^\mathrm{T}\), based on the rank–nullity theorem we know \(\mathrm{ker}(T)=\big \{{\vec {1}}\vec p^\mathrm{T}+\vec p{\vec {1}}^\mathrm{T}\;\big |\;\vec p\in {\mathbb {R}}^3\big \}\). For any \(H\in S_3\), we have

$$\begin{aligned} (H+{\vec {1}}\vec p^\mathrm{T}+\vec p{\vec {1}}^\mathrm{T}){\vec {w}}=H{\vec {w}}+{\vec {1}}{\vec {w}}^\mathrm{T}\vec p+\vec p=H{\vec {w}}+(I+{\vec {1}}{\vec {w}}^\mathrm{T})\vec p, \end{aligned}$$

where I is an identity matrix. According to the Sherman–Morrison formula, there is a unique \(\vec p\) such that the above right-hand side vanishes. We can verify that \({\tilde{H}}={\mathcal {B}}^\mathrm{T}T(H){\mathcal {B}}\) is the corresponding representative within the equivalence class of H, by checking \({\tilde{H}}{\vec {w}}={\vec {0}}\) and \(T({\tilde{H}})=T(H)\), which are clear from \({\mathcal {B}}{\vec {w}}={\vec {0}}\) and \({\mathcal {B}}{\mathcal {A}}=I\), respectively. \(\square \)

Remark 7.9

The results in this section can be generalized from \(\vec u\in {\mathbb {R}}^3\) to any dimension \({\mathbb {R}}^m\), by changing (7.5) into

$$\begin{aligned} {\mathcal {A}}=\begin{bmatrix} I_{m-1}\\ -{{\vec {1}}}^\mathrm{T} \end{bmatrix}, \end{aligned}$$

and changing (7.6) into

$$\begin{aligned} {\mathcal {B}}= \begin{bmatrix} I_{m-1}&{\vec {0}} \end{bmatrix} (I_{m}-{\vec {w}}\,{\vec {1}}^\mathrm{T}), \end{aligned}$$

where \(I_m\) is the \(m\times m\) identity matrix, and \({\vec {1}}\) is a vector whose components are all 1.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, Z., Du, Q. On the Ternary Ohta–Kawasaki Free Energy and Its One-dimensional Global Minimizers. J Nonlinear Sci 32, 61 (2022). https://doi.org/10.1007/s00332-022-09814-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-022-09814-9

Navigation