Next Article in Journal
Thymosin α1 and Its Role in Viral Infectious Diseases: The Mechanism and Clinical Application
Previous Article in Journal
Gadolinium-Cyclic 1,4,7,10-Tetraazacyclododecane-1,4,7,10-Tetraacetic Acid-Click-Sulfonyl Fluoride for Probing Serine Protease Activity in Magnetic Resonance Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Study of Phase Behaviors of Symmetric Linear B1A1B2A2B3 Pentablock Copolymer

1
Department of Physics, Taizhou University, Taizhou 318000, China
2
State Key Laboratory of Molecular Engineering of Polymers, Department of Macromolecular Science, Fudan University, Shanghai 200433, China
3
Faculty of Materials Science and Chemical Engineering, Ningbo University, Ningbo 315211, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(8), 3536; https://doi.org/10.3390/molecules28083536
Submission received: 18 March 2023 / Revised: 8 April 2023 / Accepted: 16 April 2023 / Published: 17 April 2023

Abstract

:
The nanostructures that are self-assembled from block copolymer systems have attracted interest. Generally, it is believed that the dominating stable spherical phase is body-centered cubic (BCC) in linear AB-type block copolymer systems. The question of how to obtain spherical phases with other arrangements, such as the face-centered cubic (FCC) phase, has become a very interesting scientific problem. In this work, the phase behaviors of a symmetric linear B1A1B2A2B3 (fA1 = fA2, fB1 = fB3) pentablock copolymer are studied using the self-consistent field theory (SCFT), from which the influence of the relative length of the bridging B2-block on the formation of ordered nanostructures is revealed. By calculating the free energy of the candidate ordered phases, we determine that the stability regime of the BCC phase can be replaced by the FCC phase completely by tuning the length ratio of the middle bridging B2-block, demonstrating the key role of B2-block in stabilizing the spherical packing phase. More interestingly, the unusual phase transitions between the BCC and FCC spherical phases, i.e., BCC → FCC → BCC → FCC → BCC, are observed as the length of the bridging B2-block increases. Even though the topology of the phase diagrams is less affected, the phase windows of the several ordered nanostructures are dramatically changed. Specifically, the changing of the bridging B2-block can significantly adjust the asymmetrical phase regime of the Fddd network phase.

1. Introduction

The nanoscale nanostructures self-assembled from the block copolymer (BCP) systems have attracted widescale interest due to their potential technological applications in a variety of fields and in daily life [1,2,3,4,5]. For example, the self-assembled structures from BCPs can be used as organic electronics [6], for water purification [7], energy materials [8], photonic materials [9], as well as in drug delivery [10,11]. Generally, BCPs are composed of chemically distinct homopolymer connected by covalent bonds. The repulsive interaction between the different chemical blocks leads to phase separation; however, the connectivity of the blocks drives the phase to separate in the microscopic scale. From a thermodynamic point of view, the formation of different nanoscopic domains is governed by the delicate balance of interfacial energy and stretching energy. Specifically, the stable nanoscale structures that are self-assembled from AB-type BCPs are body-centered cubic (BCC), hexagonal cylinder (C6), bicontinuous phases-Gyroid (G) and Fddd (O70), and lamella (L), determined by a competition between the two opposite trends [12,13,14,15]. Beyond these conventional structures, some vagarious nanostructures have attracted more and more attention in both experiment and theory, including the helical structure [16,17,18,19], quasicrystalline phase [20,21,22], hierarchical structure [23,24,25,26], Frank–Kasper phases [27,28,29,30,31], and so on. For example, many well-known biological systems are presented in the form of helical structures, including DNA and RNA. However, the origin of the homochirality is rarely understood. Recent research has shown that the double helical structure can be formed from a number of typical BCP systems both in theory and experiment, indicating that the BCP system can provide a powerful platform with which to understand and reveal the mysterious origin of the life. It has become a crucial issue to reveal the relationship between the initial chain topologies of multiblock copolymers and the final thermodynamically stable self-assembled nanostructures. In general, the architectures can significantly change the phase behaviors of the AB-type BCPs. However, the theoretical study shows that the phase behaviors of symmetric triblocks (ABA) and alternating (ABAB…) linear BCPs are relatively unchanged [32]. Therefore, it is generally agreed that the AB linear multiblock copolymers have similar phase diagrams.
In 2018, Sakurai and coworkers synthesized a series of S1BS2 asymmetric triblock copolymers to investigate the relationship between chain architecture and ordered structures [33]. Encouragingly, their experimental results showed that, even though the total molecular weight, volume fraction, and thermal annealing temperature were kept unchanged, the thermodynamically stable ordered structure was found to change by tuning the relative length between S1 and S2 blocks. These experimental results are consistent with Matsen’s prediction qualitatively, which is implemented by the self-consistent field theory (SCFT) [34]. Notice that the linear AB-multiblock copolymers have more tunable parameters to adjust the length ratio of each A/B-block; therefore, it provides a broad parameter space to regulate and research the self-assembly of AB-type BCP systems. Furtherly, Kim et al. investigated the phase behaviors of various linear tetrablock copolymers (S1I1S2I2) with fixed symmetric volume fraction (fPSfPI ≈ 0.5 and fPS1f PS2) [35]. Interestingly, they observed the C6 and G phases by tuning the relative length ratio fPI1, which is also in accordance with the prediction of SCFT [36].
As one of the most successful methods, SCFT plays a critically important role to explain and predict the thermodynamically stable state of the BCP systems [37,38,39,40]. Due to its strong ability to calculate the free energy accurately as well as to visualize the distribution of each segment in the ordered phase, SCFT has been regarded as the standard theoretical tool for the study of the self-assembly of the BCPs. In 2014, Xie et al. predicted a series of novel binary mesocrystals from B1AB2CB3 multiblock terpolymers, including NaCl, CsCl, ZnS, α-BN, AlB2, CaF2, TiO2, and so on, in which a minority of A- and C-blocks self-assembled into the spherical domains and a majority of B-blocks formed the matrix [41]. The main idea is that the length of middle B2-block can control the average coordination number (CN) of the binary mesocrystals, while the asymmetry of CN is modulated by the relative length of the terminal B1- and B3-block. This distinguished work opens a window for the inverse design of BCPs to obtain the targeted structure, and thus makes it possible to obtain total binary crystals and even ternary crystals in the BCP systems. The reversed design principle was not only applied to ABC-type BCP systems, but also to the AB multi-BCP systems in recent years.
Recently, several mechanisms of the self-assembly of BCPs has been proposed, including local segregation [30,42], the solubilization effect [34], packing frustration [43], a stretched bridge [41], and combinatorial entropy [44]. For example, our previous work predicted that the Frank–Kasper (FK) σ and A15 phases can be stabilized from a linear A1B1A2B2 tetrablock copolymer, in which the longer A1-block forms the “core” while the short A2-block forms the “shell” to regulate the sizes and shapes of the spherical domains, respectively [42]. Further, a linear B1A1B2A2B3 pentablock copolymer has been proposed to predict the single gyroid (SG) structure, in which a minority of A-blocks self-assemble into the network [45]. By tuning the asymmetry between the B1- and B3-blocks as well as the middle bridging B2-block, the phase transition from DG to SG can be triggered, utilizing the synergistic effect of two different mechanisms, i.e., packing frustration and stretched bridging effects. It is worth noting that the hexagonal cylinders (C6) are completely replaced by the unusual square array of cylinders (C4) as well as the rectangular array of cylinders (Crec). Because the relative lengths of the multiple blocks can be adjusted, this theoretical study demonstrates that the AB linear multi-BCPs can enrich the phase behaviors compared with simple AB and ABA copolymers. Encouragingly, some of the theoretical predictions were verified by the experiments. For example, the complex FK spherical phases have been predicted from conformational asymmetry AB-type BCP melts, in which the FK spherical domains have different sizes and shapes [29]. However, the theoretical prediction was not verified by experiments until 2017. In this experiment, Bates and co-workers tactfully used a series of di-BCPs (i.e., PEP-b-PLA, PI-b-PLA, and PEE-b-PLA) with a different degree of conformational asymmetry, verifying that the FK phases can only be formed under the condition of a highly asymmetric structure in melt systems [28]. With the close cooperation between theory and experiment, the explanatory ability and predictive power of SCFT have been further confirmed.
However, up to now, the single stretched bridging effect on the AB linear multiblock copolymer self-assembly has not been considered. Recently, several typical (B1A1B2)m (m = 3, 5) star copolymers have been well studied systematically, showing that some intriguing phase behaviors have been predicted [46,47]. Particularly, the star architecture (B1A1B2)m will be converted to a symmetric linear B1A1B2A2B3 multi-BCP for m = 2. Due to a rare report about the stretching bridge in the linear AB multi-BCPs, in this paper, we focused on the symmetric linear B1A1B2A2B3 penta-BCP with fixed fB1 = fB3 and fA1 = fA2 to reveal the single stretched bridging effect on the phase behaviors.
It is generally believed that the self-assembly of symmetrical AB-linear multi-BCPs usually forms a symmetrical phase diagram, and that the dominant spherical phase is BCC rather than other spherical phases. However, our SCFT calculations demonstrate that the length ratio of middle bridging B2-block plays a key role in modifying the self-assembly behaviors of the symmetrical B1A1B2A2B3 multi-BCP. On the one hand, the BCC can be replaced by FCC completely as the length ratio of the middle bridging B2-block is within a specific value range. On the other hand, the symmetry of the phase diagram is broken by changing the relative length of the B2-block.
The following sections are arranged as follows. In Section 3, a detailed introduction of SCFT is presented. In Section 2, the results and a discussion are presented. First, the segment distribution of the relative length of the B2-block in lamellar phase is revealed. Then, we construct a series of phase diagrams in fAχN plane with fixed fA1 = fA2, fB1 = fB3 and several typical τ = 1/2, 2/3, 4/5 (τ = (fB1 + fB3)/fB) to reveal the effects of the fA and χN on the phase behaviors of the symmetric B1A1B2A2B3 pentablock copolymers. Further, we also prove that the relative ratio of B2-block can influence the phase behaviors with fixed χN = 100. Finally, we conclude this article in Section 4.

2. Result and Discussion

To make our results as reliable as possible, nine ordered structures in the multi-AB linear BCPs are considered in Figure 1. It should be noted that the hexagonal close packing sphere (HCP) is not considered in this paper, as it usually has an insignificant free energy difference compared with FCC and is located around the phase boundary [48].

2.1. Effect of Length Ratio of B2-Block on the Formation of L Structure

To the best of our knowledge, the A/B segment distributions have a marked effect on the formation of the ordered nanostructures [32]. First, we depict the possible segment distributions of the A/B block in the lamellar phases in Figure 2 to uncover the effect of length ratio of the bridging B2-block on the stability of the ordered lamellar nanostructures. Apparently, the A1- and A2- as well as B1- and B3-blocks have the same segment distributions, respectively, because of the symmetric architecture of the linear B1A1B2A2B3 penta-BCP with fixed fA1 = fA2 and fB1 = fB3. In general, the B1- and B3-blocks prefer to aggregate in the same domain with B2-block to form the B-domain in the lamellar phase shown in Figure 2b to reduce the loss of interfacial energy. Specifically, the B1- and B3-blocks aggregate into the different domain compared with B2-block as the length ratio of B2 block (1 − τ) is too long or too short, as shown in Figure 2a,c. In this case, a short B1/B3 (or B2) block will prefer to mix in the A-domain instead of the B-domain, leading to an extra interfacial energy cost but compensated with more entropy energy.
In Figure 3, the 1D segment distributions of the A/B block in lamellar phases are presented with three typical values τ = 0.1, 0.5, and 0.9, respectively, which are highly consistent with Figure 2. It is evidenced that the local phase segregation emerged regardless of whether the middle B2-block is too short or too long, resulting in the expanded domain spacing of the ordered nanostructures. Therefore, we can speculate regarding the segment distribution of A/B block in lamellar phases from the corresponding domain spacing shown in Figure 4. This means that the short B1/B3 and B2 block will aggregate into the A-domain for τ < 0.2 and τ > 0.8, respectively, while the B1/B3 and B2 blocks will aggregate into the same B-domain as the value of τ ranges from 0.2 to 0.8. This local segregation behavior has also been reported from asymmetric ABA and ABAB linear BCP systems [24,36]. Both experimental and theoretical research have confirmed that even a subtle change in the length ratio can strongly influence the phase transition.

2.2. Phase Diagram in fA-χN Plane with Three Typical τ

To reveal the effect of the middle bridging B2-block on the phase transition sequences, five typical values of τ (0, 0.1, 0.5, 0.8, 1) have been calculated for χN = 100 and fA < 0.5, as shown in Figure 5. Interestingly, the phase transition point between G and L changes from 0.314, 0.223, 0.356, 0.290, to 0.350 as the τ increases. These devious phase transition points can be attributed to the changing of the effective volume fraction of the A/B component relative to their characteristic values. Specifically, the short B2 (or B1/B3) blocks will prefer to mix in the A-domain for τ = 0.8 (τ = 0.1), resulting in an enlarged effective volume fraction of the A-component; thus, the phase boundary between G and L shifts to the left side, as shown in Figure 5b,d. For the same reason, the phase boundary between G and C6 phases have similar trends. What is more noteworthy is that the BCC phase is replaced by FCC for τ =0.8 in Figure 5b.
Without the loss of the generality, three phase diagrams in the fAχN plane with three typical τA = 1/2, 2/3, and 4/5 are constructed in Figure 6. Generally, the main features of these phase diagrams are similar to AB di-BCP as well as to ABA tri-BCP [32]. For τ = 1/2 and 2/3, the two phase diagrams are almost symmetrical, resulting from the symmetrical architecture of the AB linear multi-BCP and the uniform stretching of A- and B-blocks. As the τ increased to 4/5, the phase diagram becomes more asymmetric compared with the case of τ = 1/2 and 2/3, where the cylindrical and spherical phase regions expanded remarkably in the right part. This particular phase transition sequence is evidenced by the free energy comparisons of the candidate phases shown in Figure 7, indicating the phase sequence of L → G → C6 → FCC.
To uncover the underlying sophisticated phase transition mechanism between these spherical and cylindrical phases, not only the free energy, but also the corresponding interfacial energy and entropy energy comparisons, are plotted as a function of fA in Figure 8a–c, respectively. This shows that the stable C6 phase region is replaced by spherical phase for fA > 0.76, as the obtained interface energy from C6 phase cannot compensate the huge entropy loss. Figure 8a indicates that there is a very small free energy difference between the A15 and s phases, and the free energies of BCC and FCC phases are almost degenerative. The A15 and s phases have a favorable entropy energy while the loss of interfacial energy is larger than the BCC and FCC phases shown in Figure 8b,c, suggesting that the A15 and s phases are metastable in this symmetrical linear B1A1B2A2B3 penta-BCP system.

2.3. Phase Diagram in fA-τ Plane with Fixed cN

Generally, the phase diagrams in the fAτ plane with fixed χN = 100 is presented in Figure 9. From the perspective of the changing tendency of these different phase regions, there is a significant difference between the two-dimensional (2D) and three-dimensional (3D) structures. This implies that the phase region of the 2D cylindrical phase changes in the range of 0.028 < ΔfC < 0.179 is significant, while the changing of the phase regions of 3D gyroid (0.014 < ΔfG < 0.052) and spherical (0.006 < ΔfS < 0.073) phases are comparatively milder. Generally, the changing of the length ratio of the bridging B2-block will result in the non-uniform stretching of each B block, leading to the unfavorable entropy cost. Based on our understanding of self-assembly, the unfavorable stretching of each block is relieved more easily in 3D space compared with 2D space.
What is more noteworthy is the appearance of the phase transition sequence, i.e., BCC -> FCC-> BCC -> FCC-> BCC, shown in Figure 8 for fixed χN = 100. In order to uncover the phase transition mechanism between FCC and BCC structures, the detailed difference of free energy as well as entropy and interfacial energies are calculated and presented in Figure 10a,b, respectively. Generally, the FCC is less stable than the BCC, as the non-uniform stretching of the B-block leads to an extra packing frustration in the FCC. The FCC becomes stable over the BCC only when the gain of interfacial energy can compensate for the loss of entropy energy.
To further visualize and explore the stabilization mechanism of the FCC and BCC spherical phases, the isosurface plots of different A- and B-blocks in BCC and FCC spherical phases with six typical parameters are presented in Figure 11. The detailed parameters are listed below each subfigure. As the spherical domains can be formed from A- and B-components, respectively, these typical parameters can be divided into two categories (i.e., (a1) (b1) (c1) and (a2) (b2) (c2)). The first part is shown in the left panel, where the short A-blocks form spheres while the long B-blocks form the matrix. When τ ≈ 0.8 (Figure 11(a1)), the short B2-block aggregate near the A/B interface and the free end of the long B1- (and B3-) block can extend to the vertices of the Voronoi cell. The length difference between B2 and B1/B3 can reduce the packing frustration of B-blocks, thus stabilizing FCC. Similarly, when τ ≈ 0.3 (Figure 11(c1)), the short B1 /B3-blocks aggregate near the A/B interface, and the long B2-block extends to the vertices of the Voronoi cell. However, when τ ≈ 0.5 (Figure 11(b1)), fB2/2 ≈ fB1; therefore, none of the B blocks have the advantage of extending to the vertices of the Voronoi cell. As a result, the BCC is more stable than the FCC in this case. The second part is shown in the right panel, where the short (or part of) B- blocks form spheres and the long A- blocks form the matrix. When τ ≈ 0.8 (Figure 11(a2)), the short B2 block cannot be phase separated and dissolved in the A-matrix. However, as shown in Figure 11(a2), the concentration of B1 block increases at some special positions, thus reducing the interfacial energy. In the BCC phase, the B1-block prefers to aggregate at the edges of the Voronoi cell. In FCC phase, the B1-block prefers to aggregate at the vertices of the Voronoi cell, which is more concentrated than in the BCC. Therefore, the FCC phase is more stable than the BCC phase. The same is true in the case of τ ≈ 0.3 (Figure 11(c2)), since B1 and B3 blocks prefer to aggregate at the vertices of the Voronoi cell in FCC. When τ ≈ 0.5 (Figure 11(b2)), all B-blocks are long enough for phase separation from the A-domain. Therefore, the BCC phase is more stable than the FCC phase.

3. Theory and Method

In this study, the system consists of n B1A1B2A2B3 penta-BCP chains in the volume of V. Each penta-BCP consists of N segments with an equal segment length b and density ρ. The length of each block is specified as fB1N, fA1N, fB2N, fA2N, and fB3N (τ = (fB1 + fB3)/fB), respectively. The penta-BCP can be reduced to BAB or ABA, as τ equals 1 or 0. The variables ϕA(r) and ϕB(r) are used to characterize the distribution of the volume fraction of A-and B-blocks in the self-assembled mesoscopic structures. Under the approximation of the mean-field treatment and Gaussian chain model, the free energy per chain is in the unit of thermal energy kBT at the temperature T, and the kB is the Boltzmann constant, which can be expressed as:
F / nk B T = In Q + 1 V dr { χ N ϕ A ( r ) ϕ B ( r ) ω A ( r ) ϕ A ( r ) ω B ( r ) ϕ B ( r ) η ( r ) ( 1 ϕ A ( r ) ϕ B ( r ) ) }
where the ϕk(r) is the mean field conjugate to ϕk(r) (k = A or B). Moreover, η(r) is the Lagrange multiplier used to force the incompressibility condition: ϕ A ( r ) + ϕ B ( r ) = 1.0 . Moreover, the variable quantity Q is the partition function interacting with the mean field ωA and ωB produced by the surrounding chains, which is determined by
Q = 1 V dr   q ( r , s ) q + ( r , s )
where the q(r, s) and q+(r, s) are the propagator starting from the distinguishable ends, satisfying the following modifying diffusion equations:
q ( r , s ) s = 2 q ( r , s ) ω ( r , s ) q ( r , s )
q + ( r , s ) s = 2 q + ( r , s ) ω ( r , s ) q + ( r , s )
where the w(r,s) = ωk(r), where s belongs to K blocks (K = A, B). In the above equations, the R g = (N/6)1/2b is chosen as the unit length. The initial conditions of the propagator functions are q(r,0) = 1.0 and q+(r,1) = 1.0.
Minimization of the free energy leads to the following SCFT equation:
ω A ( r ) = χ N ϕ B ( r ) + η
ω B ( r ) = χ N ϕ A ( r ) + η
ϕ B ( r ) = 1 Q [ 0 f B 1 dsq ( r , s ) q + ( r , s ) + f B 1 + f A 1 f B 1 + f A 1 + f B 2 dsq ( r , s ) q + ( r , s ) + 1 f B 3 1 dsq ( r , s ) q + ( r , s ) ]
ϕ A ( r ) = 1 Q [ f B 1 f B 1 + f A 1 dsq ( r , s ) q + ( r , s ) + f B 1 + f A 1 + f B 2 f B 1 + f A 1 + f B 1 + f A 2 dsq ( r , s ) q + ( r , s ) ]
The above equations are solved using the pseudo-spectral method, and the Anderson-mixing is implemented to accelerate the calculation [48,49,50,51,52,53]. The rectangular cubic is chosen to solve the periodic boundary conditions, where the lattice length is less than 0.1 Rg and the chain contour is divided to 200 points.

4. Conclusions

In summary, the self-assembly behaviors of a symmetric linear B1A1B2A2B3 penta-BCP have been studied using the SCFT. Several interesting self-assembly behaviors have been observed in the phase diagrams with respect to the fAχN and fAτ planes. One of the most striking features is the emergency of alternating phase transition sequences between the FCC and BCC phases as τ increases, indicating that the length of the B2-bridging block can play a sophisticated role to regulate the phase behaviors. It has been demonstrated that interfacial energy is a dominant factor in the stabilization of the FCC over the BCC. Particularly, the 3D complex FK σ and A15 spherical phases are less stable than the BCC and FCC phases, as the gain of entropy energy cannot compensate for the loss of interfacial energy in this symmetric B1A1B2A2B3 penta-BCP. Additionally, the phase region of the O70 phase is extremely influenced due to the non-uniform stretching of the bridging B2-blcok. Our SCFT calculations reveal the importance of the middle bridging B2-block in regulating the phase behaviors of symmetric linear B1A1B2A2B3 penta-BCP.

Author Contributions

Conceptualization, B.Z.; methodology, B.Z.; calculation, W.Y. and B.Z.; resources, B.Z. and Y.X.; data analysis, B.Z. and Q.D.; plotting, Q.D. and B.Z.; writing—original draft, W.Y. and B.Z.; writing—review and editing, B.Z., Q.D. and Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

We gratefully acknowledge the funding support from the National Natural Science Foundation of China (grant 22203060 and 21973051) and Zhejiang Provincial Natural Science Foundation of China (grant LQ20A040005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Weihua Li for support and encouragement.

Conflicts of Interest

The authors declare no competing financial interest.

Sample Availability

Not applicable.

References

  1. Darling, S.B. Directing the Self-assembly of Block Copolymers. Prog. Polym. Sci. 2017, 32, 1152–1204. [Google Scholar] [CrossRef]
  2. Li, W.H.; Muller, M. Directed Self-Assembly of Block Copolymers by Chemical or Topographical Guiding Patterns: Optimizing Molecular Architecture, Thin-Film Properties, and Kinetics. Prog. Polym. Sci. 2016, 54–55, 47–75. [Google Scholar] [CrossRef]
  3. Mai, Y.; Eisenberg, A. Self-assembly of Block Copolymers. Chem. Soc. Rev. 2012, 41, 5969–5985. [Google Scholar] [CrossRef] [PubMed]
  4. Bates, C.M.; Bates, F.S. 50th Anniversary Perspective: Block Polymers-Pure Potential. Macromolecules 2017, 50, 3–22. [Google Scholar] [CrossRef]
  5. Thurn-Albrecht, T.; Schotter, J.; Kastle, C.A.; Emley, N.; Shibauchi, T.; Krusin-Elbaum, L.; Guarini, K.; Black, C.T.; Tuominen, M.T.; Russell, T.P. Ultrahigh-density Nanowire Arrays Grown in Self-assembled Diblock Copolymer Templates. Science 2000, 290, 2126–2129. [Google Scholar] [CrossRef]
  6. Crossland, E.J.; Kamperman, M.; Nedelcu, M.; Ducati, C.; Wiesner, U.; Smilgies, D.M.; Toombes, G.E.; Hillmyer, M.A.; Ludwigs, S.; Steiner, U.; et al. A Bicontinuous Double Gyroid Hybrid Solar Cell. Nano Lett. 2009, 9, 2807–2812. [Google Scholar]
  7. Hampu, N.; Werber, J.R.; Chan, W.Y.; Feinberg, E.C.; Hillmyer, M.A. Next-Generation Ultrafiltration Membranes Enabled by Block Polymers. ACS Nano 2020, 14, 16446–16471. [Google Scholar] [CrossRef]
  8. Gong, C.; Sun, S.; Zhang, Y.; Sun, L.; Su, Z.; Wu, A.; Wei, G. Hierarchical Nanomaterials via Biomolecular Self-assembly and Bioinspiration for Energy and Environmental Applications. Nanoscale 2019, 11, 4147–4182. [Google Scholar] [CrossRef]
  9. Morgan, S.; Stefan, G.; Silvia, V.; Ulrich, W.; Ullrich, S. Block Copolymer Self-assembly for Nanophotonics. Chem. Soc. Rev. 2015, 44, 5076–5091. [Google Scholar]
  10. Gaucher, G.; Dufresne, M.H.; Sant, V.P.; Kang, N.; Maysinger, D.; Leroux, J.C. Block Copolymer Micelles: Preparation, Characterization and Application in Drug Delivery. J. Control. Release 2005, 109, 169–188. [Google Scholar] [CrossRef]
  11. Kaori, M.T.; Yuichi, Y.; Anjaneyulu, D.; Sorato, I.; Theofilus, A.; Tockary, K.T.; Kensuke, O.; Kazunori, K. Effect of Shear Stress on Structure and Function of Polyplex Micelles from Poly(ethylene glycol)-poly(L-lysine) Block Copolymers as Systemic Gene Delivery Carrier. Biomaterials 2017, 126, 31–38. [Google Scholar]
  12. Matsen, M.W.; Schick, M. Stable and Unstable Phases of a Diblock Copolymer Melt. Phys. Rev. Lett. 1994, 72, 2660–2663. [Google Scholar] [CrossRef]
  13. Bates, F.S.; Fredrickson, G.H. Block Copolymers-Designer Soft Materials. Phys. Today 1999, 52, 32–38. [Google Scholar] [CrossRef]
  14. Matsen, M.W.; Bates, F.S. Origins of Complex Self-Assembly in Block Copolymers. Macromolecules 1996, 29, 7641–7644. [Google Scholar] [CrossRef]
  15. Tyler, C.A.; Morse, D.C. Orthorhombic Fddd Network in Triblock and Diblock Copolymer Melts. Phys. Rev. Lett. 2005, 94, 208302. [Google Scholar] [CrossRef] [PubMed]
  16. Jinnai, H.; Kaneko, T.; Matsunaga, K.; Abetz, C.; Abetz, V. A Double Helical Structure Formed from an Amorphous, Achiral ABC Triblock Terpolymer. Soft Matter 2009, 5, 2042–2046. [Google Scholar] [CrossRef]
  17. Dobriyal, P.; Xiang, H.Q.; Kazuyuki, M.; Chen, J.T.; Jinnai, H.; Russell, T.P. Cylindrically Confined Diblock Copolymers. Macromolecules 2009, 42, 9082–9088. [Google Scholar] [CrossRef]
  18. Yu, B.; Sun, P.C.; Chen, T.H.; Jin, Q.H.; Ding, D.T.; Li, B.H.; Shi, A.-C. Confinement-Induced Novel Morphologies of Block Copolymers. Phys. Rev. Lett. 2006, 96, 138306. [Google Scholar] [CrossRef]
  19. Li, W.H.; Qiu, F.; Shi, A.-C. Emergence and Stability of Helical Superstructures in ABC Triblock Copolymers. Macromolecules 2011, 45, 503–509. [Google Scholar] [CrossRef]
  20. Hayashida, K.; Dotera, T.; Takano, A.; Matsushita, Y. Polymeric Quasicrystal: Mesoscopic Quasicrystalline Tiling in ABC Star Polymers. Phys. Rev. Lett. 2007, 98, 195502. [Google Scholar] [CrossRef]
  21. Miyamori, Y.; Suzuki, J.; Takano, A.; Matsushita, Y. Periodic and Aperiodic Tiling Patterns from a Tetrablock Terpolymer System of the A1BA2C Type. ACS Macro Lett. 2020, 9, 32–37. [Google Scholar] [CrossRef] [PubMed]
  22. Duan, C.; Zhao, M.T.; Qiang, Y.C.; Chen, L.; Li, W.H.; Qiu, F.; Shi, A.-C. Stability of Two-Dimensional Dodecagonal Quasicrystalline Phase of Block Copolymers. ACS Macro Lett. 2018, 51, 7713–7721. [Google Scholar] [CrossRef]
  23. Nap, R.; Sushko, N.; Erukhimovich, I.; Ten Brinke, G. Double Periodic Lamellar-in-Lamellar Structure in Multiblock Copolymer Melts with Competing Length Scales. Macromolecules 2006, 39, 6765–6770. [Google Scholar] [CrossRef]
  24. Fleury, G.; Bates, F.S. Perpendicular Lamellae in Parallel Lamellae in a Hierarchical CECEC-P Hexablock Terpolymer. Macromolecules 2009, 42, 1691–1694. [Google Scholar] [CrossRef]
  25. Zhang, L.; Lin, J. Hierarchically Ordered Nanocomposites Self-Assembled from Linear-Alternating Block Copolymer/Nanoparticle Mixture. Macromolecules 2009, 42, 1410–1414. [Google Scholar] [CrossRef]
  26. Xu, Y.C.; Qian, M.S. Formation of Perpendicular Three-Dimensional Network Nanostructures in ABC-Star Copolymers. Langmuir 2022, 38, 7889–7897. [Google Scholar]
  27. Lee, S.; Bluemle, M.J.; Bates, F.S. Discovery of a Frank-Kasper σ Phase in Sphere-Forming Block Copolymer Melts. Science 2010, 330, 349–353. [Google Scholar] [CrossRef]
  28. Schulze, M.W.; Lewis, R.M.; Lettow, J.H.; Hickey, R.J.; Timothy, M.G.; Marc, A.H.; Bates, F.S. Conformational Asymmetry and Quasicrystal Approximants in Linear Diblock Copolymers. Phys. Rev. Lett. 2017, 118, 207801. [Google Scholar] [CrossRef] [PubMed]
  29. Xie, N.; Li, W.H.; Qiu, F.; Shi, A.-C. σ Phase Formed in Conformationally Asymmetric AB-Type Block Copolymers. ACS Macro Lett. 2014, 3, 906–910. [Google Scholar] [CrossRef]
  30. Liu, M.J.; Qiang, Y.C.; Li, W.H.; Qiu, F.; Shi, A.-C. Stabilizing the Frank-Kasper Phases via Binary Blends of AB Diblock Copolymers. ACS Macro Lett. 2016, 5, 1167–1171. [Google Scholar] [CrossRef] [PubMed]
  31. Dorfman, K.D. Frank–Kasper Phases in Block polymers. Macromolecules 2021, 54, 10251–10270. [Google Scholar] [CrossRef]
  32. Matsen, M.W. Effect of Architecture on the Phase Behavior of AB-Type Block Copolymer Melts. Macromolecules 2012, 45, 2161–2165. [Google Scholar] [CrossRef]
  33. Sakurai, S.; Shirouchi, K.; Munakata, S.; Kurimura, H.; Suzuki, S.; Watanabe, J.; Oda, T.; Shimizu, N.; Tanida, K.; Yamamoto, K. Morphology Reentry with a Change in Degree of Chain Asymmetry in Neat Asymmetric Linear A1BA2 Triblock Copolymers. Macromolecules 2017, 50, 8647–8657. [Google Scholar] [CrossRef]
  34. Matsen, M.W. Equilibrium Behavior of Asymmetric ABA Triblock Copolymer Melts. J. Chem. Phys. 2000, 113, 5539–5544. [Google Scholar] [CrossRef]
  35. Ahn, S.; Kim, J.K.; Zhao, B.; Duan, C.; Li, W.H. Morphology Transitions of Linear A1B1A2B2 Tetrablock Copolymers at Symmetric Overall Volume Fraction. Macromolecules 2018, 51, 4415–4421. [Google Scholar] [CrossRef]
  36. Zhao, B.; Jiang, W.B.; Chen, L.; Li, W.H.; Qiu, F.; Shi, A.-C. Emergence and Stability of a Hybrid Lamella-Sphere Structure from Linear ABAB Tetrablock Copolymers. ACS Macro Lett. 2018, 7, 95–99. [Google Scholar] [CrossRef]
  37. Matsen, M.W. The Standard Gaussian Model for Block Copolymer Melts. J. Phys. Condens. Matter 2002, 14, R21–R47. [Google Scholar] [CrossRef]
  38. Drolet, F.; Fredrickson, G.H. Combinatorial Screening of Complex Block Copolymer Assembly with Self-Consistent Field Theory. Phys. Rev. Lett. 1999, 83, 4317. [Google Scholar] [CrossRef]
  39. Matsen, M.W. Self-Consistent Field Theory and Its Applications. In Soft Matter; Polymer Melts and Mixtures; Gompper, G., Schick, M., Eds.; Wiley-VCH: Weinheim, Germany, 2007; Volume 1, pp. 87–178. [Google Scholar]
  40. Grason, G.M.; DiDonna, B.A.; Kamien, R.D. Geometric Theory of Diblock Copolymer Phases. Phys. Rev. Lett. 2003, 91, 058304. [Google Scholar] [CrossRef]
  41. Xie, N.; Liu, M.J.; Deng, H.L.; Li, W.H.; Qiu, F.; Shi, A.-C. Macromolecular Metallurgy of Binary Mesocrystals via Designed Multiblock Terpolymers. J. Am. Chem. Soc. 2014, 136, 2974–2977. [Google Scholar] [CrossRef]
  42. Zhao, B.; Wang, C.; Chen, Y.; Liu, M.J. Frank–Kasper Phases Self-Assembled from a Linear A1B1A2B2 Tetrablock Copolymer. Langmuir 2021, 37, 5642–5650. [Google Scholar] [CrossRef] [PubMed]
  43. Xu, Z.W.; Li, W.H. Control the Self-assembly of Block Copolymers by Tailoring the Packing Frustration. Chin. J. Chem. 2022, 40, 1083–1090. [Google Scholar]
  44. Gao, Y.; Deng, H.L.; Li, W.H.; Qiu, F.; Shi, A.-C. Formation of Nonclassical Ordered Phases of AB-Type Multiarm Block Copolymers. Phys. Rev. Lett. 2016, 116, 068304. [Google Scholar] [CrossRef] [PubMed]
  45. Xie, Q.; Qiang, Y.C.; Li, W.H. Single Gyroid Self-Assembled by Linear BABAB Pentablock Copolymer. ACS Macro Lett. 2022, 11, 205–209. [Google Scholar] [CrossRef]
  46. Lynd, N.A.; Oyerokun, F.T.; O’Donoghue, D.L.; Handlin, D.L., Jr.; Fredrickson, G.H. Design of Soft and Strong Thermoplastic Elastomers Based on Nonlinear Block Copolymer Architectures Using Self-Consistent-Field Theory. Macromolecules 2010, 43, 3479–3486. [Google Scholar] [CrossRef]
  47. Chen, L.; Qiang, Y.C.; Li, W.H. Tuning Arm Architecture Leads to Unusual Phase Behaviors in a (BAB)5 Star Copolymer Melt. Macromolecules 2018, 51, 9890–9900. [Google Scholar] [CrossRef]
  48. Matsen, M.W. Fast and Accurate SCFT Calculations for Periodic Block Copolymer Morphologies Using the Spectral Method with Anderson Mixing. Eur. Phys. J. E 2009, 30, 361–369. [Google Scholar] [CrossRef] [PubMed]
  49. Shi, A.-C. Developments in Block Copolymer Science and Technology; John Wiley & Sons, Ltd.: New York, NY, USA, 2004; pp. 265–293. [Google Scholar]
  50. Fredrickson, G.H. The Equilibrium Theory of Inhomogeneous Polymers; Oxford University Press: New York, NY, USA, 2006. [Google Scholar]
  51. Tzeremes, G.; Rasmussen, K.Ø.; Lookman, T.; Saxena, A. Efficient Computation of the Structural Phase Behavior of Block Copolymers. Phys. Rev. E 2002, 65, 041806. [Google Scholar] [CrossRef]
  52. Rasmussen, K.Ø.; Kalosakas, G. Improved Numerical Algorithm for Exploring Block Copolymer Mesophases. J. Polym. Sci. Part B Polym. Phys. 2002, 40, 1777–1783. [Google Scholar] [CrossRef]
  53. Thompson, R.; Rasmussen, K.; Lookman, T. Improved Convergence in Block Copolymer Self-Consistent Field Theory by Anderson Mixing. J. Chem. Phys. 2004, 120, 31–34. [Google Scholar] [CrossRef]
Figure 1. Schematic of candidate ordered phases considered in this work, including the lamellar (L) phase and networks of gyroid (G) and Fddd (O70) phases, two different cylinder phases (C6, C4), and four different spherical phases (BCC, FCC, A15, σ).
Figure 1. Schematic of candidate ordered phases considered in this work, including the lamellar (L) phase and networks of gyroid (G) and Fddd (O70) phases, two different cylinder phases (C6, C4), and four different spherical phases (BCC, FCC, A15, σ).
Molecules 28 03536 g001
Figure 2. The plots of possible distributions of A- and B-blocks in lamellar phase, A and B blocks are shown in Red and Blue color respectively. (a) The short B1-(B3) and long B2-blocks segregate into the different domains; (b) The B1-(B3) and B2-blocks segregate within the same domains; (c) The long B1-(B3) and short B2-blocks segregate into the different domains.
Figure 2. The plots of possible distributions of A- and B-blocks in lamellar phase, A and B blocks are shown in Red and Blue color respectively. (a) The short B1-(B3) and long B2-blocks segregate into the different domains; (b) The B1-(B3) and B2-blocks segregate within the same domains; (c) The long B1-(B3) and short B2-blocks segregate into the different domains.
Molecules 28 03536 g002
Figure 3. The comparisons of density profiles in the lamellar phase for three typical τ = 0.1, 0.5, and 0.9 shown in (ac), respectively, where the B1/B3 block, A1/A2 block, and B2 block are shown in blue solid line with triangle symbol, red solid line with circle symbol, and blue solid line with square symbol, respectively.
Figure 3. The comparisons of density profiles in the lamellar phase for three typical τ = 0.1, 0.5, and 0.9 shown in (ac), respectively, where the B1/B3 block, A1/A2 block, and B2 block are shown in blue solid line with triangle symbol, red solid line with circle symbol, and blue solid line with square symbol, respectively.
Molecules 28 03536 g003aMolecules 28 03536 g003b
Figure 4. The period of lamellar phase as a function of τ for fixed χN = 100 and fA = 0.5.
Figure 4. The period of lamellar phase as a function of τ for fixed χN = 100 and fA = 0.5.
Molecules 28 03536 g004
Figure 5. The phase transition sequences of the ordered nanostructures with fixed χN = 100 and fA < 0.5 for several typical τ = 0, 0.1, 0.5, 0.8, and 1, respectively. The blue symbol indicate the phase transiton point between disorder and spherical phases, green symbol indicate the phase transiton point between spherical and cylindrical phases, red symbol indicate the phase transiton point between cylindrical and gyroid phases, and black symbol indicate the phase transiton point between gyroid and lamellar phases.
Figure 5. The phase transition sequences of the ordered nanostructures with fixed χN = 100 and fA < 0.5 for several typical τ = 0, 0.1, 0.5, 0.8, and 1, respectively. The blue symbol indicate the phase transiton point between disorder and spherical phases, green symbol indicate the phase transiton point between spherical and cylindrical phases, red symbol indicate the phase transiton point between cylindrical and gyroid phases, and black symbol indicate the phase transiton point between gyroid and lamellar phases.
Molecules 28 03536 g005
Figure 6. Phase diagrams in fAχN plane with three typical τ = 1/2, 2/3, and 4/5 (ac), respectively.
Figure 6. Phase diagrams in fAχN plane with three typical τ = 1/2, 2/3, and 4/5 (ac), respectively.
Molecules 28 03536 g006aMolecules 28 03536 g006b
Figure 7. Comparisons of free energy along changing fA of the candidate phases with fixed τ = 0.8 and χN = 80, where the free energy of C6 is treated as the reference.
Figure 7. Comparisons of free energy along changing fA of the candidate phases with fixed τ = 0.8 and χN = 80, where the free energy of C6 is treated as the reference.
Molecules 28 03536 g007
Figure 8. Comparisons of (a) free energy, (b) interfacial energy, (c) entropy energy along changing fA of the part of the candidate phases with fixed τ = 0.8 and χN =80, where the FCC is treated as the reference.
Figure 8. Comparisons of (a) free energy, (b) interfacial energy, (c) entropy energy along changing fA of the part of the candidate phases with fixed τ = 0.8 and χN =80, where the FCC is treated as the reference.
Molecules 28 03536 g008
Figure 9. Phase diagram in fAτ plane for B1A1B2A2B3 pentablock copolymer with fixed χN = 100.
Figure 9. Phase diagram in fAτ plane for B1A1B2A2B3 pentablock copolymer with fixed χN = 100.
Molecules 28 03536 g009
Figure 10. (a) Free energy, (b) interfacial and entropy energies comparison between BCC and FCC as a function of τ (0.1 < τ < 0.9) with fixed χN = 100 and fA = 0.14, where the free energy of BCC is treated as the reference.
Figure 10. (a) Free energy, (b) interfacial and entropy energies comparison between BCC and FCC as a function of τ (0.1 < τ < 0.9) with fixed χN = 100 and fA = 0.14, where the free energy of BCC is treated as the reference.
Molecules 28 03536 g010aMolecules 28 03536 g010b
Figure 11. The isosurface plots of different blocks in BCC and FCC phases, where the value of fA, τ, and isovalues (Iso) are listed below each subfigure with fixed χN = 100 for all the figures. In the left panel, the isosurface of A- and B- blocks are presented for three typical τ ( 0.8, 0.5 and 0.3) shown in (a1,b1,c1), where the spherical domain formed from A-component. In the right panel, the isosurface of A- and B- blocks are presented for three typical τ ( 0.8, 0.5 and 0.3) shown in (a2,b2,c2), where the spherical domain formed from B-component.
Figure 11. The isosurface plots of different blocks in BCC and FCC phases, where the value of fA, τ, and isovalues (Iso) are listed below each subfigure with fixed χN = 100 for all the figures. In the left panel, the isosurface of A- and B- blocks are presented for three typical τ ( 0.8, 0.5 and 0.3) shown in (a1,b1,c1), where the spherical domain formed from A-component. In the right panel, the isosurface of A- and B- blocks are presented for three typical τ ( 0.8, 0.5 and 0.3) shown in (a2,b2,c2), where the spherical domain formed from B-component.
Molecules 28 03536 g011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, B.; Dong, Q.; Yang, W.; Xu, Y. Theoretical Study of Phase Behaviors of Symmetric Linear B1A1B2A2B3 Pentablock Copolymer. Molecules 2023, 28, 3536. https://doi.org/10.3390/molecules28083536

AMA Style

Zhao B, Dong Q, Yang W, Xu Y. Theoretical Study of Phase Behaviors of Symmetric Linear B1A1B2A2B3 Pentablock Copolymer. Molecules. 2023; 28(8):3536. https://doi.org/10.3390/molecules28083536

Chicago/Turabian Style

Zhao, Bin, Qingshu Dong, Wei Yang, and Yuci Xu. 2023. "Theoretical Study of Phase Behaviors of Symmetric Linear B1A1B2A2B3 Pentablock Copolymer" Molecules 28, no. 8: 3536. https://doi.org/10.3390/molecules28083536

Article Metrics

Back to TopTop