36
Provided for non-commercial research and educational use. Not for reproduction, distribution or commercial use. This chapter was originally published in Polymer Science: A Comprehensive Reference published by Elsevier, and the attached copy is provided by Elsevier for the author's benefit and for the benefit of the author's institution, for non- commercial research and educational use including without limitation use in instruction at your institution, sending it to specific colleagues who you know, and providing a copy to your institution's administrator. All other uses, reproduction and distribution, including without limitation commercial reprints, selling or licensing copies or access, or posting on open internet sites, your personal or institution's website or repository, are prohibited. For exceptions, permission may be sought for such use through Elsevier's permissions site at: http://www.elsevier.com/locate/permissionusematerial Genzer J, Khalatur PG, and Khokhlov AR (2012) Conformation-Dependent Design of Synthetic Functional Copolymers. In: Matyjaszewski K and Möller M (eds.) Polymer Science: A Comprehensive Reference, Vol 6, pp. 689723. Amsterdam: Elsevier BV. © 2012 Elsevier B.V. All rights reserved.

Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Provided for non-commercial research and educational use.Not for reproduction, distribution or commercial use.

This chapter was originally published in Polymer Science: A ComprehensiveReference published by Elsevier, and the attached copy is provided by Elsevier for

the author's benefit and for the benefit of the author's institution, for non-commercial research and educational use including without limitation use in

instruction at your institution, sending it to specific colleagues who you know, andproviding a copy to your institution's administrator.

All other uses, reproduction and distribution, including without limitationcommercial reprints, selling or licensing copies or access, or posting on open

internet sites, your personal or institution's website or repository, are prohibited.For exceptions, permission may be sought for such use through Elsevier's

permissions site at:

http://www.elsevier.com/locate/permissionusematerial

Genzer J, Khalatur PG, and Khokhlov AR (2012) Conformation-Dependent Designof Synthetic Functional Copolymers. In: Matyjaszewski K and Möller M (eds.)

Polymer Science: A Comprehensive Reference, Vol 6, pp. 689–723.Amsterdam: Elsevier BV.

© 2012 Elsevier B.V. All rights reserved.

Page 2: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

6.18 Conformation-Dependent Design of Synthetic Functional Copolymers J Genzer, North Carolina State University, Raleigh, NC, USA PG Khalatur, Ulm University, Ulm, Germany; Russian Academy of Sciences, Moscow, Russia AR Khokhlov, Moscow State University, Moscow, Russia

© 2012 Elsevier B.V. All rights reserved.

6.18.1 Introduction 689 6.18.2 Theoretical Approaches 690 6.18.2.1 Two Paradigms in Sequence Design 690 6.18.2.2 How Can We Characterize Copolymer Sequence? 690 6.18.2.2.1 HP model 690 6.18.2.2.2 Long-range correlations versus short-range correlations 691 6.18.2.2.3 Information complexity 692 6.18.2.3 Conformation-Dependent Sequence Design: Structure Dictates Sequence 692 6.18.2.3.1 Polymer-analogous transformations 693 6.18.2.3.2 Copolymerization under heterogeneous conditions 702 6.18.2.3.3 Evolutionary approach and theoretically informed sequence design 709 6.18.3 Synthesis of Designed Copolymers 714 6.18.3.1 ‘Grafting through’ Approach 714 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous Reactions 715 6.18.4 Concluding Remarks 717 References 720

6.18.1 Introduction

Over the past few decades, we have witnessed increasing inter-est in developing new types of synthetic copolymers via design of their comonomer sequences. Most of the physical and mechanical properties of copolymers, which comprise two or more covalently bonded sequences of chemically distinct monomer species, depend not only on the chemical composi-tion but also on the arrangement of the comonomers in the polymer chain. There may be significant differences, for exam-ple, between two polymer systems with the same chemical composition, but different comonomer distributions (i.e., one of which has the comonomers distributed randomly in the chain, while the other has long blocks of each monomer type). In many cases, just the copolymer sequence dictates the structure and properties of the copolymer. To illustrate this, we will mention only two familiar examples.

Synthetic block copolymers (BCPs) can self-assemble into highly ordered supramolecular structures, showing a rich mor-phological behavior.1 Depending on the architecture, chemical composition, chain flexibility, and mutual interaction of the chemically distinct segments, BCPs can be tailored in to order spontaneously, via a thermally driven microphase separation process, form a number of equilibrium morphologies – periodic arrays of domains with tunable dimensions – such as body-centered-cubic (spherical), hexagonal (cylindrical), lamellar, and bicontinuous (gyroid) mesophases. These ‘weak crystalline’ structures, which are ‘coded’ in the monomer sequence distribution of single macromolecules, represent excellent candidates for preparing functional nanostructured materials with characteristic length scales from 1 to 103 nm in a bottom-up procedure. In particular, those systems exhibiting long-range periodicity have shown feasibility for many

applications including, but not limited to, a new generation of electronic and optoelectronic nanodevices, quantum dots, and catalyst arrays for nanotube growth.2–5 However, the reali-zation of this potential depends to a large extent on our ability to control the lengths of blocks and their distribution along the chain.

Typically, proteins fold to organize a very specific globular conformation, known as the protein’s native state, which is, in general, reasonably stable and unique. It is this well-defined three-dimensional conformation of a polypeptide chain that determines the macroscopic properties and function of a pro-tein. The folding mechanism and biological functionality are related directly to polypeptide sequence; a completely random amino acid sequence is unlikely to form a functional structure. In this view, polypeptide sequence forces a protein to be more than a collapsed heteropolymer, but rather to assume a highly specific three-dimensional structure. Hence, a fundamental issue is how functional protein sequences, which determine biologically active structures, differ from random sequences. Understanding the relationship between a protein’s sequence and its native structure is one of the hallmarks in modern science.6–8

While some synthetic copolymers – alternating, di-, tri-, and multiblock copolymers – possess an ordered-sequence distri-bution of monomeric units along the chain backbone, random copolymers (RCPs) represent a special class of macromole-cules, which possess a disordered monomer sequence distribution. A series of papers have been published which established that tuning the copolymer chemical composition and comonomer sequence distribution profoundly affects the characteristics of RCPs.9–12 Of those characteristics intrinsic to the polymeric materials, most work performed to date addressed the role of the chemical composition. By judiciously

Polymer Science: A Comprehensive Reference, Volume 6 doi:10.1016/B978-0-444-53349-4.00176-X 689

Page 3: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy690 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

choosing the polymerizing comonomers, RCPs have the poten-tial to act, for example, as polymer blend compatibilizers13–15

22–24or adhesion promoters,16–21 to mention a few. Bratko et al. investigated the bulk properties of RCPs and observed that the tunability of the polymer’s blockiness may endow RCPs with the ability to act as pattern recognition agents for similarly patterned surfaces.

Having good control over both the chemical composition and the comonomer sequence distribution in RCPs is essential for many applications. While tailoring the content of comono-mers in RCPs is relatively straightforward, adjusting the distribution of monomers along the polymer backbone is not. Today, the majority of all polymeric materials is produced using the free-radical polymerization technique.25–31 For a limited range of comonomers, anionic and cationic polymer-izations are also used.32,33 Unfortunately, however, in conventional free-radical copolymerization in solution, control of the incorporation of monomer species into a copolymer chain is practically impossible. Furthermore, in this process, the propagating macroradicals usually attach monomeric units in a random way, governed by the relative reactivities of poly-merizing comonomers. This lack of control confines the versatility of the free-radical process, because the microscopic polymer properties, such as chemical composition distribution and tacticity are key parameters that determine the macroscopic behavior of the resultant product.

The absence of control of the incorporation of mono-mers into the polymeric chain implies that many macroscopic properties cannot be influenced to a large extent. Therefore, much effort has been directed toward the development of controlled radical polymerization (CRP) methods for the preparation of various copolymers (for a review, see Reference 31). CRPs offer the possibility of producing polymers with relatively well-defined proper-ties, while at the same time maintaining the simplicity of radical processes.34 These methods are based on the idea of establishing equilibrium between the active and dormant species in solution phase. In particular, the methods include three major techniques called stable free-radical polymerization, degenerative chain transfer technique,31

and atom transfer radical polymerization, pioneered by 35 36Ando et al. and Matyjaszewski et al. Although such

syntheses pose significant technical problems, these difficul-ties have all been successively overcome in the past few years. Nevertheless, the procedure of preparation of the resultant copolymers with controlled monomer sequence distribution remains somewhat complicated.

Summarizing the above, presently there is no comprehen-sive understanding of the interrelation between chemical sequences in synthetic copolymers and the conditions of their synthesis. One has to glance merely at recent literature in synthetic polymer science to realize that the problem of sequence–property relationship is by no means studied com-prehensively. In this contribution, we will discuss some new synthetic strategies, which have been developed recently to provide effective control of the chemical sequences. Our main focus will be on RCPs with tunable sequences.

Apart from the introductory section, the chapter is subdi-vided into two major parts: ‘theoretical approaches’ and ‘synthesis of designed copolymers’.

6.18.2 Theoretical Approaches 6.18.2.1 Two Paradigms in Sequence Design

Broadly speaking, sequence design may be defined as an approach aimed at finding an optimum sequence that provides desired properties of the resultant polymer. This requires a scoring function that may typically be based on physical prin-ciples, knowledge-based approaches, or a specifically designed function. Alas, insofar as the terms ‘sequence design’ and ‘sequence engineering’ imply, a rational, planned approach to the creation or modification of copolymer structure and func-tion, both still remain beyond our capabilities in a general way.

There are two main paradigms in the sequence design pro-blem. In protein science, the de novo sequence design problem consists of finding a sequence of amino acids that fold into a target globular structure. This problem is sometimes called the inverse protein folding problem. Many current methods for de novo protein sequence design consist of numerically mutat-ing a sequence until a maximum stability is achieved for the target structure that is usually considered as a ground state. There are a number of excellent reviews that cover this sub-ject.37–42 In synthetic polymer chemistry, emphasis is on the development of new methods of synthesis, on the control of (co)polymer stereochemistry and architecture, and on the design of high-performance polymeric materials tailored for specific uses and properties. The difference between the two sequence design concepts is related to several essential differ-ences between natural and nonnatural copolymers. We mention here only a few of them.

The order of amino acids in a polypeptide chain produced by the synthetic apparatus of the living cell is always the same for a particular protein so that all the protein sequences of a given type are structurally identical copies in every cell in a living organism. We cannot distinguish one individual protein sequence from another. For most synthetic copolymers, pro-duced industrially or synthesized in research laboratories, the occurrence of a certain degree of sequence disorder is inevita-ble. Therefore, if we are speaking about a synthetic copolymer sequence, we mean, explicitly or implicitly, that averaging over many different sequences has been carried out. For a protein to function, it must be in its highly specific native conformation that is stable only in a narrow temperature region. Secondary and tertiary structural motifs are the central issues to protein structure. In contrast, the properties and functions of synthetic copolymers are not so tightly related to their conformation. Moreover, we are mainly interested in the nonunique polymer conformations.

This list is far from exhaustive, but should rather be taken as simply a set of examples. In the present chapter, we will deal mainly with synthetic macromolecules and practically will not touch on biopolymers.

6.18.2.2 How Can We Characterize Copolymer Sequence?

6.18.2.2.1 HP model There are many kinds of monomers used to make up copoly-mers. These differ in physical and chemical properties. One of the most important differences (essential features) is their solubility, that is, how much they like or dislike a given solvent (e.g., water). For example, while protein structure is generally determined by permutations of the 20 amino acids, the

Page 4: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 691

primary distinction between amino acid side chains can be regarded from the very rough viewpoint to be depended only on their hydrophobicity.43 This proposition is supported by de novo protein design experiments showing that the binary pattern of hydrophilic (more polar) and hydrophobic (less polar) amino acids plays a significant role in secondary44 and tertiary45 structure determination. Indeed, the most important difference between the monomeric units of globular proteins is that some amino acid residues are hydrophobic, while others are hydrophilic or charged.46–50 Following this line, simple binary monomer alphabets have been widely exploited in simulation studies of the thermodynamics and stability of folded lattice proteins42,51,52 and random heteropolymer globules.53,54 Hence, the chemical and atomistic details of different monomeric units may not be necessary to understand many basic properties of copolymers, including synthetic ‘two-letter’ copolymers.

In what follows, we will mainly use the so-called HP model51 that distinguishes only two kinds of monomeric units, namely hydrophobic (H) and hydrophilic (or polar, P). This coarse-grained model of a linear hydrophobic/hydrophilic macromolecule reflects the spirit of minimalist models, in that it is simple yet based on a physical principle.

Despite the apparent simplicity of the HP model, the infor-mation complexity of an N-unit sequence encoded only by H and P symbols can be arbitrarily high for large N (strictly speaking, in the N → ∞ limit). Indeed, using the two-letter ‘•–’ Morse alphabet, we can write the whole human history in a message of reasonable length.

6.18.2.2.2 Long-range correlations versus short-range correlations Diblock and repeated-block copolymers are the simplest exam-ples of binary copolymers made up of two different monomer

species. More sophisticated distributions of chemically differ-ent groups along the chain are characteristic of random and random-block copolymers, including uncorrelated or ideal RCPs and so-called correlated RCPs. In the former class, the corresponding chemical sequences are uncorrelated that corre-sponds to Bernoulli or zeroth-order Markov processes.55 In the latter class, the correlation in the sequences of both types of segments is defined by means of a first-order Markov process. It is important to emphasize that in both cases the correlations characterizing the distribution of monomeric units along copo-lymer chains decay exponentially. The type of primary structure that emerges in these copolymers can be characterized as ‘ran-dom with short-range correlations (SRCs)’. Such correlations will always show up after the copolymerization process, if the probability of addition of chemically distinct units to the grow-ing chain depends on the type of the unit, which was added on the previous polymerization step.31 There are, however, such copolymers for which this is not the case. In this review, we will consider just these copolymers. To clarify the point in question, we will undertake again a brief excursion into biology.

The key biological macromolecules – proteins, DNA, and RNA – are responsible for functions, which are incomparably more complex and diverse than the functions, which we are normally discussing for synthetic polymers. The molecular basis for this ability to perform sophisticated functions is asso-ciated with unique chemical sequences of these biopolymers. In particular, a protein sequence as a whole determines the globular conformation and hence biological function. If this sequence is cut into two pieces, those pieces normally neither correspond to a soluble globule nor have any biological func-tion (Figures 1(a) and 1(b)). The same is true for statistically complex DNA sequences, which encode in a four-letter alphabet all genetic information and exhibit significant correla-tion on different scales.56

(a)

(c) (d)

(b)

Figure 1 (a, b) If a protein sequence is cut into two pieces, the pieces neither correspond to a native globule nor have any biological function. (c, d) The sequence distribution for rather large sections of a synthetic RCP is practically identical to that of the whole polymer.

Page 5: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

� �

� �

692 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

All these peculiarities are connected to long-range correlation (LRC) effects, and the corresponding sequence, which cannot be divided into shorter subsequences with similar statistical pat-terns and functional features, may be termed an ‘inseparable sequence’. Such sequence integrity and LRCs are not character-istic of the majority of synthetic linear copolymers, the primary structure of which is chemically homogeneous on a large scale. Indeed, the sequence distribution for rather large sections of synthetic copolymers is normally practically identical to that of the whole polymer (Figures 1(c) and 1(d)).

Thus, there are two fundamentally different classes of sequences: those that exhibit LCR and those that do not possess this property.

It is instructive to discuss briefly the quantities used to measure LRC. To monitor the long-range statistical properties of sequences, the method developed by Peng et al.57,58 in their search for LRC in DNA sequences is usually employed. In this approach, each HP copolymer sequence is transformed into a sequence of +1 and −1 symbols, which are considered as steps of a one-dimensional random walk. Shifting the sliding win-dow of length λ along this sequence step-by-step, the number of +1 and −1 symbols inside the window is counted at each step. This number γk(λ) is a new random variable that depends on the position k of the window along the sequence. The variable γk(λ) has a certain distribution. Its average is deter-mined by the overall sequence composition, and its variance is given by D2 ¼ ⟨γ2ðλÞ⟩ − ⟨γðλÞ⟩2, where ⟨…⟩ represents the λ average over all windows of size λ and generated sequences.

If the sequence is uncorrelated (normal random walk) or there are only local correlations extending up to a characteristic range (Markov chain), then the value of Dλ scales as λ½ with a window of a sufficiently large λ. A power law Dλ ∝ λα with α >½ will then manifest the existence of LRC. At this point, we note that LRC is not equivalent to ‘random-blocky’ sequences. In the case of the latter, as we will see, the exponent α is always equal to ½ if the sequence length is large enough.

In some cases, because of large fluctuations, conventional scaling analysis cannot be applied reliably to the relatively short sequences characteristic of natural/synthetic polymers or generated in typical simulations. To avoid this problem, one can use the so-called detrended fluctuation analysis method, which was specifically adapted to handle problems associated with short nonstationary sequences.57,58 This approach leads to a special function FD(λ) that characterizes the detrended local fluctuations within a window of length λ. Generally, FD(λ) shows the same behavior as Dλ. There are many other methods for monitoring LRCs in copolymers.59–61

6.18.2.2.3 Information complexity In a sense, the presence of LRC in a given copolymer sequence characterizes the ‘information complexity’ of the sequence. It is natural to expect that the content of information in the sequences of biopolymers (i.e., proteins, DNA, RNA) is relatively high in comparison with random sequences where it should be almost zero.62 If we then analyze the primary structures of the globular proteins and compare them with the simple primary structure of conventional synthetic copolymers, we should draw the perceptive conclusion that ‘protein-originated texts’ are much more informative and specific.

What is complexity? There is no good general definition of complexity, though there are many attempts to define it.63,64

Intuitively, complexity lies somewhere between order and dis-order or between regularity and randomness. Complexity has been measured by logical depth, metric entropy, information content (Shannon’s entropy and related characteristics65), fluc-tuation complexity, and many other techniques. These measures are well suited to specific physical or chemical appli-cations, but none describe the general features of complexity.

A simple measure capable of indicating how far a copoly-mer sequence differs from trivial random sequences is based on the so-called Jensen–Shannon (J S ) divergence measure.66–68

Let us explain how it can be defined. Let S ={s1,…,sN} be a sequence of N symbols. For two

subsequences S1 = {s1, …, sn} and S2 = {sn+1, …, sN} of lengths n and N – n, the difference between the corresponding discrete probability distributions f1(s1, …, sn) and f2(sn+1, …, sN) is quantified by the Jensen–Shannon divergence

n N−n J S ðS1,S2Þ=N ¼ hðSÞ− hðS1Þ þ hðS2Þ ½1� N N

where S = S1⊕S2 (concatenation) and h(S) is Shannon’s entropy65 of the empirical probability distribution obtained from block frequencies in the corresponding subsequence. Of course, Shannon’s entropy depends on the definition of a set of words in the sequence. For two-letter HP copolymers, one can adopt the following set of words (uniform blocks): H, HH, HHH, …, P, PP, PPP,…; that is, word (block) is defined by its length ℓ and type. In this case, Shannon’s entropy per monomer can be written as68

X h ¼ −

Nw fHðℓÞlog2 fHðℓÞ þ fPðℓÞlog2 fPðℓÞ ½2� 2N

l

where fH(ℓ) and fP(ℓ) are the frequencies of words of length ℓ composed of letters H and P, respectively, and Nw is the total number of words.

The Jensen–Shannon divergence J S is zero for subse-quences with the same statistical characteristics; it takes higher values for increasing differences between the statistical patterns in the subsequences, and reaches its maximum value for a certain set of distributions. In particular, both purely random and any regular (multiblock) copolymers of infinite length show J S ¼ 0. We normally expect that a completely random sequence or a sequence with long uniform blocks contains less information than a sequence containing many different blocks (words) of medium length. Of course, we cannot unambigu-ously distinguish only by sequence analysis between what might be called quantity and quality of information.

It is clear that information complexity and LRCs cannot emerge just as a result of random permutations of monomeric units in a copolymer chain. Some coupling of chemical mod-ifications to other factors is necessary. It is most straightforward to introduce the interrelation of chemical modifications and chain conformations. Such a consideration leads naturally to a concept of conformation-dependent sequence design (CDSD) in synthetic polymer chemistry.

6.18.2.3 Conformation-Dependent Sequence Design: Structure Dictates Sequence

In a series of publications,69–71 a concept of CDSD of functional copolymers has been introduced (for reviews, see

Page 6: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 693

Reference 72–77). The essence of the approach is based on the assumption that a copolymer obtained under some prepara-tion conditions is able to ‘remember’ features of its original (parent) conformation in which it was built up and to store the corresponding information in the resultant chemical sequence. In other words, this concept takes into account a strong cou-pling between the conformation and primary structure of copolymers during their synthesis.

Ideologically, the approach69–71 bears some similarities with that proposed earlier in the context of the problems of protein physics;42,46–50 however, it aims at synthetic copoly-mers rather than biopolymers. The original idea for protein design consists of running through sequences of amino acids to determine which sequence (or sequences) has the lowest energy in a unique (target) conformation. We have stressed the differences between sequence design characteristic of nat-ural and synthetic copolymers.

Presently, there are three known CDSD techniques: (1) the chemical modification (polymer-analogous transformation (PAT)) of homopolymers; (2) the step-growth copolymeriza-tion of monomers with different properties under special conditions; (3) the evolutionary and theoretically assisted sequence design approaches. We will address all these techniques.

6.18.2.3.1 Polymer-analogous transformations Controlling the chemical composition and architecture of polymeric materials is important in many technological appli-cations. While in most instances the resultant polymer can be prepared conveniently using existing synthetic methods, in other situations generating functional polymers can be challenging due to various reasons. These include the difficulty of finding a suitable polymerization media, the inability of existing synthetic methods to polymerize certain monomers, and complications arising from uncontrolled chemical side-reactions on a given monomer. These barriers can be over-come to some extent by implementing postpolymerization protocols (PPPs), also referred to as PATs, where a homopoly-mer is modified chemically with a selective reagent. Modification of polymers by PAT has been known for a long time (see, e.g., the Nobel Prize lecture of Hermann Staudinger78) and presently is well investigated. This synthetic strategy can be exemplified, for instance, by the bromination of polystyrene (PS).79 Polymer-analogous reactions represent deliberate changes of functional groups that are carried in macromolecular chains with the general objective of maintain-ing the polymerization degree of the original macromolecules.

Recent years have witnessed increased interest in utilizing PAT to create specialty polymers. In fact, some industrial pro-cesses, such as modification of polyolefins,80 employ PATs routinely. Progress in this field has recently been documented in a comprehensive review by Gauthier et al.81 While PAT can be used to convert a polymer made of a given type of segments, say A type, to a macromolecule made of chemically different units, say B type, by converting A into B, there exists a vast class of ‘in between’ materials that involve A1−xBx copolymers, where x is a fraction of the A units that have been converted chemi-cally into B moieties. Thus, by adjusting the degree of chemical modification of the bare A-homopolymer, one can prepare heteropolymers consisting of various degrees of chemical mod-ification along the polymer backbone. In most instances, the

distribution of the newly added B units in resulting chains is random, thereby leading to A1−xBx RCPs. However, by fine-tuning the condition of the PAT chemical reaction, one can also alter the degree of randomness of the A and B species. In particular, this can be done using a synthetic strategy based on the CDSD approach.

6.18.2.3.1(i) Protein-like copolymers It is known that the hydrophilic P units in globular proteins are mainly covering the surface of the globule, giving rise to their stability against intermolecular aggregation, while hydrophobic H units are forming primarily the core of the globule.82 Such requirement is rather restrictive, that is, it is satisfied only for a very small fraction of all possible primary structures. Moreover, since the HP correlations defined in such a way depend on the conformation of the globule as a whole (i.e., on the protein ternary structure), they should be characterized as long-range ones. From a practical viewpoint, the question is whether such primary structures can be obtained for binary copolymers, not obligatorily of biological origin.

In the late 1990s, Khokhlov and Khalatur69,70 devised a clever PAT-based methodology for adjusting the spatial distri-bution of comonomer species in two-letter RCPs. It is easy to do this by computer simulation, and much more difficult in real experiments. In both cases, however, the corresponding procedure should involve the four stages that are schematically depicted in Figure 2.

Stage 1: We take a bare homopolymer chain with excluded volume interactions in a good solvent (see Figure 2(a)).

Stage 2: Strong attraction between all monomeric units is switched on and a homopolymer globule (‘parent’ confor-mation) is formed (Figure 2(b)). Of course, when we are speaking about real experiments, by switching on the attrac-tion we should understand the jump of temperature, addition of poor solvent, and so on.

Stage 3: This step is much more easily realizable in computer experiments. We should simply consider the ‘instant snap-shot’ of the globule and ‘color’ the units on the surface and in the core in different ‘colors’, that is, we assign the index P to those units that are on the surface of the globule and call these units hydrophilic and assign the index H to the units in the core of the globule and call these units hydrophobic. By doing so, we fix this primary structure (Figure 2(c)). In real experiments, the ‘coloring’ of the surface can typically be done with a chemical reagent Z entering in the reaction with monomeric units and converting them from hydropho-bic to hydrophilic, H + Z → P. In chemical language, it is called PAT. If the amount of the reagent is small enough, only surface monomeric units will be contaminated, the core remaining hydrophobic. Another important feature is to have fast enough ‘coloring reaction’ and slow enough intermolecular aggregation (which will always take place under the conditions when globules are formed). To slow down the aggregation, the neutral thickeners of the aqueous solutions may be used.

Page 7: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

(a) (b)Stage 1

Stage 2

Homopolymer coil Homopolymer globule

Stage 3

(d) (c)

Stage 4

Protein-like copolymer ‘Coloring’ procedure

694 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 2 Schematic representation of the CDSD chemical modification of a homopolymer chain: (a) initial homopolymer coil; (b) initial (‘parent’) conformation formed by a homopolymer globule; (c) chemical modification (surface ‘coloring’); (d) resultant protein-like copolymer. Modified chain segments are shown in green.

Stage 4: This last step is necessary for computer realization. The uniform strong attraction of units should be switched off, and different interaction potentials should be introduced for H and P units (Figure 2(d)). As a result, we obtain a specific copolymer called ‘protein-like copolymer’ (PLC).69

Initially, the protein-like HP sequences were generated in References 69 and 70 for a lattice chain, using a Monte Carlo method and the bond-fluctuation model.83 When the chain is a random heteropolymer, an average over many different sequences distribution must be carried out explicitly to pro-duce the final properties. Therefore, the sequence design scheme69,70 was repeated many times, and the results were averaged over different initial configurations.

It turned out that many statistical properties of protein-like and RCPs with the same HP composition are very different. In order to distinguish whether this difference is due to the special sequence design described above, or just due to the different degree of blockiness, one can introduce for compar-ison also a ‘random-blocky’ primary sequence. We assume that random-blocky HP copolymers have the same chemical com-position and the same average length L of uninterrupted H or P sequence as PLCs, but in other respects, the HP sequence is random. In Reference 69, the distribution of block length ℓ was taken in the Poisson form: f( −L ℓ

ℓ) = e L /ℓ!. In Figure 3, we present the typical distributions of H and P

monomers along the chain for regular (multiblock) copolymer as well as for truly random copolymers, random-blocky copo-lymers, and PLCs. From the comparison of the primary structures of the random and PLCs, we can see that the average lengths of H and P blocks in the PLC are notably longer. In contrast, the copolymer with the random-blocky architecture, having the same average length of H and P sections, LH and LP, as for the PLC, exhibits a different distribution of these blocks

along the chain. The main feature of the protein-like sequences is the presence of rather long uniform H and P sections.

In Figure 3, we also show the variance Dλ as a function of λ for protein-like sequences with a 1:1 HP composition. For comparison, the data for two other types of sequences are also presented. One of them is a purely random 1:1 sequence; it demonstrates Dλ ∝ λ1/2 scaling, as expected. Comparing this curve with simulation results,84 we see immediately that the protein-like sequence is not random; rather well-pronounced correlations exist in it. Thus, it is interesting to compare the simulation data with the Poisson distribution adjusted to achieve the same 1:1 composition and the same degree of blockiness as for the corresponding PLC. This model sequence exhibits a somewhat more rapid variation of Dλ at small λ, but ultimately the law Dλ ∝ λ1/2 is obeyed for large values of λ. Nevertheless, this random-blocky model is also seen to be unsatisfactory for the statistical behavior of a protein-like sequence throughout the interval of λ examined, 2 < λ < 512. Although the simulation data do not fit accurately to any power law Dλ ∝ λα, the slope of the observed Dλ dependence corre-sponds to a significantly larger value than 1/2, up to about α = 0.85, thus indicating pronounced LRCs in a protein-like sequence. Thus, the primary structure emerging in the case of PLCs can be characterized as ‘quasi-random with long-range correlations’. Here, we stress again that there is the fundamen-tal difference between the sequences with LRC and random-blocky sequences.

The findings discussed above are surprisingly similar to those known for DNA sequences, which appeared as a mosaic of coding and noncoding patches.56–58 Indeed, like DNA chains, the PLC sequence also contains two types of alternating sections forming a certain pattern. It is known that the noncod-ing regions in DNA do not interrupt the correlation between the coding regions (and vice versa), and the DNA chain is fully

Page 8: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

Regular1

1.6−1

Protein-like Random 1 1.2 Random

Random-block

D λ 0.8

−1 10

slope 0.5Random block Log

0.4 1

0.0

−1 −0.4

Protein-like 0.0 0.5 1.0 1.5 2.0 1

Log10λ

−1 200 400 600 800 1000

Monomer unit

� � � �

� �

Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 695

Figure 3 Left panel: Typical distributions of H and P monomeric units along the chain for regular (multiblock) sequence having fixed block length of 8 units, as well as for purely random sequence with average block length L = 2, random-blocky sequence with L = 6.4, and protein-like sequence with L = 6.4. The H units are denoted as +1 and the P units as −1. Right panel: Variance of the number of different monomeric units in the fragment of sequence of size λ for protein-like, random, and random-blocky copolymers. The sequence length N = 1024.

correlated throughout its whole length. As a result, the D2 (orλ FD2 ) curve does not contain the linear portion D2

λ ∝λ. In princi-ple, the same behavior is observed for protein-like sequences.77

Govorun et al.84 have shown, both by exact analytical theory and by computer simulations, that the corresponding chemical sequence is nonalternating and demonstrates the specific LRC, which can be described by the statistics of the Lévy-flight type.85 For such probabilistic processes, an observable stochas-tic variable x exhibits large jumps (flights), called ‘Lévy flights’ (LFs), characterized by power-law (rather than exponential) probability distribution function f(x). The possibility for exact analytical description of sequences resulting from surface ‘col-oring’ (Figure 2(c)), comes from the fact that the statistics of polymer chains inside dense globule is Gaussian, that is, it is described by the ordinary diffusion equation. One has only to worry about correct boundary conditions, and this problem was resolved in Reference 84. The probability distribution of block lengths ℓ in PLC can be written as84

∞ � X b2 2πb nπb nπ fHðℓÞ ¼ nð−1Þnþ1sin exp − ℓ ½3�

3R� R� 6 R� n¼1

for H units and

b=R�−1∞ ζ 2 X sin ζ n1 n 21=3−1

fPðℓÞ ¼ 23ð21=3−1Þ n¼1 ζ n− sin ζ n cos ζ n " # b2 2ζ n exp − ℓ ½4� 6 R�ð21=3−1Þ

for P units. Here, b is the bond length, R* = b(3N/8π)1/3 defines the characteristic size of the bare homopolymer globule, and ζn

satisfies ζn = (1–2−1/3)tan ζn.

Analytical theory84 suggests that the LF-type statistics, albeit with a broader crossover region, is expected even if parental

(globular) conformation used to generate protein-like macro-molecules is not maximally compact, but rather a globule somewhat closer to the Θ-point. Note that, strictly speaking, this distribution is not a ‘true’ LF distribution, which falls to zero as 1/|x|α+1 (where 0 < α < 2) and, therefore, has an infinite variance, but represents the so-called truncated Lévy flight

84,85(TLF) with a finite variance for finite chain lengths N. As has been shown,84 the copolymers generated with the

use of eqns [3] and [4] reproduce the coloring procedure described in References 69–71: monomers located in the core of globule are set to be H-type, while monomers belonging to a globular surface are assigned to be of P-type. An analogous result was obtained in a later work86 for the model that takes into account more chemical details.

The presence of LRCs in designed sequences is a very impor-tant feature. It is easy to understand that these correlations are due to the fact that assigning of the type of chain segments (H or P) under the preparation conditions described above depends on the conformation of the parent globule as a whole, not on the conformation of small sections of the initial homopolymer chain. From this viewpoint, one may say that such a sequence encodes in two-letter alphabet the spatial (core–shell) structure of a copolymer globule. Obviously, this functional feature can be realized if and only if a general statistical pattern is attributed to the sequence as a whole, and cannot be obtained by joining of independent statistical pat-terns of two or more subsequences of smaller lengths.

Because of the finite size of the bare globule subjected to CDSD, the longest correlations that can be found in this case are also finite. However, the range of these correlations is much larger than for usual synthetic copolymers obtained, for example, via radical copolymerization under homogeneous conditions.31

The question of whether natural proteins originate from random sequences of amino acids was addressed in many

Page 9: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy696 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

(a)

Hydrophilic loop Hydrophobic core

(c) (b)

Figure 4 Typical snapshots of the globular conformations of (a) protein-like, (b) random, and (c) random-block copolymers of the same length (N = 512). Hydrophobic segments are shown in light-red color and hydrophilic segments in green color.

publications. It was demonstrated that protein sequences are not completely random sequences.87 In particular, the statisti-cal distribution of hydrophobic residues along chains of functional proteins is nonrandom.88 Furthermore, protein sequences derived from corresponding complete genomes dis-play a distinct multifractal behavior characterized by the so-called generalized Rényi dimensions (instead of a single fractal dimension as in the case of self-similar processes).89

It was also found that the coil-to-globule transition for PLCs, induced by the strong attraction of H segments, occurs at higher temperatures and leads to the formation of denser globules and has faster kinetics than for random and random-block counterparts.69–71,77 The reason for this is illu-strated in Figure 4 where typical snapshots of globules formed by protein-like and RCPs are shown.

One can see that the HP heteropolymer obtained as a result of the simple one-step ‘coloring procedure’ can self-assemble into a segregated core–shell microstructure, thus resembling some of the basic properties of globular proteins. The core of protein-like globule is much more compact and better formed as compared to that observed for RCPs; it is surrounded by the loops of hydrophilic segments, which stabilize the core.

Apparently, this is due to some ‘memory effect’: the core, which existed in the parent globular conformation (Figure 4(a)), was simply reproduced upon refolding caused by the attraction between H units. One may say that the fea-tures of parent conformation are inherited by the PLC. Looking at the conformations of Figure 4, it is natural to argue that PLC globules could be soluble in water and thus are open to further chemical modification, while RCP globules will most probably precipitate. It is clear that the main reason for the deviation of the properties of PLCs from those for RCPs is the special sequence design scheme,69–71 not just difference in the degree of blockiness.

6.18.2.3.1(ii) Statistically controlled monomer sequence distributions A drawback of the simplest sequence design scheme described above is that either a copolymer sequence is templated or it is random, with little control in between. In their initial

work, Khokhlov and Khalatur69–71 assumed that the polymer-analogous reaction was instantaneous, that is, the parent homopolymer before chemical modification was ‘frozen’. It is easy to understand that instantaneously modifica-tion of the parent homopolymer is almost impossible to realize experimentally since any reaction between the homopolymer units and the reactant species is kinetically limited and will take place on a much longer time scale than the monomer and coil relaxation times.

Semler and Genzer90 extended the original method of Khokhlov and Khalatur by allowing the copolymer to reconfi-gure during the chemical modification process, based on the reasonable expectation that the copolymer’s conformation would evolve during this process. To this end, they used Monte Carlo simulations to model the formation of PLCs with tunable monomer sequence distributions.

The simulations were carried out in a two-step computa-tional scheme. First, using a Monte Carlo method and a lattice chain model83 with an implicit solvent, the collapse transition was mapped out via varying the reduced monomer–monomer interaction potential ε/kBT. This procedure provided the initial homopolymer configurations for the subsequent ‘coloring’ simulation. The degree, to which the homopolymer coil is collapsed, was achieved by adjusting the inter- and intramole-cular potentials acting among the chain monomers. In the second process, another Monte Carlo simulation was con-ducted, which involved the PAT performed directly on the homopolymer chain of a given state of collapse. This simula-tion scheme used an explicit low-molecular-weight reactant that had a predetermined reaction probability with the mono-meric units along the homopolymer. A threefold excess of reactant was added at random to the simulation box. During this simulation, a monomer unit along the length of the chain had a certain probability to either react with the reactant or make a conformational move. Reactions were allowed to occur with a predetermined probability (set to either 5% or 25%) if the distance between the reactant and monomer along the chain was less than a predetermined reaction radius. The simu-lation progressed until a specified percentage of the chain was colored (two degrees of coloring was considered: 40% and 60%). Both the chain collapse and coloring simulations were performed many times in order to access the statistics of the process.

The simulation results revealed that regardless of the length of the parent homopolymer, all monomers along the fully expanded chains possessed an equal probability of coloring. In contrast, the probability of coloring monomer units in col-lapsed chains was distributed widely. Increasing the chain length resulted in monomer sequences with broader distribu-tions of colored units. Furthermore, expanded coils exhibited an upsurge in the frequency of occurrence of short sequence lengths of the colored units, while collapsed coils contained larger fractions of long sequential monomer lengths. A higher frequency of long-colored blocks was also detected upon increasing the chain length of the parent homopolymer. A simple statistical analysis revealed that while the sequence distribution of the colored species in polymers colored in an expanded state was random, chains colored under collapsed conformations transformed into heteropolymers, which pos-sessed statistically self-similar sequence distributions. The latter sequences can be identified as protein-like ones.

Page 10: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 697

The fact that distinct coloring sequences can be achieved by coloring parent homopolymers over time (rather than perform-ing instantaneous coloring as performed in the original sequence design scheme of Khokhlov and Khalatur69–71) has important implications for preparation of RCPs with distinct and adjustable monomer sequences.

6.18.2.3.1(iii) Hydrophobic modification of hydrophilic polymers The segregated core–shell microstructures, consisting of a hydrophobic core surrounded by a hydrophilic shell, are of great practical interest as their mechanical properties are mainly influenced by the core polymer and the chemical properties and solubility mainly by the shell monomer units. These mate-rials have attracted increased attention because they not only maintain the functions of both the core and shell components but also exhibit additional excellent optical, electrical, and magnetic properties.91 Polymeric core–shell microstructures can be obtained through phase separation,92 which is realized by inductive polymerization,93 solvent extraction and evapora-tion,94 self-assembly of amphiphilic block copolymers,95 or sequential precipitation,96 which is intrinsically a self-assembly and phase-separation process. However, these methodologies are usually complicated and tedious.

A simple CDSD-based approach to synthesis of copolymers with a specific sequence capable of forming core–shell micro-structures has been suggested.76 The idea is rather similar to that discussed in the previous sections but bears an important difference: instead of the hydrophilization of globular surface, it is suggested to perform a sequential hydrophobization of a hydrophilic polymer chain (poly-P) dissolved in a good sol-vent, using a low-molecular-weight hydrophobic modifier (H) poorly soluble in this solvent. The approach is illustrated sche-matically in Figure 5.

In a dilute solution, when the polymer is in a coiled state (Figure 5(a)), the diffusion of hydrophobic particles into the coil is normally faster than the chemical reaction.97 In this case, the local concentration of particles H inside the swollen coil is practically the same as in the bulk. Therefore, we expect that at

(a) (b)

(d) (c)

Figure 5 Schematic representation of the hydrophobic modification of a hydrophilic polymer in a solvent, which is selectively poor for hydrophobic modifier and modified chain segments. Modifying agent and hydrophobic monomers are shown in light-red.

the initial stage, the reaction will lead to a RCP: some of the P monomeric units will attach H reagent and thereby they will acquire amphiphilic (A) properties: P + H → A (Figure 5(b)). As long as the number of the modified A units is not too large, the chain remains in a swollen coil-like conformation (Figure 5 (b)). However, when this number becomes sufficiently large, the hydrophobically modified polymer segments tend to form intrachain micelle-like aggregates (Figure 5(c)). This is due to the loss in translational entropy of covalently bonded H species after their grafting to the hydrophilic backbone. Structurally, the intramolecular micelles are similar to reverse micelles formed by free low-molecular-weight surfactants, and like ordinary micelles they should solubilize hydrophobic species. The presence of the intrachain aggregates can dramatically alter the reaction conditions. Because of preferential adsorption, the poorly soluble modifier will diffuse inside the A-rich regions, thus leading to spatially inhomogeneous distribution of the concentrations (Figure 5(c)). One can expect that further mod-ification of the hydrophilic sections of the polymer chain will occur predominantly within these microglobular regions. In the course of chemical reaction, they would progressively increase in size and then coalesce. As a result, a nonrandom microblock distribution of the chemically modified chain seg-ments emerges. Finally, we should have hydrophobically modified segments inside rather compact conformation formed by the resultant hydrophilic–amphiphilic (PA) copoly-mer. In that way, one can expect the formation of core–shell morphologies with inner (poorly soluble) core and outer (well-soluble) hydrophilic cover (Figure 5(d)). In principle, the polymer-analogous reaction can be terminated at any desir-able time by quenching (e.g., by lowering the temperature) or by stopping the supply of reactive compounds.

The algorithm used in Reference 98 simulated a reaction in a three-dimensional cube, utilizing periodic boundary condi-tions. Initially, an N-unit hydrophilic homopolymer (poly-P) in a coil state and N freely diffusing hydrophobic (H) mono-mers were placed in the cube. For the conditions considered, despite the attraction between H monomers, they were soluble due to their high translational entropy. To treat the hydropho-bic interactions, an approach based on the solvation free energy (SASA) model99 was used. Upon contact between a monofunc-tional reactant H particle and a P bead on the chain, a bond could be formed between the two. This led to the formation of an ‘amphiphilic’ monomer unit (PH amphiphilic dumbbell): P+H → A. The reaction was considered as a sequence of the alternating steps: grafting of a new H monomer to the chain and the subsequent long relaxation of the chain. The process was terminated when the required number of the hydrophilic monomer units was transformed into the A type. The composi-tion of the resultant copolymer was constrained so that there were 75% hydrophilic and 25% amphiphilic monomer units. As a result, a hydrophilic–amphiphilic PA copolymer with a certain distribution of P and A units along the hydrophilic polymer backbone was obtained. To gain better statistics, �103 hydrophobically modified copolymer chains for each set of the parameters were independently generated and then the required average characteristics were found.

To give a visual impression of the simulated system, Figure 6 presents a typical snapshot of an amphiphilic copoly-mer having 256-unit hydrophilic backbone with 64 attached hydrophobic side groups. It is seen that, using the synthetic

Page 11: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

100

γ = 0

10−1 γ = 0.3γ = 0.6

) l 10−2

(Af

10−3

10−4

4 8 12 16 20 i /N

698 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 6 Left panel: Typical snapshot of a hydrophobically-modified copolymer. Hydrophilic chain segments are shown in green color and hydrophobic side groups in red color. Right panel: Average number fraction of blocks consisting of ℓ neighboring amphiphilic monomers occurring in a 256-unit copolymer chain hydrophobically modified in a solvent having different selectivity for modifying agent. In this model, the solvation free energy, Δμs, required to transfer an N-site molecule from vacuum to polar solvent (e.g., in water) is approximated by the sum of linear terms γiSi, where Si is the solvent-accessible surface area (SASA) for site i and γi is the solvation parameter for the site.99 In this approximation, Δμs can also be considered as the free energy cost for the transfer of a particle from the interior of a hydrophobic cluster to the solvent. The solvation parameter γ is an estimate of the free energy of transfer divided by exposed surface area. For hydrophilic particles, γ ≤ 0. For hydrophobic particles, solvent quality becomes poorer with increasing γ. In reality, when the temperature is fixed, the change in γ can be due to variation of the solvent composition.

strategy described above, one can indeed end up with a copo-lymer having a dense hydrophobic core surrounded by a hydrophilic shell.

From the picture presented in Figure 6, one can expect that the sequential hydrophobization of a polymer coil should lead to a copolymer with a nonrandom sequence distribution. This is indeed the case. As an example, let us consider the average number fractions of blocks consisting of ℓ neighboring amphi-philic monomers, fA(ℓ), occurring in a copolymer chain. Some results are shown in Figure 6 on a semilogarithmic scale.

We expect that for a random distribution of monomers A incorporated into the polymer chain, the function fA(ℓ) should decay exponentially with increasing ℓ. In fact, such a behavior is observed for the copolymer modified in a solvent, which is good for the low-molecular-weight modifier, that is, at γ = 0. When the solvation parameter γ is increased and the solvent becomes selectively poorer for the modifying agent, the values of fA(ℓ) are skewed toward A sections of greater length, which implies a copolymer with blocky tendencies.55

A further worsening of the solvent quality leads to a strong deviation from the exponential decay of fA(ℓ), thus indicating a nonrandom copolymer sequence distribution.

Thus, copolymers of the same composition can have quali-tatively different sequence distributions depending on the solvent in which the chemical transformation is performed. In a solvent selectively poor for modifying agent, hydrophobi-cally modified copolymers were found to have the sequence distribution with LRCs, whereas in a nonselective (good) sol-vent, the reaction always leads to the formation of random (Bernoullian) copolymers. In the former case, the chemical microstructure cannot be described by any Markov process, contrary to the majority of conventional synthetic copolymers.55

In general, the microsegregated structures observed for hydrophobically modified polymers (Figure 6) are similar to core–shell globules obtained via ‘coloring’ procedure (Figure 2). However, it should be kept in mind that the

experimental realization of the hydrophilic modification of the surface of a hydrophobic globule was shown to be rather unreliable, because of the difficulty to stabilize dense globules in the solution for the time sufficient to implement a PAT.74 On the other hand, the simulations76 show that the method based on the hydrophobic modification of soluble polymers should be quite universal and robust.

6.18.2.3.1(iv) Design of PLCs with tunable randomness Discussing the further development of the sequence design scheme based on PAT involving macromolecular substances modified by a low-molecular-weight reactant, it is of interest to consider how differences in the solubility of a homopolymer and modified monomeric units affect the sequence of the resulting copolymer. One would expect that as the modifica-tion reaction evolves, solubility of the newly created segments along the designed copolymer affects the chain conformation and ultimately the distribution of the chemically different units inside the designed copolymer. It is believed that while color-ing the A homopolymer with B species that are more soluble in the solvent should result in opening up the chain, B monomers that are less soluble than the A units would move toward the center of the collapsed polymer. The situation when the mod-ified monomer unit exhibits lower solubility in the solvent than the original unmodified monomer appears to have been the case in experimental work by Semler et al.,100 who carried out a polymer-analogous reaction by brominating parent PS coils in selective solvents.

Recently, Strickland et al.101 have used computer simula-tions to examine situations in which the modified monomeric unit exhibited lower, identical, or greater solubility when com-pared with the original unmodified monomer. They presented the results of discontinuous molecular dynamics (also referred to as event-driven molecular dynamics) simulations of polymer-analogous reactions with special focus on the case in which the chemically modified monomers had different solu-bility than the parent homopolymer. A system of polymer

Page 12: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

A0.90B0.10 A0.80B0.20 A0.70B0.30 A0.60B0.40 A0.50B0.50

25 degree of polymerization (N) 0.

5

100 =

150 BA

20 200 R

250 300

15 1.0

> =

2 g

BA

<R

R

10

5.0

5 =

BA

R

0 0.0 0.1 0.2 0.3 0.4 0.5 0.6

10.0

Degree of coloring =

BA

R

Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 699

molecules, modeled as chains containing 100–300 square-well monomers of A type, was allowed to equilibrate at a selected temperature. After equilibration, athermal reactant particles were placed in the simulation box. When the reactant particles came in contact with A-type monomers, the monomer was modified to become a B-type monomer. This reaction was continued until a desired number of the parent monomers was modified. Since the solvent in the simulations was implicit, the solubility of A- and B-type monomers was adjusted by varying the A–A and B–B interaction strengths. The ratio of B–B inter-action strength to A–A interaction strength, RBA = |εBB|/|εAA|, was varied from 0.5 to 10, thus allowing to vary the solubilities of A and B species. Interactions between non-bonded A and B units along the chain were modeled as athermal, that is, εAB =0.

The simulations showed that the sequence distribution of A and B species in the resultant copolymer is affected by the interaction strengths acting between the A–A and B–B mono-mers, that is, their solubilities. In general, when the solubility of the B monomers is not the same as the A monomers (εAA ≠ εBB), the chain conformation changes during the coloring reaction by either expanding (when B is more soluble than A) or contracting (when B is less soluble than A). It was found that there exists a minimum degree of modification before substan-tial chain reconfiguration can take place; typically, reconfiguration requires a minimum of 20% conversion of A to B. For high values of RBA (i.e., very low B solubility), the B monomers tend to move toward the center of the globule, causing the globule to invert. Strong A–A interaction and good B solubility led to the formation of tight globules and hence produced copolymers with increased tendency toward random-blocky comonomer distribution. In contrast, weak A–A interaction and good B solubility led to extended homo-polymer conformations, which produced copolymers with random comonomer distribution.

More interesting and somewhat unexpected behavior occurred when the solubility of B was poor compared to that of A. One might expect that in cases of poor B solubility, the

copolymer would contract in size during the coloring process. Instead, as can be seen by examining the data in Figure 7 (left panel), under poor solvent condition, the globule size increases with increased degree of modification, an effect which, at first glance, is counterintuitive. This initial swelling of the copolymer is explained by the fact that B monomers move into the globule away from the solvent. Thus, as the poor-solvent homopolymer is modified, it begins to expand as if it were in good solvent due to the solubility of B. In essence, the good solubility of B reduces the effective theta temperature of the system as the chain is being modified. One can also observe the opposite trend when the solubility of B species is poor. In this case, the initially swollen chain trans-forms from coil-to-globule as the chain is modified. In other words, the poor solubility of B species increases the effective theta temperature of the system as the chain is being modified. This effect is similar to that discussed for a hydrophobically modified copolymer in the previous section. Some typical con-formations of the resultant copolymers are shown in Figure 7 (right panel).

The results from the simulations are summarized in Table 1, which shows that the changes in chain conformation during the PAT can be divided into four major categories, depending on the initial conformation of the parent homopolymer (globule or coil) and the relative solubility of B (more or less soluble).

Statistical analysis of the generated copolymer sequences revealed that one can effectively control the degree of random-ness by simply varying the value of kBT/|εAA|.

6.18.2.3.1(v) Chemical modification of surface-tethered homopolymers Very recently, Strickland et al.102 developed a sequence design methodology based on chemical modification of homopoly-mer macromolecules grafted to flat impenetrable surfaces. As discussed above, previous studies employed PAT by placing the parent homopolymers in a selective solvent and subse-quently letting the chemical reagent modify the units on the

Figure 7 Left panel: Mean-square radius of gyration as a function of the degree of coloring (i.e., the mole fraction of B units, x, in A1−xBx) for polymers of chain length ranging from 100 to 300 at RBA (=|εBB|/|εAA|) = 10.0 and kBT/|εAA| = 0.2 (poor solvent). Right panel: Conformation of a 300-mer A1−xBx prepared by ‘coloring’ a parent A 300-mer (gray balls) with the coloring species (red balls) to various degrees of coloring, x, ranging from 0 to 0.5 at kBT/|εAA| = 0.2 and RBA = |εBB|/|εAA| = 0.5, 1.0, 5.0, and 10.0.

Page 13: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy700 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Table 1 Effect of A and B solubility on copolymer conformation during the coloring process

Parent conformation B monomer solubility Change in chain size upon increasing x

Coil (soluble) Coil (soluble) Globule (insoluble) Globule (insoluble)

Good Poor Good Poor

Minor increase or no change Slow contraction Rapid expansion at x ≈ 0.30, followed by slow expansion Rapid expansion at x ≈ 0.20, followed by slow collapse

homopolymer that were sterically accessible. Since the parent homopolymer adopted either a coil or globule conformation and, most importantly, was always isolated, the chemical modification occurred along the entire backbone of the macro-molecule. Selective shielding of specific portions of the chain was not possible. This limitation can be overcome by tethering the parent chains to a surface prior to coloring. Not only does the surface restrict the approach of the reactant to one end of the chain, but neighboring chains would also laterally shield each other depending upon the chain grafting density. While chemical modification of surface-tethered homopolymers has been performed experimentally by several groups,103–108 no quantitative information exists on the distribution of the mod-ified species in the resultant copolymers.

In their simulations, Strickland et al.102 employed discon-tinuous molecular dynamics to study the coloring process in macromolecular tethers made of various lengths. Specifically, they explored the effect of the grafting density of the parent polymer and the interplay between the solubilities of the unmodified and modified segments along the polymeric grafts on the comonomer distribution in designed copolymers.

The computer simulation proceeded in the following way. First, M homopolymer chains comprising various numbers N of square well monomers (50-, 100-, and 300-mers) were attached to a surface (the z = 0 plane) at various surface densi-ties, ρ = M/S, where S is the area of the tethering surface. Then the homopolymer grafts were equilibrated at a select tempera-ture and surface density. After equilibration, reactant particles (Z) were placed randomly in the simulation box with lateral (x,y) periodic boundary conditions. Upon contact between an A-monomer and an Z-particle, the A-monomer was ‘colored’ and converted from A into B species (A þZ→ B). The Z-particle was removed and then randomly placed back into the box to maintain a constant reactant concentration. All A–Z interac-tions were considered effective and any A→B conversion was accepted. This polymer-analogous reaction was carried out until a desired number of B-type monomers was achieved, producing A1−xBx designed copolymers, where x is the mole/ number fraction of the B segments. The solubility of A- and B-type monomers was adjusted by varying the system reduced temperature, kBT/|εAA|, and the B–B interaction strength, εBB. The ratio of the B–B interaction strength to the A–A interaction strength (RBA = |εBB|/|εAA|) was varied from 0.5 to 10. In this manner, the newly created B unit can be modeled as either more (|εBB| < |εAA|) or less (|εBB| > |εAA|) soluble than A. Interactions between nonbonded A and B units along the chain were modeled as athermal, that is, εAB = 0. Typically, the reactions were run until the chain was composed of at least 60% B units. The distribution of A and B units along the polymer chain was calculated to determine the effect of solu-bility, chain length, surface density, and temperature on the

comonomer distribution of the surface-tethered designed copolymers.

The simulations102 demonstrated that the monomer sequence distribution in the resultant copolymer depends on the interplay among (1) the length and the grafting density of the A-based homopolymer anchors, (2) the solubility of the parent homopolymer, and (3) the solubility of the B coloring units. In particular, the monolayer density profile, end-monomer probability profile, and the sequence distribu-tion of A and B species along the designed copolymer chain ware found to be functions of the parent homopolymer solu-bility and temperature, as measured by kBT/|εAA|, and modified monomer solubility, as measured by the RBA ratio. The blocki-ness of short copolymers formed at low surface density (especially below the mushroom-to-brush overlap density) is similar to that of copolymers formed in the bulk due to the low-level of chain–chain interactions. When chain length was increased to the point where neighboring chains overlaped, the blockiness increased with increasing surface density. In general, increasing the system temperature reduced copolymer blocki-ness. High temperature promoted the extension of the chains away from the tethering surfaces, allowing for reactions along the length of the polymers. In contrast, low temperature pro-moted the collapse of the chains, concentrating the reactions at exposed portions of the polymer. It was shown that the effect of the surface density on blockiness depended on the chain length. In general, for a given increase in surface density, the blockiness of a longer-chained system increased more than that of a shorter-chained system. Increasing surface density at tem-peratures below the theta-temperature (Tθ) had little effect on blockiness but had a large effect at temperatures above Tθ. In particular, at kBT/|εAA| ≫ Tθ and high surface density, the dis-tribution of the segments in the brush resembles closely that of diblock copolymers where the block close to the surface com-prises primarily unmodified segments A and the block at the periphery of the tethered chain is made of a chemically mod-ified chain section.

The simulations carried out by Genzer and coworkers also revealed how the coloring develops over time as a function of the solvent quality for both A and B units, as well as the grafting density of the parent homopolymer on the flat surface. It was shown that grafting density plays a pivotal role in determining the comonomer distribution along the designed copolymers, which can range from truly random to diblock-copolymer-like.

It should be noted that the present simulation work may provide design guidelines that go beyond those pertaining to the formation of grafted designed copolymers. Specifically, one can envision situations where the coloring species are replaced with nonpolymeric objects, such as metallic or oxide nanoparticles, thus producing polymer-based composites grafted to surfaces. Such systems have been realized experimentally109–118 and

Page 14: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 701

studied theoretically.109 Results from those studies indeed reveal that the penetration depth of the nanoparticles inside polymeric grafts depends on the interplay between chain length, grafting density, and size of the nanoparticle.115,117–119

6.18.2.3.1(vi) Adsorption-tuned copolymers and molecular dispenser The idea of CDSD via PAT (postprocessing) can be generalized in many respects.77 Indeed, a special chemical sequence can be obtained not only from a globular conformation; any specific polymer chain conformation can play the role of a parent one.

The simplest example of this kind is connected to the con-formation of a homopolymer partly adsorbed onto a flat substrate (Figure 8, top panel). Let us assume that the chain segments being in direct contact with the surface in some typical instant conformation are chemically modified. This can take place when the surface catalyzes some chemical trans-formation of the adsorbed segments. One can expect that after desorption, such a copolymer will have special functional properties: it will be ‘tuned to adsorption’.

Following this line, Zheligovskaya et al.120 compared the adsorption properties of copolymers with special ‘adsorption-tuned’ primary structures (adsorption-tuned copolymers, ATCs) with those of truly RCPs and random-block copolymers. Monte Carlo simulations revealed that specific features of the ATC primary structure promoted the adsorption of ATC chains, in comparison with their random and random-block counter-parts under the same conditions. In other words, the resultant copolymer sequence ‘memorized’ the original state of the adsorbed homopolymer chain. Its statistical properties exhibit LRCs of the LF type similar to those found for PLCs obtained via PAT of a homopolymer globule.121

Zheligovskaya et al.120 have simulated the adsorption of ATCs. The critical adsorption energy and conformational prop-erties were compared with those of RCPs with the same content of adsorbed segments and random-block copolymers with the same composition and the same average numbers of adsorbed and nonadsorbed blocks. It was found that the difference in the primary structure of the chains leads to the difference in the critical adsorption energy ε* and the characteristics of adsorbed chains. In particular, the ATC chains have the smallest (by the absolute value) critical adsorption energy ε*. The RCPs are characterized by the largest ε*, as compared to other copoly-mers. This fact is simply explained by the difference in block lengths: the RCPs have the shortest blocks. This is consistent with the analytical results122 for regular binary copolymers, according to which the absolute value of the critical adsorption energy decreases with increasing block length at the same frac-tion of adsorbed and nonadsorbed segments. At the same time, the difference in the critical adsorption energy for the random-block and ATC chains (which are characterized by the same average block lengths) can be explained only by the details of the ATC primary structure. Because the ATC chains have significantly longer end nonadsorbed blocks, as compared to their random-block counterparts, the adsorbed segments are placed more compactly in each ATC chain. This specific feature of the ATC primary structure promotes adsorption of ATCs. Thus, the difference in the adsorption behavior of ATC, ran-dom, and random-block chains can be rationally explained by taking into account the specific features of the primary structure of these chains.

The obtained results support the general idea of a CDSD: the generated ATC sequence ‘memorizes’ some features of the specific parent conformation of the adsorbed homopolymer. In particular, the position of adsorbed segments is tuned in the

(a) (b) (c)

(a) (b) (c)

Figure 8 Top panel: Schematic representation of the sequence design procedure leading to an adsorption-tuned copolymer. (a) Initial partly adsorbed homopolymer, (b) chemical modification of adsorbed chain segments, (c) resultant copolymer. Modified segments are shown in red. Bottom panel: Stages of preparation of copolymer envelope. (a) Adsorption of homopolymer chain on a colloidal particle, (b) ‘coloring’ of the polymer chain (blue color corresponds to chemically modified monomer units and red color to adsorbed units) and introduction of cross-links (shown as green sticks) to stabilize hollow-spherical structure, and (c) elimination of the core particle.

Page 15: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy702 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

best way for subsequent adsorption. Therefore, it is not surpris-ing that this memorized hidden information became apparent as soon as we consider the adsorption of ATC chains. In other words, the polymer ‘learns to be adsorbed’ in the parent con-formation, and this ‘experience’ is used in the subsequent ‘life’ of this copolymer.

Velichko et al.123 suggested the model of a so-called ‘mole-cular dispenser’. This idea is a further development in the direction of CDSD. Namely, they considered the conformation of a homopolymer chain adsorbed on a spherical colloidal particle (Figure 8, bottom panel) and performed design of sequence for this state of macromolecule. The motivation behind this design procedure is that if we eliminate the parent colloidal particle after the design is completed (e.g., by etch-ing), the resultant copolymer will be tuned to adsorb selectively another colloidal particle of a parental size σp. For instance, if such copolymer is exposed to a polydisperse colloidal solution of particles of different size, it will selectively choose to form a complex with the particle having the same radius as that in parental conditions. That is why such a macromolecular object can be called a molecular dispenser.

It should be noted that the development of such polymer systems is stimulated by existing experimental works. In particular, the experimental methods of preparation of nanometer-sized hollow sphere structures have been sug-gested124–129 because of their possible usage for encapsulation of molecules or colloidal particles. The prepara-tion of hollow sphere structures, generally, is based on self-assembling properties of block copolymers in a selective solvent, that is, on the formation of polymer micelles with a nanometer-sized diameter. Further cross-linking of the shell of the micelle and photodegradation130 of the core part produce nanometer-sized hollow cross-linked micelles.

The sequence design procedure proposed in Reference 123 can be described in more detail as follows. First, we consider a homopolymer chain attracting to a colloidal nanoparticle. Such chain is forming an adsorbed complex with the particle (its typical conformation is shown in Figure 8(a), bottom panel). Only a fraction of chain segments is in direct contact with colloidal particle, while other segments form flower-like loops. Then we ‘color’ the segments in the loops in ‘blue’, while the segments near colloidal particle remain ‘red’, that is, they are attracting to this particle. If the sequence design is stopped at this stage, the pronounced selectivity of the complex forma-tion with another particle of parental size σp is not reached.123

However, if additional cross-links are introduced between adsorbed (red) units, thus fixing the cage structure of the cen-tral cavity (Figure 8(b), bottom panel), the macromolecule emerging after elimination of colloidal particle (Figure 8(c), bottom panel) does indeed exhibit the features of a molecular dispenser. This sequence design scheme was realized in the computer simulations.123

To characterize the complexes formed by molecular dispen-ser and colloidal particles, the probability P(σ,T) of finding a complex made of the copolymer envelope and the particle of a

123 Itgiven size, σ, was calculated as a function of temperature T. was found that the selectivity of the complex formation strongly depends on the number of cross-links in the envelope. The particles of parent or smaller size (σ ≤ σp) are fully absorbed by the central cavity, because the corresponding fitting was ensured by the sequence design procedure. In contrast, particle

of a larger size (σ > σp) is too big for a central cavity, and thus the complex formed does not saturate all the possibilities for the attraction of adsorption-active (red) units to the surface of the particle. As to small particles, they easily pene-trate inside the molecular dispenser, but the resultant complex is not stable (especially at high temperature) because of small surface of such particles. All these factors lead to a nonmono-tonous behavior of P(T,σ): there is the peak in P(T,σ) at σ ≈ σp

for moderately cross-linked copolymer envelopes.123

6.18.2.3.2 Copolymerization under heterogeneous conditions 6.18.2.3.2(i) How can we implement CDSD polymerization? Radical copolymerization at heterogeneous conditions offers additional opportunities not available in homogeneous (solu-tion) copolymerization.30,31 These include the intrinsic possibilities of exploiting the heterogeneities of the reaction system to control the chemical microstructure of the synthe-sized copolymers, making possible new paradigms for synthesis and production of polymeric materials. We discuss some new CDSD-based synthetic strategies, which have been developed in recent years to provide effective control of the chemical sequences.

The CDSD copolymerization differs in principles and in results from the conformation-dependent PAT.76,77 The corre-sponding process can be considered as a variant of template polymerization (also called molecular imprinting) based on the noncovalent binding of polymerizing monomers to a template.

The essence of this technique consists of the copolymeriza-tion of monomers differing in their affinity to the template and therefore different distribution in the reaction system. In con-trast to conventional types of template polymerization,131,132

in the CDSD copolymerization, all the monomers are bifunc-tional and thus form linear, not cross-linked, polymers. The sequence of the segments in the resultant molecularly imprinted copolymer is determined by the template-controlled conformation of the propagating macroradical.75–77 Thus, the CDSD copolymerization regime is possible only in reaction systems with a strongly inhomogeneous spatial distribution of monomer concentrations.75–77 As an origin of inhomogene-ities, one can mention interfaces, nanoparticles, molecular and supramolecular aggregates (e.g., for macroradicals capable of forming globular conformations), and so forth. They should selectively adsorb one of the comonomers (Figure 9).

Moreover, the spatial scale of the concentration inhomo-geneities should be comparable to the size of the growing macroradical. In addition, the polymerizing monomers should retain their properties after they are incorporated into a polymer chain. In the case of a copolymerization near an interface, one of the types of monomers and chain seg-ments should be preferentially distributed near the adsorbing surface, whereas the other types of monomers and chain seg-ments should be preferentially located in the solution. As a result, the chemical sequence and conformation of the grow-ing macroradical become mutually dependent. Such interdependence determines the further way of chain growth through the comonomer concentrations in a small reaction volume (see Figure 9).

Finally, the rate of copolymerization should be slow enough. This guarantees that, during the chemical reaction,

Page 16: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 703

(a) (b) (c)

Figure 9 Schematic representation of the CDSD copolymerization process in the cases when one of the comonomers is selectively absorbed by a polymer globule (top) or adsorbed on a surface (bottom). (a) Growing chains during the copolymerization (reaction zone around the growing chain end is marked with dashed line); resulting copolymers in (b) globular (adsorbed) and (c) coil-like states. Regions where absorbed (adsorbed) monomers dominate are shown in gray. The conditions in the reaction system compel an initially homopolymeric chain to take the desired template conformation, for example, a globular one. One part of the monomer units is screened from the solution in this conformation, while the other part is accessible for chemical modification.

the equilibrium concentration fields remain approximately constant, and the growing chain possesses an equilibrium con-formation between successive attachments of the monomers. Therefore, the CDSD regime is realized when

max fτD,τrelg < τR ½5� where τR is the reaction time characterizing the polymerization rate, τD is the characteristic diffusion time for monomers, and τrel is the chain relaxation time that depends on the current chain length Nτ. For an ideal chain, the Rouse relaxation time scales with N as τrel � N2.133 It is clear that eqn [3] corresponds to a kinetically controlled regime.

6.18.2.3.2(ii) Copolymerization with simultaneous globule formation Let us assume that we are performing a radical copolymeriza-tion of moderately hydrophobic (H) and hydrophilic (P) monomers in an aqueous medium. The conditions for copoly-merization (e.g., the temperature and solvent composition) should be chosen in such a way that when the emerging chain is long enough it can form a globule. As long as the current chain length Nτ is not too large, the growing hydro-phobic–hydrophilic (HP) macroradical remains in a coil-like conformation. However, when its length becomes sufficiently large, the chain tends to form a two-layer globule with a hydro-phobic core and a polar envelope. The origin of this effect is connected to the loss in translational entropy of covalently bonded hydrophobic species after their incorporation into the growing chain.

The presence of a hydrophobic–hydrophilic interface can dramatically change the reaction conditions. The hydrophobic core will selectively absorb hydrophobic species from the solu-tion (Figure 9), and this will result in a redistribution of monomer concentrations between the core and bulk solution. Because the probability of attachment for each comonomer is determined by its concentration in a relatively small reaction volume near an active chain end, the active center inside the hydrophobic core will mainly attach more hydrophobic spe-cies; on the other hand, when the active center is located on the globule surface, it will mainly attach polar (soluble)

monomers. In this way, the two-layer globule will grow, retain-ing its core–shell structure with a predominantly hydrophobic core and a hydrophilic outer envelope (cf. Figure 9).

Using Monte Carlo and molecular dynamics simulations, 134,135 Berezkin et al. have studied the process of irreversible

radical copolymerization of hydrophobic and hydrophilic monomers in a common solvent. The conformation-dependent polymerization was modeled as a step-by-step chemical reaction of the addition of H and P monomers to the growing copolymer chain, under the assumption that a depolymerization reaction was not allowed. To simplify the analysis, it was assumed that the copolymerization is ideal with equal reactivity ratios. The type of attached monomers and the probabilities of their addi-tion were determined from the average concentrations of the reactive monomers in a reaction volume Vτ around the moving active end of the macroradical of a given length Nτ during a reaction time τR. It was also assumed that the reaction bath was sufficiently large and the reactive species were sufficiently dilute so that there was significant time for the growing macroradicals to propagate independently. The preferential sorption of hydro-phobic monomers in the core of the arising globule was explicitly taken into account.

The simulations134,135 showed that using the conformation-dependent polymerization mechanism, one can indeed end up with a designed copolymer having a dense hydrophobic core surrounded by a hydrophilic shell. Figure 10 presents a typical snapshot of the resultant copolymer globule and an example of the radial distribution that characterizes the distribution of H and P units with respect to the center of the globule. These findings prove explicitly the core–shell microstructure of the globule obtained via irreversible copoly-merization in the solvent, which is moderately poor for hydrophobic species. In general, the same microsegregated structure is observed for core–shell globules obtained via ‘col-oring’ procedure.69–71

The average chemical composition of the resulting HP poly-mer is determined by the bulk concentrations of the monomers and by the solubility of hydrophobic monomers in a globular core. For approximately equal contents of H and P segments in the growing copolymer, the solution should contain an excess

Page 17: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

5 H segments 4 P segments

) r 3

(W 2

1

0 0 2 4 6 8 10

r /σ

704 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 10 Left panel: Typical core–shell microsegregated structure that is obtained via conformation-dependent copolymerization with simultaneous globule formation in a selective solvent. Hydrophobic chain segments are shown in red and hydrophilic segments are shown in green. Right panel: Radial distribution of hydrophobic and hydrophilic segments (with respect to the center of the globule) for a copolymer chain of 512 segments with a 1:1 HP composition. σ is the size of chain segment.

of P monomers. In general, the chemical composition of the synthesized copolymer deviates strongly from the statistical one. Thus, the change in the solution concentration of poly-merizing monomers allows the copolymer composition to vary over a wide range. Interestingly, although the reactivities of both monomers were equal,134,135 their copolymerization led to a copolymer that was preferentially enriched with hydro-phobic segments. Therefore, a strong deviation from the principle of equal reactivity of Flory136 is observed for the CDSD polymerization regime. The type of monomer attached to the growth center during the simulation under kinetic con-trol (at large τR values, see eqn [5]) is determined by the conformation and primary structure of the growing chain as a whole, not only by the local concentration of reactive mono-mers near the active end of the macroradical. As a result of such cooperativity, the formation of sequences with specific LRCs of the LF type was observed.

Because of the specific mechanism of chain propagation, the synthesized copolymers can have a gradient primary struc-ture; that is, the HP composition can change along the growing chain during copolymerization.98 This result follows from the analysis of the distribution of monomers along the growing chain. Very short chains do not contain sufficiently large number of hydrophobic units to form a core–shell con-formation, and the growth of these chains is similar to that observed for random free-radical copolymerization.31 The composition of these chains is also close to the monomer composition in solution. Then, as the macroradical becomes longer and the number of connected H units increases, a dense hydrophobic core can be formed. This core absorbs H monomers, and their fraction in the resultant copolymer increases. In this case, the probabilities of monomer addition are also not constant because of the changes in the ratio between the volumes of the hydrophobic core and the polar shell during the chain growth. Therefore, the gradient primary structure is formed because of a change in the chain confor-mation and a continuous redistribution of comonomers between the globule and the solution in the course of the polymerization; in this way, a compositional drift is produced along the chain.

Note that the statistical model developed by Berezkin et al.134,135 to predict sequence in free-radical polymerization is not applicable in living radical polymerization because of the reversible activation/deactivation procedure between radicals and capping agents.137

6.18.2.3.2(iii) Radical graft copolymerization near a selectively adsorbing surface In a series of publications,138–141 a statistical model was devel-oped for studying graft copolymerization in the vicinity of a surface. This approach represents an irreversible radical copo-lymerization of selectively adsorbed A and B comonomers with different affinity to a chemically homogeneous impenetrable surface, allowing for a strong short-range monomer (A)-surface attraction. Thereby, one of the types of monomers and chain segments are preferentially distributed near the adsorbing sur-face, while other monomers and chain segments are located preferentially in the solution, as schematically depicted in Figure 9. As a result of such concentration inhomogeneities, the chemical sequence and conformation of growing macro-radical become mutually dependent that corresponds to the CDSD regime. It is found that under certain conditions, the adsorption copolymerization yields copolymers with gradient monomer sequence distribution.

Gradient copolymers (GCs) are characterized by local composition varying monotonically along the chain. They have attracted growing attention over recent years. The range of pos-sible applications of these copolymers is quite wide.142–146 In the solutions or melts they can form diverse mesostructures. GCs can be used as surfactants and, particularly, as compatibilizers in polymer blends.145 Grafting of these copolymers yields ‘smart surfaces’, which can change the structure and properties, depend-ing on external conditions.146

GCs are usually obtained via living polymeriza-tion.31,137,147–150 The gradient arises here due to the continuous variation of the monomer ratio during synthesis. However, this technique allows producing sequences with a relatively simple statistics only. The CDSD-based technique represents an alternative approach to the synthesis of copoly-mers with nontrivial statistics, including GCs, via free-radical

Page 18: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 705

polymerization. As indicated above, the sequence of monomer units formed during synthesis and the equilibrium conforma-tion of the reacting macromolecule should be interdependent. This interdependence results from the self-organization of monomers and copolymer macromolecules in the reaction system. We emphasize again that the conformation-dependent copolymerization is most effective when the chain propagation is slower than the diffusion of monomers and the formation of the equilibrium macroradical conformation.

The propagation of the grafted chain was considered as a sequential addition of monomer units of the type A or B to the free active end of macroradical.141 There are four possible ways in which monomer can be added:

kAA� A• þ A →� AA•

kAB� A• þ B → � AB•

kBA� B• þ A → � BA•

kBB� B• þ B → � BB• ½6� where kXY are reaction rate constants of the interaction between active center X and monomer Y. Relations between these con-stants are usually expressed through the reactivity ratios rA = kAA/kAB and rB = kBB/kBA. For simplicity, it was assumed that the rate of propagation does not depend on the chemical structure of the active center (rA = rB = 1), so the ideal copoly-merization was considered. The average concentrations depended on local monomer concentrations in the solution and near the surface. The concentration of the nonadsorbable monomer B was the same everywhere, while the concentration of monomer A in the adsorption layer exceeded that in the solution bulk by a factor q = exp(–ε/kBT) ≡ exp(u), where ε is the energy of a single adsorption contact and u denotes the dimen-sionless adsorption energy. Because chain propagation is usually a kinetically controlled process, the macroradical changes its conformation many times between the propagation acts, and the probability of finding the active center near the surface, fS, should be averaged over the macroradical confor-mations for a current chain length n. This averaging is a significant theoretical problem. However, exact calculation of the statistical weight of each macroradical conformation is possible if the macroradical is considered as an ideal (Gaussian) chain on a lattice.

The computational procedure was realized as follows. The first monomer unit of the grafted chain was placed at the origin. The surface coincided with the XY plane; the z axis was normal to the surface and directed to the bulk of the solution (cf. Figure 11, top left panel). If the distance between the nth monomer unit and the surface zn = 0, then the unit contacted with the surface, and if this unit belonged to the A type, then it was adsorbed.

In order to find the partition function and statistical proper-ties of the growing macroradical, Berezkin et al.141 introduced a variable αn(r) that was defined as follows: αn(r) = 1 if the nth unit located at the point r ={x,y,z} is adsorbed, and αn(r) = 0 if the unit is not adsorbed. If the macroradical of length n has in some conformation k adsorption contacts with the surface, and its active center is located at the point r, the possible number of such conformations is denoted as βn(k,r). The following recur-rence equation determines the value of βn(k,r):

βnðk,rÞ ¼

Xβn−1

k−αn 0

R

ðrÞ,r � ½7� 0ðrÞ

where R′(r) is the set of all possible locations of the monomer unit (n-1) if the nth unit is located at r; r′ is the point belonging to the set R′(r). Then, the partition function of an N-unit chain is given by

nXXA ZN ¼ β ð k

N k,rÞq ½8V

� k¼0

Here, nA is the number of A monomer units in sequence and V denotes the total number of lattice points in the system. Before the addition of the next monomer to the sequence of current length n, the βn(k,r) values are calculated for all r, where βn(k,r) ≠ 0. Then the probability to meet the active center in the adsorption layer fS is found from the equation

nA f ðNÞ ¼ Z −1 S N

XXβN k,r qk 9

Vz 0 k 0

ð Þ ½ �¼ ¼

where Vz=0 is a subset of lattice points with z = 0. The attach-ment of comonomers is determined by their average local concentrations, which, in turn, depend on the corresponding bulk concentrations, 0 and 0�A �B, and on the energy parameter u. The technique developed in Reference 141 allows one to obtain an exact analytical solution of the problem for the Gaussian chain model.

Figure 11 (top right panel) shows that at a fixed solution composition, the adsorption leads to an increase in the fraction �A of adsorbing monomer (A) units in the copolymer, as expected. This is explained by the growth of monomer A con-centration in the adsorption layer.

To analyze the local compositional inhomogeneities along the chain of synthesized copolymers, one can use the probabil-ities of finding a monomer unit of type A at the nth position from the beginning of a growing macromolecule, �A(n). It is clear that for a truly RCP, in which chemically different seg-ments follow each other in statistically random fashion, the �A

(n) function should coincide with the average fraction of A segments for any n. For a random-block copolymer, the frac-tion of one component averaged over many sequences should also be uniform along the chain.

Figure 11 (bottom left panel) presents the local copolymer composition �A(n). It is seen that when u > 0, the simulated polymerization process yields copolymers with a well-defined gradient structure. For such copolymers, the �A(n) function smoothly decreases with n. It should be kept in mind t hat speaking about GCs, we have to consider their ensemble generated by the same synthetic process, not a single chain which can, in principle, have an arbitrary statistics. Note that in the range u = 0.3–0.4, the composition gradient is most pronounced. Interestingly, this adsorption energy is very close to the critical adsorption energy uc estimated for the same model.

To examine the influence of the adsorption energy on the gradient in more specific way, one can introduce a quantitative measure of the gradient as the variance of the preaveraged local copolymer composition:

N N D ¼ N−1

X½f ðnÞ�2− ½N−1 A

Xf 2A ðnÞ� ½10

n¼1 n¼1 �

Page 19: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

(a)

1.0

0.8

(b) 0.6 u

φA 0.0 0.4 0.5

1.0 2.0

0.2 3.0 5.0

(c) 0.0 0.0 0.2 0.4 0.6 0.8 1.0

0 φA

0.52

u 6 0.1 0.2

0.51 5

0.3 10) 4

�n 0.4 1.0

)( AA φ φ (D

0.50 2

u = 0.322 u* c c ≈ 7

0 0.49 0.1 1 10

0 100 200 300 400 500 u n

706 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 11 Top left panel: Graft adsorption copolymerization near the surface that selectively adsorbs one of the comonomers: (a) the reaction system before the polymerization; (b) the chain propagation; (c) the resulting sequence. Top right panel: Fraction of A monomer units �A in 500-unit copolymers vs. the monomer A fraction in the bulk. The adsorption energies, u, are shown in the legend. Bottom left panel: Local copolymer composition (averaged over 104 independent sequences) as a function of monomer position in the chain. Bottom right panel: Value of gradient in composition vs. the dimensionless adsorption energy, at �A = 0.5 and N = 500.

The value of D is shown in Figure 11 (bottom right panel) as a function of u. It is seen that the dependence of D on the adsorption energy goes through a well-pronounced maximum located at u ≈ uc when the largest compositional nonuniformity is observed (cf. Figure 11, bottom left panel).

A detailed analysis of the model showed that the key factor determining the microstructure of the resultant copolymer is the dependence of the probability fS(n) on the current macro-radical length n. This could be considered as a ‘memory function’. Its behavior is strongly influenced by the adsorption

energy u. There are three regimes of the adsorption copolymer-ization that correspond to different adsorption energies:

1. In the nonadsorbing regime (u ≪ uc), the probability for the active center to be located near the surface scales as fS(n) ∝ n−1, where n is the current macroradical length. Therefore, in this regime, the chain propagation leads to asymptotical convergence of the local copolymer composi-tion to that of RCPs synthesized in a solution under the same conditions. For finite values of n, a fast change in the

Page 20: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 707

local chemical composition is observed for initial sections of the growing macroradical.

2. In the weak adsorption regime (u ≈ uc), the probability fS is nonzero for any n and u. As a result, the local copolymer composition in the long-chain limit differs dramatically from that of a RCP synthesized in the solution bulk. In the vicinity of uc, the magnitudes of the compositional nonuni-formity and gradient are maximal, and the gradient extends along the entire chain for any chain length.

3. For the strong adsorption regime (u ≫ uc or u > u*c ), the copolymer statistics does not depend on u and is deter-mined solely by the solution monomer concentrations.

Thus, the change in the adsorption energy can provide versatile and accurate control of the statistics of the resulting copolymers obtained via the conformation-dependent adsorption copoly-merization. It is known that the gradient (tapered) nature of copolymers, which can be synthesized in free-radical polymer-ization processes, is due to a drift in the free monomer composition during solution polymerization.31 Such copoly-mers can be considered a special type of block copolymers in which the composition of one component varies along the chain. With a decreasing difference in the monomer reactivity rations, the formation of gradient statistical copolymers rather than gradient (tapered) block copolymers occurs. However, in the model polymerization process considered in References 138–141, all the polymerizing species had the same reactivities, and the monomer concentrations remained unchanged during synthesis. Therefore, the changes in the probabilities of the addition of components to the growing macroradical are due to the evolution of its chemical composition; this result is typical for the CDSD regime.

6.18.2.3.2(iv) Radical copolymerization near a patterned surface Because the spatial scale of the monomer concentration gradi-ent in the reaction system discussed in the previous sections was comparable with the size of the macroradical, the chemical inhomogeneity has been observed for the generated copolymer sequence as a whole. To limit and control the size of the gradient regions, the spatial scale of the concentration inhomo-geneities in the reaction system should be limited. The easiest way to achieve this is to perform the copolymerization near a solid patterned surface with discrete adsorption sites. In this case, each of the sites represents a small independent source of concentration disturbances lying at a certain distance from other sites. Surfaces with a regular distribution of adsorption sites are of most interest because they can allow fine control of the primary structure of the copolymers obtained with the CDSD technique.

Another motivation for this work is the development of copolymers that are tuned to a certain surface, that is, copoly-mers that have a ‘memory’ of the preparation conditions and are able to reproduce their specific conformation in the vicinity of the surface with predefined chemical heterogeneity. It is believed that copolymers designed in this way could have potential for the recognition of patterned substrates through the formation of stable adsorption complexes with planar or spherical substrates composed of two chemically distinct sites, one of which has a preferential affinity for one of the

comonomers. Obviously, these copolymers could have a great potential in molecular technology and biotechnology.

A Monte Carlo simulation of irreversible template copoly-merization near a chemically heterogeneous surface with a regular (hexagonal) distribution of discrete adsorption sites was performed and reported in References 151 and 152. The sites could selectively adsorb one of the two polymerizing monomers and the corresponding chain segments from solu-tion. The focus of this study was on the influence of the polymerization rate, adsorption energy ε, and distance between adsorption sites rs on the chain conformation and chemical sequence of the resulting AB copolymers and, specifically, on the coupling between polymerization and selective adsorption.

Under the preparation conditions corresponding to the CDSD regime, the formation of quasi-regular copolymers with blocky primary structure was observed.151 In such copo-lymers, there are two types of alternating sections. One of them contains randomly distributed A and B segments. The second one consists mainly of strongly adsorbed A segments. The average length of the random sections is proportional to the distance separating the nearest neighbor adsorption sites rs. The average length of the A-rich sections is determined by the adsorption capacity of the adsorption sites. Therefore, by vary-ing the interaction parameters and the distribution of adsorption sites on the substrate, one can design and synthesize copolymers with different surface-induced chemical sequences in a controlled fashion.

Under the preparation conditions corresponding to the CDSD regime, the formation of quasi-regular copolymers with blocky primary structure was observed. A typical snapshot from the simulation is presented in Figure 12(a). In such copolymers, there are two types of alternating sections. One of them contains randomly distributed A and B segments, the second one consists primarily of strongly adsorbed A segments. The average length of the random sections is proportional to the distance separating the nearest neighbor adsorption sites rs. The average length of the A-rich sections is determined by the adsorption capacity of the adsorption sites. Therefore, by varying the interaction parameters and the distribution of adsorption sites on the substrate, one can design and synthesize copolymers with different surface-induced chemical sequences in a controlled fashion.

In particular, the variation of the strength of the effective monomer (A)–substrate interaction ε allows us to determine the asymptotic regimes corresponding to random copolymer-ization or CDSD. The average intramolecular chemical composition �A emerging in the simulation is shown as a function of ε in Figure 12(b) for the case in which the solution concentrations of A and B monomers are equal, C0

A ¼ C0 B. In the

weak adsorption regime, ε/kBT ≲ 6, random copolymerization dominates (�A ≈ 0.5); in the region ε/kBT ≳ 15, CDSD domi-nates. The threshold value of the energy parameter ε when random copolymerization and CDSD have about the same probability is ε/kBT ≈ 8 for a given choice of the polymerization parameters and simulation model.

To analyze the correlations in the copolymer sequences, one can use the so-called two-point chemical correlators:97

X ΘðiÞ ðiÞ ðiÞ

= n ½11�αβ ¼ nαβ αβ α ,β

Page 21: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

(a)

(b) 1.0 (c)

0.9 0.6

ε /kBT 0.8

0.5 20

A 15

ϕ 0.7

) AA

i(

0.4

Θ 100.6 0.3

8 6 0.5

0.2 0 5 10 15 20 0 20 40 60 80 100 ε /kBT i

708 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 12 (a) Snapshot illustrating a typical conformation of a 512-unit copolymer chain with excluded volume interaction synthesized near a patterned surface in the strong adsorption regime. Chain segments of A and B type are shown as red and green sticks, respectively. Spheres depict adsorption sites. (b) Average chemical composition as a function of the adsorption energy parameter for the case when the solution concentrations of A and B monomers are equal. (c) Chemical correlators for AA diades for different values of the adsorption energy parameter ε.

where nðiαβÞ is the number of αβ (AA, BB, and AB) pairs and i is

the ‘chemical distance’ between these pairs along the chain. Chemical correlator ΘðiÞ

αβ represents the probability of finding a pair of α and β segments (α, β = A, B) among all the possible pairs separated by i segments along the chain in a given copo-lymer sequence. Some of the results of the calculation151 are shown in Figure 12(c) for a few different values of ε.

The dependences of ΘðiÞ AA on i indicate that quasi-regular

copolymers are synthesized in the CDSD regime when the adsorption interaction is sufficiently strong. Indeed, we see that these copolymers are characterized by a periodic variation of the composition along the chain. For the system simulated, the period of this variation is about 20 segments. Therefore, the formation of copolymers with blocky primary structures can be observed.

The following mechanism of the formation of quasi-regular copolymers was suggested.151 Chain growth begins near an adsorption site at which the concentration of strongly adsorbed A monomers is high. These monomers are attached preferably to the end of the growing macroradical and form an initial chain section comprising mainly A segments. This section gra-dually covers the nearest adsorption site. Because of the limited adsorption capacity, the screened adsorption site looses the ability to attract the active end group of the macroradical. Afterwards, the reaction volume moves into solution away from the substrate in which random copolymerization occurs. The random chain section grows until it reaches the nearest free adsorption site. Subsequently, the formation of a new adsorbed

section commences again. Such cycles are repeated many times. As a result, the formation of a quasi-regular copolymer with alternating A-rich and random AB sections is observed.

Thus, the copolymer sequence consists of repeating blocks. Each of these blocks comprises a short ‘gradient sequence’ formed by strongly adsorbed sections and random weakly adsorbed sections. The average length of the adsorbed section is nearly constant and is related to the adsorption capacity of the adsorption center. In contrast, the average length of the random section depends on its conformation between the adsorption centers, and in principle, it can vary over a wide range. It should be emphasized that the quasi-regular copoly-mer is formed only when the random bridges connecting neighboring adsorption sites are strongly stretched and the variation of their length is sufficiently small. The strongly stretched regime leads to the periodical composition varia-tions. Figure 12(c) shows that this regime is realized for ε/kBT ≥ 15.

The distribution of B blocks, which are included mostly in nonadsorbed chain sections, decays exponentially and thus should obey Bernoullian statistics that correspond to a zeroth-order Markov process.55 The average length of such blocks is close to 2, which is the same as that of a RCP. In the case of A blocks, the distribution function fA(ℓ) also decays exponentially in the initial region, which corresponds to short blocks included in the random chain sections. For longer A blocks, however, the distribution becomes significantly broader and has a local maximum at ℓ � 10.152 Hence, one

Page 22: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 709

can conclude that the distribution of A blocks strongly deviates from that known for random sequences.

By varying the distance between nearest adsorption sites, rs, one can control the composition variation period of the synthe-sized copolymer. From the chemical correlators defined by eqn [11], it is easy to find the average number of segments in the repeating chain sections, N , for different rs values. It is instruc-tive to analyze the relation between N and rs. As expected, a

μpower law N ∝ r is observed. It is clear that the value of the s exponent μ in this dependence should range be between 1 (for a completely stretched chain) and v−1 with v ≈ 0.6 (for a random coil with excluded volume133). The calculation152 yields μ ≈ 1.33 for N ≥15. This supports the aforementioned assump-tion that the repeating chain sections are strongly stretched between the adsorption sites. The same conclusion can be drawn from the visual analysis of typical snapshots similar to that presented in Figure 12(a).

6.18.2.3.3 Evolutionary approach and theoretically informed sequence design 6.18.2.3.3(i) Design as a simulation of evolutionary process The concept of evolution of primary sequences of biopolymers has attracted significant interest from biologists, chemists, and physicists for a long time.153–156 As has been discussed, it is natural to expect that the content of information in the sequences of biopolymers (proteins, DNA, RNA) is relatively high in comparison with random sequences where it should be almost zero.62 Presumably, the information complexity of early ancestors of present day biopolymers has been increased in the course of molecular evolution when the copolymer sequences became more and more complicated.154 The study of various possibilities of this evolution of copolymer sequences is just the area where the concept of evolution can be used in the context of polymer science.

It is worthwhile to note that since the information content of sequences can be represented as a mathematically defined quantity, the whole process of evolution of biopolymer sequences can be specified in exact mathematical terms. The formulated fundamental problem is extremely difficult to han-dle because of the absence of direct information on the early prebiological evolution. Therefore, of particular interest are ‘toy models’ of evolution of sequences that show different possibi-lities for appearance of statistical complexity and of LRCs in the sequences.

Evolutionary computation approaches represent optimiza-tion methods. They are conveniently presented using the metaphor of natural evolution: a randomly initialized popula-tion of individuals evolves following a crude parody of the Darwinian principle of the survival of the fittest. New indivi-duals are generated using simulated evolutionary operations such as mutations. The probability of survival of the newly generated solutions depends on their fitness (how well they perform with respect to the optimization problem at hand): the ‘best’ are kept with a high probability, and the ‘worst’ are rapidly discarded.

Some computer models describing the evolution of copo-lymer sequences have been proposed previously.46,47 Most of them are based on a stochastic Monte Carlo optimization principle (Metropolis scheme). Such optimization algorithms start with arbitrary sequences and proceed by making random substitutions biased to minimize relative potential energy of

the initial sequence and/or to maximize folding rate of the target structure.

As has been emphasized, the problem that we address here is somewhat different from that usually discussed in the context of protein physics. We do not intend to search for unique three-dimensional (native) conformations with fast folding rate. On the contrary, we are interested in the state with large entropy. In general, our aim is to learn whether it is possible to make with synthetic copolymers a step along the same line as molecular evolution.

Ascending and descending branches of sequence evolution. The aim of the study68 was to introduce explicitly the concept of sequence evolution into the CDSD scheme. Using the method of molecular dynamics, the conformation-dependent evolu-tion of model HP copolymer sequences was simulated.68 The sequence evolution mechanism involved the generation of an initial protein-like sequence by inspecting a homopolymer globule and by attributing H type to the monomeric units in the core of this globule and P type to the units on the surface of the globule. The resulting copolymer was then transferred to a coil conformation and then refolded by means of strong attrac-tions acting among the H units. The HP sequence was further modified, depending on the position of a monomer in the core or on the surface of a newly formed globule. Such modifica-tions, leading to changing the primary HP sequence, are repeated many times (�103). With this evolutionary process that can be called ‘repeated coloring’, structures and sequences are formed self-consistently.

A 128-unit flexible-chain heteropolymer with an HP com-position fixed at 1:1 was simulated for the condition when hydrophobic H monomers strongly attract each other, thus stabilizing a dense globular core. Concurrently, the affinity energy εPP among two hydrophilic P monomers was considered as a parameter (the interaction between H and P monomers is pffiffiffiffiffiffiffiffiffiffiffiffiffiffigiven by εHP ¼ εHHεPP). For this model system, various conformation-dependent and sequence-dependent properties, including information-theoretic-based quantities, can be calculated.

Depending on the attraction energy between polar segments εPP, it is possible to find two regimes (branches) of evolution (regimes I and II).68 If εPP is smaller than some crossover energy ε*PP (regime I), the evolution can lead to a second-order-like transition in sequence space from the sequences with a protein-like primary structure capable of forming a core–shell globule to the degenerated (nonprotein-like) sequences having long uniform H and P blocks (cf. Figure 13, left panel). This transition is also accompanied by large changes in the confor-mational properties of copolymer.

Figure 13(a) (right panel) shows the mean square gyration radius, R2, plotted versus εPP. As seen, R2 is a weakly decreasing g g function of εPP in the range εPP > ε*PP and demonstrates a rapid growth when εPP decreases and becomes less than ε*PP. The critical value ε*PP is found to be smaller than the critical energy at which coil-to-globule transition takes place in a homopoly-mer chain of the same length.

The degenerated primary structure resembles a di- or tri-block sequence (‘core–tail’ or ‘tadpole-like’ conformation), as seen in Figure 13.

Thus, when the attraction between the hydrophilic seg-ments is not sufficiently strong, we deal with the ‘descending branch’ of the evolution, which leads to nonprotein-like

Page 23: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

(a) (b) (a) 40

30

2 R

g 20

10

00 1 2 3 4 5(b) 0.10

(bi

t)

0.08

N/JS 0.06

0.04 0 1 2 3 4 5

εPP/ε* PP

710 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 13 Left panel: Snapshots of two typical conformations of designed copolymers obtained after long evolution (3.63 � 108 MD time steps) of sequences. (a) Core–tail (tadpole-like) structure at εPP = 0. (b) Core–shell structure at εPP = 0.3 kBT. Hydrophobic and hydrophilic units are colored red and green, respectively. Right panel: (a) mean square gyration radius and (b) Jensen–Shannon divergence measure as a function of the attraction energy εPP between hydrophilic segments, after the sequence evolution procedure. The characteristic energy of H–H interactions is fixed at εHH = 2kBT, thus stabilizing a dense globular core.

sequences having low information content and low complex-ity. In contrast, in the second regime (at εPP ≥ ε*PP), the complexity of protein-like structures is found to increase and therefore we have the ‘ascending branch’ of the evolution.

Using the Jensen–Shannon divergence, J S (cf. eqn [1]), as a measure of complexity for the generated sequences, one can obtain an interesting result (cf. Figure 13(b), right panel). The most important feature is that the J S value is a nonmonoto-nous function of εPP, whereas Shannon’s entropy, Shannon’s index, and many other sequence-dependent parameters change always gradually.68

For the sequences generated in the evolutionary process described above, it was shown that at εPP ≥ ε*PP (regime II), the degree of complexity, as measured by J S divergence, can be considerably higher compared to that observed for the regime I, at ε*PP < ε*PP. The complexity increases slightly with decreasing εPP, reaches a maximum just on the boundary of regimes I and II, and then drops sharply (Figure 13(b), right panel). Therefore, in the regime II, the evolution preserved the copoly-mer sequence of high complexity, whereas in the regime I, the information content of the sequence has degenerated in the course of evolution.

Simultaneous evolution of sequences and conformations. Chertovich et al.157 reported on simultaneous evolution of sequences and conformations. This design procedure leads to the final state that depends on the set of interaction parameters and on the rearrangements both in conformational space and in sequence space. These rearrangements are characterized by the usual thermodynamic temperature, T, for conformational space as well as effective sequence rearrangement temperature, Ts. Namely, after certain number of Monte Carlo steps in con-formational space (in the course of this process the system equilibrates at temperature T), the possibility of mutation of

two randomly chosen monomeric units is attempted: the monomeric unit H converts into P and vice versa. The move is accepted provided this process leads to a decrease in the energy of a protein-like globule. If the globular energy is increased by the amount ΔE, this move is accepted with the probability defined by exp(–ΔE/kBTs).

In general, we can define the evolution of sequences for any value of Ts. However, the simulation for three characteristic cases can be most easily understood.

1. The inequality T ≪ Ts means that all the moves in sequence space are accepted independent of conformation. This cor-responds to random mutations, and the final sequence (after long evolution) will be that of a random HP copoly-mer (no information complexity).

2. The inequality T ≫ Ts means that only the moves leading to a decrease in the globular energy E are accepted. Such an evolu-tion should lead to a sequence corresponding to a minimum of globular free energy in conformational space. It was shown68 that in the absence of any attractive interactions between P units the final sequence after evolution should have hydrophobic core with very few hydrophilic (or polar) loops and a long hydrophilic tail. This sequence is close to that of HP diblock copolymer, and should not exhibit any information complexity, as has been stated above.

3. The case T = Ts corresponds to annealed HP sequence. This case is equivalent to the situation when monomeric unit H can be converted into P by attaching some ligand L:P ⇄ H+L. We assume that the number of ligands is fixed to maintain 1:1 HP composition; however, they can choose which mono-

meric unit to bind to. This defines, in particular, the chain

Page 24: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 711

sequence. Annealed HP sequences in the context of polymer globules were first considered by Grosberg.158

When Ts does not correspond to any of the characteristic cases described above, the evolution of sequences coupled with con-formations can be still defined in the same way. One has to only remember that if T ≠ Ts and both T and Ts are finite, there is a flow of heat between conformational space and sequence space, so that full thermodynamic equilibrium is impossible. Still, we can be in a stationary regime corresponding to the sequences tending to a certain fixed point, and possessing (or not possessing) information complexity.

The simulations and theoretical arguments157 predict that at high Ts, the sequence free energy Fs dominates and the sequence tends to be completely random, corresponding to the mini-mum of Fs. At low Ts, the evolution selects those sequences, which correspond to the low value of the conformation free energy (structure of the core–tail type). In the intermediate regime, there is an interplay between the conformational ther-modynamic force and the sequence contribution (evolution pressure). Therefore, the formation of nontrivial structures and sequences is possible. Even in the absence of attraction between P units, the final sequences remain protein-like (i.e., the final copolymer has the core–shell structure of a globule) and therefore maintain certain information complexity.

6.18.2.3.3(ii) Theoretically assisted sequence design The examples of sequence design described above belong to a spectrum of in silico methodologies that can be called knowledge-based model approach. A knowledge-based approach requires some form of a priori knowledge about a particular target model (e.g., core–shell globule) and is there-fore limited in its applicability by the models that are available. As has been noted, there is another sequence-design strategy where existing theoretical concepts are employed. In this sec-tion, we will discuss a simple theoretically assisted algorithm that introduces a ‘selection pressure’ under which a random two-letter (AB) copolymer sequence can mutate and transform into the sequence tuned to microphase separation transition (MIST). In particular, we are interested in determining how a sequence of A’s and B’s should be organized in order to reach maximum characteristic length scale for MIST at a given tem-perature, AB composition, and polymer chain length N. It is natural to call such a sequence ‘MIST-tuned sequence’.

It is well known that heteropolymers can self-assemble into highly ordered patterns of microstructures, both in solution and

4,159–161 in bulk. This subject has been reviewed extensively.1–

The driving force for structure formation is competing interac-tions, that is, immiscibility of chemically different monomer species, on the one hand, and covalent bonding of units within the same macromolecule, on the other hand. The latter factor prevents the separation of the system into homogeneous macro-scopic phases. Spontaneous ordering of copolymers, which occurs with changing (typically decreasing) temperature and is accompanied by an explicit symmetry change, is treated as MIST or order-disorder transition (ODT) and the respective tempera-ture is the ODT temperature. With further decrease of temperature, the ODT is often followed by various order–order transitions between the different ordered morphologies.

Khalatur et al.162 proposed a theoretically assisted design scheme that optimizes monomer distributions and leads to the MIST-tuned sequence capable of forming microphase-separated structures with a required (micro) domain spacing r*. In this evolutionary approach, a simulation is performed to search for ‘point mutations’ that favor the microphase separation of copolymer melt. A random AB sequence is taken as an initial generation (G = 0), which is considered as a ‘common ancestor’ of a given run. Then a procedure of the evolution (annealing) of the sequence starts. The iterative procedure consists of many mutation steps. At each step – with every ‘click of the evolutionary clock’ – two monomers are chosen randomly and, if they happen to be of different types, an attempt is made to exchange their types (A ↔ B). This point mutation changes the copolymer sequence (S) and the corresponding sequence-dependent intrachain cor-relation function ω(q,S) calculated in the reciprocal q-space. The ω(q,S) function is used as an input for some theory (see below) that gives the transition temperature T* (or the critical value of the Flory–Huggins parameter χ* ∝ 1/T*) and the wave number of maximum instability q*(=2π/r*). These quantities can be viewed as a scoring (fitness) function f that describes how well a particular sequence is optimized to match the target property. There should be a feedback in the system. In other words, to accept or reject current mutation, one should com-pare two fitness functions, old and new ones. This is done following the Metropolis ideology, using some design para-meter, which is called design (or sequence) temperature. This parameter characterizes the tolerance to mutations in sequence space or, in other words, an ‘evolution pressure’. In fact, this is the same parameter as that discussed in Section 6.18.2.3.3(i).

If the ODT temperature T* is viewed as a fitness function f = T*(S), the resulting change in the transition temperature ΔT* after an attempted mutation is found and the probability p to fix the mutation is guided by the standard Metropolis algorithm: p =1, if ΔT* ≥ 0, otherwise p = exp(ΔT*/Ts), where Ts is the fictitious temperature referred to as sequence design temperature. It is assumed that after N point mutations, the N-unit sequence is passed on to the next generation, G → G +1. Such modifications, leading to changing in copolymer sequence, are repeated many times (nG � 106). For each trajec-tory corresponding to a stationary regime, one may interpret the set of sequences generated in the course of the evolutionary process as n ‘G different species’ originating from a common ancestor. It is reasonable to constrain the sequence composi-tion so that there are fAN units of type A and (1–fA)N units of type B.

In the weak-segregation regime, the phase behavior of a polymer melt composed of flexible-chain macromolecules can be described on the basis of the random-phase approxima-tion (RPA)159 or the polymer integral equation reference interaction site model (pRISM) theory163,164 that allow finding the conditions under which the spatially homogeneous state of the system becomes unstable.

Using the simplest incompressible random phase approxi-mation, the critical value of the Flory–Huggins parameter, χ*, and the corresponding transition temperature, T* ∝ 1/χ*, are determined by the condition that the scattering intensity S(q) reaches its maximum value at a particular nonzero wave number q*. Within the RPA, the scattering intensity is given by

Page 25: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

(a) (b) 0.10 1.50

N = 128 N = 128 N = 256 N = 256

N = 512 N = 5120.08 1.25

N = 1024 N = 1024

0.06 1.00

N

1/2

*/ N

T */r

0.04 0.75

0.02 0.50

0.00 0.25 0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0

Ts Ts

712 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

S−1 RPA ðqÞ ¼ RðqÞ−2χ ½12�

with

1 ð ωAAðqÞ ω

qÞ ¼ þ BBðq 2ωABðqÞ R

Þ

δω q fB fA þ

fA fB

�½13� ð Þ

where the intramolecular correlation functions ωαβ characterize the conformation of a macromolecule and its sequence distri-bution, fA = NA /N and fB = NB/N, �A and �B are the volume fractions of the corresponding monomer species, and δω is defined as

δωðqÞ ¼ ωAAðqÞωBBðqÞ− f −1A f−1B ω

2 ABðqÞ ½14�

To a first approximation, one can treat polymer chains on the basis of the highly simplified unperturbed model without intramolecular excluded volume interactions. This allows us to considerably simplify the problem by calculating the ωαβ

using the Gaussian function exp(–q2σ2|i–j|/6) which describes the distribution of segments i and j belonging to the species A and B inside an N-unit polymer.

In pRISM theory, the MIST process is directly reflected in the normalized static structure factor, S(q). As a copolymer system is cooled and microdomains are forming, the peak scattering intensity grows in a mean-field manner corresponding to the linear portion of the S(q) curve in the coordinates S−1(q)–χ. Extrapolation of this linear portion to divergent intensity defines an apparent mean-field spinodal temperature .164

‘ ’

For the T*(S) fitness function used in the sequence-selection process,162 it was found that if the sequence design temperature Ts is too high, the design can be viewed as a random walk in sequence space, and it yields random sequences, as expected. In contrast, when Ts is too low, A and

B units tend to be separated within the chain and the evolu-tionary algorithm leads to trivial symmetric diblocks for fA = fB =½. In the Ts → 0 limit, one observes a typical mean-field behavior: T* ∝ N and r* ∝ N1/2 (for long symmetric diblocks, T*/N = 0.096 or χ*N = 10.4). This is well seen in Figure 14 where the transition temperature T* and the characteristic length scale r* (=2π/q*) are shown as a function of Ts for different N. In between the Ts → 0 and Ts → ∞ regimes, there

cÞis a certain critical design temperature Ts ð

≈ 0:78 at which the transition from random to nonrandom sequences occurs.

What is rather unexpected here is that we find nonzero design temperature near which the r* value has a maximum for sufficiently long chains, N ≥ 200. In other words, there are such sequences, which produce microphase-separated struc-tures with larger characteristic length scales than simple symmetric diblocks (at least, in the weak segregation regime). Another important observation is that the transition in sequence space from random to nonrandom sequences occurs as a first-order-like transition. Indeed, the calculations pre-

cÞdicted that near Ts ð the distribution function W(r*) for the

ensemble of generated sequences demonstrates a clear bimodality.162

The most intriguing observation that can be done for this transitory regime is that sufficiently long sequences providing the maximum of r* do not correspond to simple diblocks but rather they have a gradient shape, or it maybe better to say, an S-like shape (Figure 15(a)). Of course, among all generated sequences, symmetric diblocks have the highest value of T* (i.e., their critical χ parameter is lowest). However, S-like sequences can show larger characteristic length scales.

We now analyze the sequence design process with the value of r* used as a fitness function. The maximization of the fitness

Figure 14 (a) Transition temperature T* and (b) characteristic length scale r* (=2π/q*) as a function of the sequence design temperature for different chain lengths N at 1:1 AB composition. Both T* and r* are averaged over the ensemble of generated sequences (�106).

Page 26: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

(a) 1.0 Ts= 0.20 Ts= 0.40 T

0.5 s= 0.60 Ts= 0.70 Ts= 0.74

1 − Ts = 0.80

) i( A 0.0

p 2 random

−0.5

−1.0 0.0 0.2 0.4 0.6 0.8 1.0

i /N

(b) 1.0

0.5

1 −) i( A 0.0

p 2

rs = 0.01

−0.5 rs = 0.05

rs = 0.50

−1.0 0.0 0.2 0.4 0.6 0.8 1.0

i /N

Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 713

Figure 15 Composition profiles shown as a function of reduced monomer number i/N for 512-unit ‘MIST-tuned’ copolymer sequences having 1:1 AB composition. The present definition assumes that the A-type monomers are coded by symbol +1, whereas symbol −1 is assigned to the B-type monomers. For an ideal RCP in which chemically different units follow each other in a statistically random fashion, the probability pA that monomer A is located at the i th position in the chain is ½ for any i. (a) Optimization of the transition temperature T* at different Ts leads to sequences that have a gradient (or S-like) distribution of A and B monomers, as schematically depicted in the right panel. (b) Optimization of the characteristic length scale r* at different rs leads to sequences that have a tapered (or ‘bolas-like’) distribution of A and B monomers, as schematically depicted in the right panel.

function f = r*(S) with the sequence design parameter rs also leads to a gradient-like composition profile (Figure 15(b)). In this case, however, we deal with a copolymer whose primary structure is similar to that known for tapered or ‘bolas-like’ copolymers exhibiting strong composition inhomogeneity along their chain. When rs → 0, the r* value for MIST is signifi-cantly larger than that observed for diblock copolymers with the same AB composition and the same chain length.162 Note that both RPA and pRISM lead to qualitatively similar results.

Using the Jensen–Shannon divergence, J S , as a measure of complexity for the generated sequences, one can obtain an interesting result (cf. Figure 16). The most important feature is that J S value is a nonmonotonous function of the sequence design temperature Ts. Another intriguing result is that the Jensen–Shannon divergence measure shows a maximum in the vicinity of Ts

ðc Þ .

For the sequences generated in the sequence design process described above, it was found that at Ts ≈ Tsðc Þ, the degree of complexity, as measured by J S divergence, can be considerably higher as compared to that observed for very low and high Ts. The complexity increases with Ts decreasing, reaches its max-imum just near Tðc

s Þ, and then sharply drops (Figure 16).

Therefore, in the vicinity of Tsðc Þ, the copolymer sequences of

120

N = 128 100 N = 256 N = 512

80 N = 1024

JS/b

it

60

40

20

0 0.0 0.5 1.0 1.5 2.0

Ts

Figure 16 The Jensen–Shannon divergence measure as a function of the sequence design temperature for different chain lengths N.

high complexity are generated, whereas at low Ts, the informa-tion content of the sequence is degenerated in the course of the design. One can say that the most interesting and unusual

Page 27: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy714 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

sequences appear at the ‘edge of chaos’ when Ts is close to the transition temperature Ts

ðcÞ .

6.18.3 Synthesis of Designed Copolymers

Three experimental approaches have been employed to synthe-size PLCs. One of them utilizes synthesis of copolymers from monomers having various solubilities in a common solvent. Changing the solution conditions leads to local agglomeration of monomers; the synthesis performed under these conditions produces PLCs. One drawback of this approach is that copoly-mers thus formed comprise typically multiple fractions having varying chemical composition and comonomer sequences. Various separation methods have to be employed to fractionate the resulting macromolecules into monodisperse (composi-tion and monomer sequence) fractions. Another method reported in the literature involves ‘grafting through’ approach of (typically hydrophilic) oligomeric or polymeric grafts onto a backbone of a parent macromolecule. The parent polymer is a ‘truly random’ copolymer containing responsive segments (i.e., those that respond to variation in temperature, i.e., N-isopropylacrylamide) and hydrophilic reactive units (i.e., glycidyl methacrylate) that will serve as attachment points for the macrografts. The characteristics of these PLCs will depend on the density and length of the grafts, the chemical composition of the grafts and the spatial distribution of the grafts along the parent macromolecule. In case that thermore-sponsive segments are present in the parent RCP, the latter can be adjusted by varying the solution temperature during the ‘grafting through’ reaction. Finally, the last approach leading to PLCs reported in the literature follows the original sequence design scheme suggested by Khokhlov and Khalatur.69–77

Specifically, a parent homopolymer is placed in a selective solvent, where it adopts either an expanded (good solvent), Gaussian (theta solvent), or collapsed (poor solvent) confor-mation. A ‘chemical coloring’ species capable of modifying chemically monomers inside the parent homopolymer is added to the solution and it reacts with those monomers in the parent homopolymer that are accessible to the reaction thus converting them into another species. As will be discussed in detail below, this chemical modification can follow either che-

þB −Bmical addition (A → C) or elimination (A → C). If no conformational change takes place during the reaction, the distribution of the newly formed ‘coloring’ segments in the PLC will be governed by the original conformation of the parent macromolecule. In reality, the situation may be more complex because the solubility and thus the overall con-formation of the PLC changes with increasing the ‘degree of coloring’. Depending on the interplay between the solubility of the original and newly created units in the copolymer, the chain may either expand (solubility of the newly added species is better than that of the original monomeric units) or can further collapse or even invert (solubility of the newly added species is worse than that of the original monomeric units). This behavior was predicted in computer simulations101

(see Section 6.18.2.3.1(iv)). Below we provide more detailed description of the experi-

mental approaches employed in the synthesis of PLCs.

6.18.3.1 ‘Grafting through’ Approach

Early experimental realizations pertaining to the formation of PLCs appeared in the early 2000s. Virtanen and Tenhu165 pre-pared RCPs comprising N-isopropylacrylamide (NIPAAm) and glycidyl methacrylate (GMA) or N-acryloylsuccinimide (NASI) units and grafted amine-terminated poly(ethyleneoxide) (PEO) grafts to the GMA (or NASI) moieties.166,167 Light scat-tering studies on those systems have revealed that the conformation of P(NIPAAm-g-EO) copolymers depends on the interplay between the solubility of PNIPAAm, a thermo-responsive polymer, which exhibits the lower critical solution temperature (LCST) behavior in aqueous solutions at ≈ 32 °C, and PEO and the conditions of the conformation of the parent P(NIPAAm-co-GMA) chain to which the PEO grafts were added. Although some of the observed properties (i.e., the critical solution temperature or viscosity) could be associated with varying number of PEO grafts in systems prepared under good and nearly critical conditions (15 and 29 °C), other char-acteristics (i.e., the sizes of aggregates under poor solvent conditions, fluorescence electron paramagnetic resonance) reflect various location of the PEO grafts along the copoly-

166–168mer. The authors noted that this suggest ‘conformational memory’ of the parent polymer.

6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization

Another approach toward preparing PLCs involves direct copo-lymerization of monomers or macromonomers under variable solubility conditions. This method is based on selective solu-bility of the two monomers leading to aggregation of the monomers and resulting in more blocky character of the resul-tant RCP. In early studies, Lozinsky et al.169–172 reported on radical copolymerization of mixture of N-vinylcaprolactam (NVCL)/N-vinylimidazole (1-VIA) mixtures in 10% aqueous dimethyl sulfoxide at temperatures both below and above phase separation. The researchers isolated three fractions of NVCL/1-VIA copolymers. Differential scanning calorimetry stu-dies revealed that one fraction underwent conformational transition predicted by the theory.134,135

Siu et al.173 later reported on copolymerization of NIPAAm and N-vinylpyrrolidone (VP) at various temperatures. While PNIPAAm/PVP copolymers synthesized at temperature below the LCST, copolymerization at temperatures higher than LCST resulted in a globular ‘protein-like’ comonomer sequences.

Wahlund et al.174 copolymerized 1-VIA with NIPAAm and employed affinity chromatography to separate the resultant copolymers into two fractions with different physico-chemical characteristics. While copolymers with truly random distribu-tions of the two monomers did not interact with the metal chelate iminodiacetic acid Sepharose CL 6B (Cu2+-ID Asepharose), the other fraction comprising PLCs adsorbed readily. The authors reconciled the observed behavior by the accumulation of the hydrophilic pendant 1-VIA groups respon-sible for complexing at the hydrophilic periphery of the resultant copolymer.

One of the major drawbacks of the ‘direct copolymeriza-tion’ method is that one typically ends up with multiple fractions of the monomers and thus resulting comonomer distributions of the resulting copolymers. The authors noted

Page 28: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 715

that the various fractions have to be isolated via tedious separa-tions, however.

In a subsequent study, Kumar et al.175 used the P(NIPAAm-co-1-VIA) copolymers to purify histidine-tagged green flourescent protein (His-tag GFP) from recombinant E. coli by copper-chelate affinity precipitation. The authors noted that the comonomer distribution in the copolymer played a critical role in the interaction between the copolymer and the His-tag GFP, while complete elution of the affinity-bound Cu P(NIPAAm-co-1-VIA) was achieved with PLC comonomer sequences (imidazole groups exposed to the outer solution), no recovery was obtained with IMAC non-bound copolymer fraction (imidazole groups unexposed). More details pertaining to the behavior of these systems can be found in a review by Lozinsky.176

There were other reports in the literature that reported on direct copolymerization of various monomers that may lead to PLC.

With some exceptions,177 one of the monomers in the copolymer was a thermoresponsive unit,178 that is, NIPAAm, whose presence endowed the resultant copolymer with ther-moresponsive nature.179,180 Upon heating in water, PNIPAAm undergoes coil-to-globule transition due to breaking hydrogen bonds between NIPAAm units and water and forming intra-chain hydrogen bonds.179–181 While for PNIPAAm, this transition occurs at ≈ 32 °C, copolymerization of NIPAAm with another unit (say, U) shifts the transition point up or down, depending on the physico-chemical nature of U, chemi-cal composition of the resultant copolymer, and comonomer distribution of NIPAAm and U.

Works pertaining to studying the formation and properties of protein-like behavior included copolymers made by copoly-merizing NIPAAm with hydrophilic monomers, including, methacrylic acid (MAA),182 acrylic acid (AA),183 and sodium styrene sulfate (NaSS).184 Inter- and intramolecular association was observed that was governed by the interplay between the solubility of the hydrophilic units and temperature-dependent solubility of NIPAAm. In most instances, the monomer sequence distribution in the resultant copolymer was not mea-sured; the presence of blockiness was inferred indirectly.

Wu185,186Zhang and performed an extensive series of experiments aiming at comprehending the phase behavior of RCPs bearing a NIPAAm unit. For instance, they noted that copolymerization of NIPAAm with hydrophilic VP at tempera-tures higher and lower than its LCST, respectively, resulted in segmented and random VP distributions in the resultant P(NIPAAm-co-VP) copolymers.187 The authors reported that P(NIPAAm-co-VP) chains with a random-blocky distribution of the VP units aggregated readily to form larger mesoglobules relative to their counterparts having a random VP distribution. The formation of mesoglobules was attributed to the competi-tion between intrachain contraction and interchain association.

In another study, the researchers copolymerized NIPAAm and EO at 45 °C, that is, above the LCST of NIPAAm.188 Wu and coworkers used a palette of experimental tools including light scattering and differential scanning calorimetry and observed that the copolymer with a higher PEO content involved two transitions; one sharp transition at 33 °C and another broader one around 45 °C (both in heating and cool-ing processes). The second transition disappeared gradually with lowering the PEO content in the copolymer. The

researchers attributed the first transition to the collapse of the PNIPAAm segments. The second, higher-temperature transi-tion, was associated with steric repulsion-induced stretching and the n-clustering-induced collapse of PEO chains on the periphery of the collapsed PNIPAAm cores.

Cheng et al.189 also reported on phase behavior of NIPAAm and sodium acrylate (NaA) copolymers grafted on solid sub-strates. The researchers noted that the brushes can refold to the state at which the chains were prepared, a behavior that was consistent with the theoretical prediction for the folding of PLCs. They also commented on the formation of hydrogen bonds in the collapsed state of the copolymer chains, indicat-ing that the hysteresis in the cooling process was due to the incomplete removal of additional hydrogen bonds.

Tang et al.190 reported on the formation of ‘macromolecular thermometers’ by direct copolymerization of NIPAAm and a tetraphenylene (TPE)-based unit; TPE is nonemissive when dissolved but becomes highly emissive when aggregated. While no direct characterization of the NIPAAm and TPE sequences was shown, the authors noted that increasing the content of the TPE units resulted in polymer aggregation due to hydrophobic interactions among TPE units, as deduced from increased fluorescence signals. While increasing the solution temperature collapses NIPAAm units, it leads to breaking up TPE complexes due to increased molecular motion that ulti-mately leads to decrease of fluorescence signal. One can only speculate that variation of comonomers sequences in P(NIPAAm-co-TPE) can be observed by monitoring the fluor-escence emission as a function of temperature.

Similarly, Liu et al.191 prepared graft copolymers comprising poly(acrylic acid) (PAA) backbone and side-chain PNIPAAm grafts that exhibited pH-responsive (due to PAA) and thermo-responsive (due to PNIPAAm) behavior. The macromo-lecules were formed by AIBN-initiated direct copolymerization in solution. No effect of comonomer sequences on the phase behavior was reported, although the system would, in principle, allow for probing this effect efficiently.

Shimori and coworkers studied the formation of a copolymer of a RCP P(NIPAAm-co-AAPBA) made by direct copolymeriza-tion of NIPAAm and 3-(acrylamido)phenylboronic acid (AAPBA) and its interaction with glucose. It was reported that the temperature-induced precipitation of P(NIPAAm-co-AAPBA) increased by 1.5 and 4 °C in the presence of a small concentra-tion of glucose and fructose, respectively.192 While the authors did not comment on the effect of comonomer sequences in their paper, it is obvious that tailoring the distribution of the como-nomer units in the copolymer will have a profound effect on the bind efficacy to sugars.

Multiblock copolymers comprising NIPAAm and N,N-dimethylacrylamide (DMAAm) with variable length of blocks of both components prepared by reversible addition– fragmentation chain transfer polymerization were also investi-gated and their phase behavior was found to exhibit ‘protein-like behavior’ exhibiting intramolecular collapse upon heating, forming unimolecular flower-like micelles above the thermal phase transition temperature.193

6.18.3.3 Polymer-Analogous Reactions

A few groups have developed experimental protocols for gen-erating PLCs that follow very closely the scheme suggested by

Page 29: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy716 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

early computer simulations of Khokhlov and Khalatur.69–71

Starting from a parent homopolymer made of chemically equivalent segments, PLCs can be generated by chemically modifying those segments that are accessible for reaction. In the initial stages of the reaction, the process is governed by the conformation of the parent macromolecule; as the reaction proceeds and the solubility of the newly created copolymer changes, the conformation of the copolymer and thus accessi-bility of the modifiable segments get altered, which, in turn, affects the accessibility of the reaction sites.90,101

Manokruang and Manias194 formed RCP comprising methyl methacrylate (MMA) and MAA units by alkaline hydro-lysis of parent PMMA homopolymers with NaOH, which converted MMA into MAA. By performing the reaction under poor solvent conditions (isopropyl alcohol with small amounts of deionized water), the researchers speculated that the resulting P(MMA-co-MAA) copolymers possessed blocky distributions of the two monomers. While no attempts were made to characterize the comonomer distribution, the authors reported on sharp phase transitions in aqueous solutions as a function of pH and a corresponding hysteresis associated with increasing and decreasing the pH. The latter was attributed to inter- and intrachain interactions and stabilization of the col-lapsed state due to strong hydrophobic interactions acting among MMA units.

One of the most comprehensive studies pertaining to the formation of PLCs has been carried out by Semler et al.100 They utilized electrophilic substitution of Br onto PS195 in selective solvents as a means of preparing poly(styrene-co-4-bromostyrene) (PBrxS) with tunable distribution of the 4-bromostyrene units whose mole fraction (x) can be tune smoothly from 0 to 1. Jhon et al.196 have studied the kinetics of bromination of PS in selective solvents and established that regardless of the solvent type, the reaction is second order in bromine. The researchers have also carried out a series of experiments examining the role of solvent in the bromination reaction on the comonomer sequence distribution in PBrxS. A series of PBrxS copolymers was prepared by brominating parent PS (Mn = 30 kDa) in 1-chlorodecane (CD), 1-choroundecane (CUD), or 1-chlorododecane (CDD) at various temperatures. Because the solvents have similar dipole moments but varying different theta temperatures (TΘ), (TΘCD = 6.6 °C, TΘCUD = 32.8 °C, TΘCDD = 58.6 °C), the kinetics of bromination is similar at a given temperature for all three solvents.196

Because of the temperature- and solvent-dependent conforma-tional changes of PS, bromination was expected to produce PBrxS with different sequence distributions of 4-BrS. For instance, at 32.8 °C, PS adopts a swollen coil conformation in CD, Gaussian coil in CUD, and a collapsed conformation when dissolved in CDD. In the following, we refer to the samples as PBrxS-SOLtt, where x denotes the mole fraction of 4-BrS, SOL stands for solvent (CD, CUD, or CDD), and the symbol ‘tt’ denotes the bromination temperature in °C.

In order to establish the degree of blockiness of 4-BrS in PBrxS-SOLtt, Semler et al.100 utilized an electro-optical Kerr effect measurement in order to determine the molar Kerr con-stant (mK), which carries information about both the monomer stereoregularity and PS/4-BrS sequencing in PBrxS (cf. Figure 17). To gain a more quantitative insight into the results of the electrical birefringence measurements, the researchers calculated mK using the rotational isomeric state

(RIS) model and matrix multiplication methods of Flory for a few selected periodic sequences having the same degree of bromination (≈60%) and degree of polymerization (=300) as the experimental system. Comparison of the mKs for the RIS periodic sequences and the experimental sequences provided clear evidence that PBr0.58S-CDD33 and PBr0.63S-CUD33 have a more blocky character relative to PBr0.59S-CD33; the average block length of the 4-BrS units within PBr0.58S-CDD33 and PBr0.63S-CUD33 is in the range of 20 and 10 monomers, respectively.

The differences in monomer sequence distribution influ-ence significantly the partitioning of copolymers at interfaces. Specifically, macromolecules with longer consecutive sequences of modified monomers that interact preferentially with the substrate adhere more strongly to the substrate relative to the polymers with random distribution of the modified units. Therefore, the mobility of RCPs on substrates, as reflected by the rate of dewetting, is also expected to decrease with increasing degree of blockiness of the modified monomers.

To confirm the differences in monomer sequence distribu-tions between PBr0.59S-CD33 and PBr0.58S-CDD33, Genzer and coworkers studied dewetting of thin films of those poly-mers from solid substrates. The effect of the monomer sequence distribution in PBrxS was elucidated by depositing films of the same thickness onto substrates covered with semi-fluorinated organosilane (SFOS) self-assembled monolayers (SAMs) and annealing the specimens. The growth and coales-cence of holes in the films of various thicknesses were recorded using optical microscopy as a function of time revealed that the polymer with a higher degree of blockiness dewetted more slowly relative to the specimen having a ‘truly random’ mono-mer sequence distribution. The sample set up and the results of the experimental findings are summarized in Figure 18.

Jhon et al.197 studied the effect of polymer composition, blockiness, solvent quality, polymer concentration, and tem-perature on adsorption of PBrxS PLCs onto silica substrates. They reported that among the solvents studied, cyclohexane showed the highest PBrxS adsorption to substrates because of high affinity of the 4-BrS unit to silica, strong repulsion of 4-BrS from the solvent, and weak adsorption of cyclohexane on the substrate. Under these conditions, the amount of PBrxS adsorbed onto silica substrates increased with increasing 4-BrS content in the copolymer. Upon increasing the tempera-ture from 25 to 50 °C, the amount of PBrxS on silica decreased because styrene solubility in cyclohexane increased (the system crossed the theta-to-good solvent quality for PS), as shown by the data presented in Figure 19. While at low degrees of bro-mination, Jhon et al. could not detect any dependence of comonomer sequences in PBrxS on adsorption, increasing the degree of bromination to 35% produced adsorption isotherms that depended on the comonomer sequences. Specifically, sam-ples with higher degree of blockiness exhibited higher adsorption on silica relative to PBrxS that possessed a random distribution of PS and 4-BrS units.

Solubility of PBrxS copolymers in cyclohexane was investi-gated in great detail. Specifically, Jhon et al.198 measured the turbidity of the solution as a function of solute concentration and temperature. The cloud point was found to depend on the comonomer distribution; it is always higher for random-blocky samples, relative to truly random comonomer distributions. By performing the experiments at various PBrxS concentrations in

Page 30: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy

1 2 3 4 5 6

* *

* * * *

Br Br Br

3 (b)(a) 0.35 4 PS

5

w e

ndo

(W)

0.30 PBr0.58S-CDD33

0.25 PBr0.59S-CD33 130 125 120

PS

Hea

t flo 0.20

PBr1.0S

0.15 1 3 PS/PBr

2 1.0S0.10

4 (40/60) 5 6 80 100 120 140 160 180

Temperature (�C) 130 125 120

PBr

2 m

) 0.350.58S-CDD33

− (c)

(V

0.30 S-CD33

1 3 142 0.25 PBr0.59

4 10

5 6 �er

r co

nsta

nt 0.20 S-CUD33

PBr0.63

0.15 PBr0.58S-CDD33

130 125 120

0.10 PBr0.59S-CD33 PS

K 0.05 200 160 120 80 40 0 1 2 3 4 5 6

ppm Polymer concentration (wt.%)

Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 717

Figure 17 Characterization of poly(styrene-co-4-bromostyrene) (PBrxS) copolymers, where x is the mole fraction of 4-BrS. (a) 13C-NMR spectra of PS (top, black line), b-PBr0.58S (middle, red line), and r-PBr0.59S (bottom, blue line); the peaks in the spectra mark the resonances of phenyl ring carbons in the ortho, meta, and para positions (shown in the upper panel of the figure) in PS and PBrxSs. The 13C-NMR data confirm that the bromination occurred exclusively in the para position of the phenyl ring of PS, while the aliphatic part of the PS chain and its 13C-NMR spectrum are not affected by the bromination. The mole fraction of 4-bromostyrene in PBrxS can be obtained from the ratios of peaks 1 and 3, 2 and 4, or 6 and 5. (b) Differential scanning calorimetry (DSC) from PS (black line), PBr0.58S-CDD33 (red line), PBr0.59S-CD33 (blue line), PBr1.0S (green line), and a mixture of PS and PBr1.0S (40/60 m/m) (magenta line). The DSC results provide the Tg of each material and also serve as further evidence that the bromination reaction takes place on all PS chains present in the reaction vessel. (c) Kerr constant for PS (black line), PBr0.58S-CDD33 (red line), and PBr0.59S-CD33 (blue line) as measured on solutions (1–5 wt.%) of polymers in 1,4-dioxane (>99% purity) at 293 K.

cyclohexane and various amounts of 4-BrS in the sample, Jhon et al. began generating phase diagrams for solubility of RCPs as a function of comonomer sequences. During the course of these experiments, they have also discovered a strong isotope effect on solubility. These experiments were performed with all possible combinations of deuterium labeling: (1) PBrxS incyclohexane, (2) d-PBrxS in cyclohexane, (3) PBrxS id-cyclohexane, and (4) d-PBrxS in d-cyclohexane. In addition, the transition from coil to globule in random-blocky samples suggested that one may encounter multiple-chain collapse, as supposed to single-chain collapse present in the coil-to-globule transition of random PBrxS.

Recently, Han et al.199 carried out analysis of selective adsorption of PBrxS copolymers using interaction chromato-graphy (IC). Their IC experiments performed with a series of PBrxS samples reveal that (1) IC is capable of determining the extent of bromination in PBrxS samples regardless of the

solvent used, and (2) for some solvents, one can, in addition to measuring the chemical composition of the random PBrxS copolymer, also detect differences in monomer sequence dis-tribution. The data in Figure 20 represent a summary of the IC experiments. As discussed by Han et al., the interplay between solvent-induced conformational changes and surface adsorp-tion on chromatographic stationary phases offers new insights in chromatographic separation for the characterization of the synthetic copolymers and biomacromolecues that embody comonomer sequence distribution.

6.18.4 Concluding Remarks

Under the general heading of ‘sequence design’, there is an ever-increasing body of methodology that has been rapidly evolving over the past few years in polymer chemistry. In this

n

Page 31: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy718 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

Figure 18 (a) Samples for dewetting studies are formed by covering the silica substrate with a self-assembled monolayer of 1H,1H,2H,2H-perfluorodecyltrichlorosilane (SF-SAM), casting thin polymer films (thickness t ), and heating to temperature T such that T–Tg = 35 °C, where Tg is the bulk glass transition temperature of the polymer. Upon annealing, holes (diameter D) develop in the films and grow with increasing time. The time dependence of the hole diameters for PS (squares), PBr0.59S-CD33 (circles), and PBr0.58S-CDD33 (triangles) measured on films having a thickness of t ≈ 53–67 nm (b) and t ≈ 117–118 nm (c) are presented. The lines are meant to guide the eye.

2.5

2.0

(a) (b) (c)

) 2−m

Ads

orbe

d am

ount

(m

g

1.5

1.0

0.5 T = 25 �C T = 50 �C T = 50 �C

PS PS PS r-PBr0.10S r-PBr0.10S r-PBr0.35S

b-PBr S b-PBr S b-PBr S 0.0 0.10 0.10 0.35

0 20 40 60 80 0 20 40 60 80 0 20 40 60 80 Adsorption time (h)

Figure 19 Adsorbed amount of polystyrene (open squares), r-PBrxS (closed squares), and b-PBrxS (closed circles) as a function of adsorption time for polymers adsorbing from cyclohexane at 25 °C (a) and 50 °C (b, c). The mole fraction of 4-BrS units in PBrxS was 0.10 (a, b) and 0.35 (c). The polymer concentration in the solution was 1% (w/w). The lines are meant to guide the eye.

review, we have discussed advances that have recently been achieved in the computer simulation, theoretical understand-ing, and synthesis of designed functional copolymers. The fundamental principle of these approaches was conformational-dependent sequence design (CDSD), which takes into account a strong coupling between the conformation and primary structure of copolymers during their synthesis. The focus was on PLCs with a nontrivial primary sequence

exhibiting large-scale compositional heterogeneities and long-range statistical correlations between monomeric units. These features are intrinsically related to the CDSD scheme and they cannot be explained by the basic stochastic processes such as random sequence or Markov chain; the first has no correlations and the second only has SRCs. We have tried to demonstrate how the preparation conditions dictate the result-ing copolymer sequences. In many cases, the presence of LRCs

deposit polymerfilm

D

Polymer film

anneal aboveTg of polymers

t

Silica substrate Silica substrateSilica substrate

(b)

(a)

(c)

Hol

e di

amet

er (

μm)

120

100

80

60

40

20

00 500 1000

PS (t = 67.6 nm)

PBr0.58S-CDD33

SF-SAM

(t = 52.9 nm)

PBr0.59S-CD33 (t = 54.6 nm)

PS (t = 116.9 nm)

PBr0.58S-CDD33 (t = 118.1 nm)PBr0.59S-CD33 (t = 118.0 nm)

1500

Annealing time (s)

0 500 1000 1500

Page 32: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 719

Figure 20 Difference between the degree of polymerization of the adsorbing segments in System I (PS) and System II (4-BrS) in PBrxS-CUD and PBrxS-NB as a function of the mole fraction of 4-BrS in the PBrxS copolymer. For each system, the data split into two groups depending on the bromination temperature. Bromination of PS below T = 32.8 °C produces PBrxS RCPs with a random-blocky distribution of the 4-BrS segments (b-PBrxS). Bromination of PS above T = 32.8 °C produces PBrxS RCPs with a more random distribution of the 4-BrS segments (r-PBrxS).

in designed copolymers can bring about dramatic changes in their physical properties with respect to the corresponding copolymers whose sequence has only minor correlation between adjacent monomeric units.

While PLCs have been realized experimentally165–199 and their unique physical characteristics have been documented in multiple studies,76,77,200–210 much more remains to be under-stood about these unique macromolecular systems. First and foremost, our understanding of the coloring process in real experimental systems is limited. This stems primarily form technical difficulties associated with monitoring the sequence design (coloring) process with a sufficient spatio-temporal resolution. It is likely that, in order to gain insight into this process, multiple experimental probes would have to be employed concurrently. While some initial attempts have began recently for the PBrxS using small-angle neutron scatter-ing in combination with Kerr effect and elemental analysis,211

much more work needs to be done before one can describe the experimental coloring pathways conclusively.

As discussed earlier in the document, the actual coloring pathway may follow multiple paths depending on the solubi-lity of the parent homopolymer and that of the coloring units. While it appears from the preliminary data that during the formation of PBrxS a conformational rearrangement of the original PS coil takes place, other experimental situations may be realized that follow a different route. For instance, one can envision situations, also discussed earlier, where a compact globule opens up progressively during the coloring chemical reaction because the newly added monomer increases the

overall solubility of the mother macromolecule. An example of such a system can be quarternization of poly(dimethylami-noethyl methacrylate) (PDMAEMA) in water at high temperatures, where PDMAEMA is originally collapsed.212

Another area that merits further work involves detailed investigation of the effect of the molecular weight of the parent homopolymer on the comonomer sequence distribution in PLC. While is it desired that macromolecules with larger mole-cular weight are used in this process because they form a more compact globule and are thus expected to result in PLCs with more pronounced degrees of blockiness, technical limitations exist that set an upper limit on the molecular weight. To explain, in order to perform the coloring reaction under poor solvent conditions, one has to work with sufficiently low poly-mer concentrations (typically much below the overlap concentration, C*) in order to avoid agglomeration of the collapsed homopolymer as well as the newly formed copoly-mer globules. Because C* decreased dramatically with increasing molecular weight of the polymer, one may not produce a sufficient amount of the PLC when working with high-molecular-weight macromolecules. One possible solution to this problem could involve grafting the parent macromole-cules onto a solid substrate in order to decrease their tendency to overlap with their neighbors. This situation can be realized technically by growing the parent polymers by ‘grafting from’ methodology from supports comprising inorganic or metallic (nano)particles decorated with polymerization initiators. The density of the initiators should be sufficiently low that the growing chains remain in the ‘mushroom conformation’ and

15

10

5

0

−5

−10

−15

−200.25 0.30 0.35

Mole fraction of 4-BrS in PBrxS

0.40 0.45 0.50

MC/ACN on C18-SiOx

THF/IO on SiOxb-PBrxS

b-PBrxS

r-PBrxS

r-PBrxS

‘truly’ random copolymer

T > Θ T = Θ T < ΘD

Pad

s-D

Pad

seq

Page 33: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy720 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

are thus isolated. Sufficient amounts of polymer can be pro-duced by employing small particles whose overall surface area-to-volume is high. Using small particles also provides the advantages of minimizing any confinement effect both during the ‘grafting from’ homopolymerization as well as the coloring reaction itself. Alternatively, end-functionalized polymers can be grafted chemically by the ‘grafting onto’ methodology from solution; this technique has the advantage of avoiding any confinement effects during the ‘grafting from’ technique as well as the ability to work with polymers that have been char-acterized previously in bulk. One has to stress, however, that avoiding confinement effect completely during the coloring process is likely not possible because the collapsed macromo-lecule may (and likely will) interact strongly with the particle support.

Finally, adequate analytical methods should be employed in order to characterize the structure of the copolymer. Characterization of the overall molecular weight and chemical composition of the copolymer is determined by utilizing well-established analytical methodologies. However, determi-nation of comonomer sequence distribution is more challenging. While in some situations one can use nuclear magnetic resonance (NMR), information thus obtained is lim-ited because NMR cannot distinguish sequences beyond the length of triads. As recently demonstrated by Semler et al.,100

Kerr effect, that is, measurement of optical birefringence of solutions under electric field, may provide a convenient means of determining comonomer sequence distribution (in addition to chemical composition and tacticity) when there is sufficient contrast between the dipole moments of the mono-mers present. Kerr effect does not suffer from the same limitation of NMR.213 However, in order to interpret the experimental data by complementing the measurements with modeling using RIS model,214,215 and calculating the so-called molar Kerr constant by employing the matrix formalism devel-oped by Flory216 and others.217–222

It should be noted that PLCs prepared by either of the aforementioned methodology are not monodisperse in their length, chemical composition, and the comonomer distribu-tion. The degree of polydispersity in any properties varies from method to method and even within each method, it may change depending on the actual system synthesized. Characterizing these polydispersities is a daunting task. Even more complex is fractionation of the resultant copolymers into monodisperse fractions.

References

1. Bates, F. S.; Fredrickson, G. H. Phys. Today 1999, 52, 32. 2. Hamley, I. W. The Physics of Block Copolymers; Oxford University Press: New

York, NY, 1998. 3. Hamley, I. W. Nanotechnology 2003, 14, R39. 4. Hamley, I. W., Ed. Developments in Block Copolymer Science and Technology;

Wiley: New York, NY, 2004, p 380. 5. Stoykovich, M. P.; Nealey, P. F. Mater. Today 2006, 9, 20. 6. Blumenfeld, L. A.; Tikhonov, A. N. Biophysical Thermodynamics of Intracellular

Processes: Molecular Machines of the Living Cell; Springer Verlag: Berlin, 1994. 7. Li, H.; Helling, R.; Tang, C.; Wingreen, N. Science 1996, 273, 666. 8. Xia, Y.; Levitt, M. Curr. Opin. Struct. Biol. 2004, 14, 202. 9. Balazs, A. C.; Gempe, M. Macromolecules 1991, 24, 167–176. 10. Chai, Z. K.; Sun, R. N.; Karasz, F. E. Macromolecules 1992, 25, 6113–6118.

11. Angerman, H.; Hadziioannou, G.; ten Brinke, G. Phys. Rev. E 1994, 50, 3808–3813.

12. Simmons, E. R.; Chakraborty, A. K. J. Chem. Phys. 1998, 109, 5493–5496. 13. Faldi, A.; Genzer, J.; Composto, R. J.; Dozier, W. D. Phys. Rev. Lett. 1995,

74, 3388. 14. Genzer, J.; Composto, R. J. Macromolecules 1998, 31, 870. 15. Pellegrini, N. N.; Sikka, M.; Satija, S. K.; Winey, K. I. Polymer 2000,

41, 2701–2704. 16. Kulasekere, R.; Kaiser, H.; Anker, J. F.; et al. Macromolecules 1996,

29, 5493–5496. 17. Mansky, P.; Liu, Y.; Huang, E.; et al. Science 1997, 275, 1458–1460. 18. Benkoski, J. J.; Fredrickson, G. H.; Kramer, E. J. J. Polym. Sci., Part B: Polym.

Phys. 2001, 39, 2363–2377. 19. Eastwood, E. A.; Dadmun, M. D. Macromolecules 2002, 35, 5069–5077. 20. Montanari, A.; Müller, M.; Mezard, M. Phys. Rev. Lett. 2004, 92, 185509. 21. Moghaddam, M. S.; Chan, H. S. J. Chem. Phys. 2006, 125, 164909. 22. Bratko, D.; Chakraborty, A. K.; Shakhnovich, E. I. Chem. Phys. Lett. 1997,

280, 46–52. 23. Chakraborty, A. K.; Golumbfskie, A. J. Annu. Rev. Phys. Chem. 2001,

52, 537–573. 24. Chakraborty, A. K. Phys. Rep. 2001, 342, 2–61. 25. Greszta, D.; Madrare, D.; Matyjaszewski, K. Macromolecules 1994, 27, 638. 26. Moad, G.; Solomon, D. H. The Chemistry of Free Radical Polymerization; Elsevier:

Oxford, 1995. 27. Malmstrom, E. E.; Hawker, C. J. Macromol. Chem. Phys. 1998, 199, 923. 28. Clay, P.; Christie, D. I.; Gilbert, R. G. In Controlled Radical Polymerization;

Matyjaszewski, K., Ed.; ACS Symposium Series 685; American Chemical Society: Washington, DC, 1998, Chapter 7, pp 104–119.

29. Matyjaszewski, K.; Xia, J. Chem. Rev. 2001, 101, 2921. 30. Davis, K. A.; Matyjaszewski, K. Adv. Polym. Sci. 2002, 159, 14. 31. Davis, K. A.; Matyjaszewski, K. Statistical, Gradient, Block and Graft Copolymers

by Controlled/Living Radical Polymerizations; Springer: Berlin; New York, 2002. 32. Hsieh, H. L.; Quirk, R. P. Anionic Polymerization: Principles and Practical

Applications; Marcel Dekker: New York, NY, 1996. 33. Hillmyer, M. Curr. Opin. Solid State Mater. Sci. 1999, 4, 559. 34. Qiu, J.; Charleux, B.; Matyjazsewski, K. Prog. Polym. Sci. 2001, 26, 2083. 35. Ando, T.; Kato, M.; Kamigaito, M.; Sawamoto, M. Macromolecules 1996,

29, 1070. 36. Matyjaszewski, K.; Patten, T. E.; Xia, J. H. J. Am. Chem. Soc. 1997, 119, 674. 37. Shakhnovich, E. I. Fold. Des. 1996, 1, R50. 38. Shakhnovich, E. I. Curr. Opin. Struct. Biol. 1997, 7, 29. 39. Shakhnovich, E. I. Fold. Des. 1998, 3, 45. 40. Frauenfelder, H.; Wolynes, P. G.; Austin, R. H. Rev. Mod. Phys. 1999, 71, S419. 41. Pande, V. S.; Grosberg, A. Y.; Tanaka, T. Rev. Mod. Phys. 2000, 72, 259. 42. Shakhnovich, E. Chem. Rev. 2006, 106, 1559–1588. 43. Dill, K. A. Biochemistry 1990, 29, 7133–7155. 44. West, M. W.; Hecht, M. H. Protein Sci. 1995, 4, 2032–2039. 45. Kamtekar, S.; Schiffer, J. M.; Xiong, H.; et al. Science 1993, 262, 1680–1685. 46. Shakhnovich, E. I.; Gutin, A. M. Proc. Natl. Acad. Sci. U.S.A. 1993, 90, 7195. 47. Pande, V. S.; Grosberg, A. Y.; Tanaka, T. J. Phys. Fr. 1994, 4, 1771. 48. Pande, V. S.; Grosberg, A. Y.; Tanaka, T. Proc. Natl. Acad. Sci. U.S.A. 1994,

91, 12972. 49. Irbäck, A.; Peterson, C.; Potthast, F.; Sandelin, E. Phys. Rev. E 1998, 58, R5249. 50. Gupta, P.; Hall, C. K.; Voegler, A. C. Protein Sci. 1998, 7, 2642. 51. Lau, K. F.; Dill, K. A. Macromolecules 1989, 22, 3986–3997. 52. Sali, A.; Shakhnovich, E.; Karplus, M. Nature 1994, 369, 248–251. 53. Dill, K. A. Biochemistry 1985, 24, 1501–1509. 54. Ashbaugh, H. S.; Hatch, H. W. J. Am. Chem. Soc. 2008, 130, 9536–9542. 55. Konig, J. L. Chemical Microstructure of Polymer Chains; Wiley: New York,

NY, 1980. 56. Li, W.; Kaneko, K. Europhys. Lett. 1992, 17, 655. 57. Peng, C.-K.; Buldyrev, S. V.; Goldberger, A. L.; et al. Nature (London) 1992,

356, 168. 58. Peng, C.-K.; Buldyrev, S. V.; Havlin, S.; et al. Phys. Rev. E 1994, 49, 1685. 59. Pande, V. S.; Grosberg, A. Y.; Tanaka, T. J. Chem. Phys. 1994, 101, 8246. 60. Irback, A.; Peterson, C.; Potthast, F. Proc. Natl. Acad. Sci. U.S.A. 1996, 93, 9533. 61. Gusev, L. V.; Vasilevskaya, V. V.; Makeev, V. Y.; et al. Macromol. Theory Simul.

2003, 12, 604. 62. Gatlin, L. L. Information Theory and the Living System; Columbia University Press:

New York, NY, 1972. 63. Nicolis, G.; Prigogine, I. Exploring Complexity; W. H. Freeman: New York,

NY, 1989. 64. Kauffman, S. A. Origins of Order: Self-organization and Selection in Evolution;

Oxford University Press: New York, NY, 1993.

Page 34: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 721

65. Shannon, C. E. Bell Syst. Tech. J. 1951, 30, 50–64. 66. Lin, J. IEEE Trans. Inf. Theory 1991, 37, 145–151. 67. Ofran, Y.; Rost, B. J. Mol. Biol. 2003, 325, 377–387. 68. Khalatur, P. G.; Novikov, V. V.; Khokhlov, A. R. Phys. Rev. E 2003, 67, 51901. 69. Khokhlov, A. R.; Khalatur, P. G. Physica A 1998, 249, 253. 70. Khalatur, P. G.; Ivanov, V. I.; Shusharina, N. P.; Khokhlov, A. R. Russ. Chem. Bull.

1998, 47, 855. 71. Khokhlov, A. R.; Khalatur, P. G. Phys. Rev. Lett. 1999, 82, 3456. 72. Khokhlov, A. R.; Grosberg, A. Y.; Khalatur, P. G.; et al. In Proceedings of the

International School of Physics “Enrico Fermi”, Course CXLV: “Protein Folding, Evolution and Design”; Broglia, R. A., Shakhnovich, E. I., Eds.; IOS Press: Amsterdam, 2001; pp 313–332.

73. Khokhlov, A. R.; Khalatur, P. G.; Ivanov, V. A.; et al. In Challenges in Molecular Simulations (SIMU Newsletter); Mac Kernan, D., Ed.; CECAM: Lyon, 2002; Vol. 4, pp 79–100.

74. Khokhlov, A. R.; Khalatur, P. G. Curr. Opin. Solid State Mater. Sci. 2004, 8, 3. 75. Khokhlov, A. R.; Berezkin, A. V.; Khalatur, P. G. J. Polym. Sci., Part A: Polym.

Chem. 2004, 42, 5339. 76. Khalatur, P. G.; Berezkin, A. V.; Khokhlov, A. R. Recent Res. Dev. Chem. Phys.

2004, 5, 339. 77. Khalatur, P. G.; Khokhlov, A. R. Adv. Polym. Sci. 2006, 195, 1–100. 78. Staudinger, H. The Nobel Prize in Chemistry, 1953, Nobelprize.org. 21 Sep 2011;

http://www.nobelprize.org/nobel_prizes/chemistry/laureates/1953/. 79. Staudinger, H.; Machemer, H. Ber. Dtsch. Chem. Ges. 1926, 59, 2932. 80. Boaen, N.; Hillmyer, M. Chem. Soc. Rev. 2005, 34, 267. 81. Gauthier, M. A.; Gibson, M. I.; Klok, H.-A. Angew. Chem. 2009, 48, 48. 82. Schultz, G. E.; Schirmer, R. H. Principles of Protein Structure; Springer:

Berlin, 1979. 83. Carmesin, I.; Kremer, K. Macromolecules 1988, 21, 2819. 84. Govorun, E. N.; Ivanov, V. A.; Khokhlov, A. R.; et al. Phys. Rev. E 2001,

64, 40903. 85. Shlesinger, M. F.; Zaslavskii, G. M.; Frisch, U. Lévy Flights and Related Topics in

Physics; Springer-Verlag: Berlin, 1996. 86. Kuchanov, S. I.; Khokhlov, A. R. J. Chem. Phys. 2003, 118, 4684. 87. Yu, Z. G.; Anh, V. V.; Lau, K. S. Phys. Rev. E 2003, 68, 21913. 88. Irbäck, A.; Peterson, C.; Potthast, F. Proc. Natl. Acad. Sci. U.S.A. 1996, 93, 9533. 89. Yu, Z. G.; Anh, V. V.; Lau, K. S. Phys. Rev. E 2001, 64, 31903. 90. Semler, J. J.; Genzer, J. J. Chem. Phys. 2006, 125, 14902. 91. Zhang, J. X.; Ma, P. X. Angew. Chem. Int. Ed. 2009, 48, 964. 92. Yow, H. N.; Routh, A. F. Soft Matter 2006, 2, 940. 93. Tiarks, F.; Landfester, K.; Antonietti, M. Langmuir 2001, 17, 908. 94. Dowding, P. J.; Atkin, R.; Vincent, B.; Bouillot, P. Langmuir 2004, 20, 11374. 95. Cai, C. H.; Lin, J. P.; Chen, T.; et al. Chem. Commun. 2009, 2709. 96. Dai, T.; Chen, K.; Qing, X.; et al. Macromol. Rapid Commun. 2010, 31, 484–489. 97. Kuchanov, S. I. Adv. Polym. Sci. 2000, 152, 157. 98. Berezkin, A. V.; Khalatur, P. G.; Khokhlov, A. R. Polym. Sci. A 2005, 47, 66. 99. Wesson, L.; Eisenberg, D. Protein Sci. 1992, 1, 227. 100. Semler, J. J.; Jhon, Y. K.; Tonelli, A.; et al. Adv. Mater. 2007, 19, 2877. 101. Strickland, L.; Hall, C. K.; Genzer, J. Macromolecules 2009, 42, 9063. 102. Strickland, L. A.; Hall, C. K.; Genzer, J. Langmuir 2010, 26, 8810–8820. 103. Brantley, E. L.; Jennings, G. K. Macromolecules 2004, 37, 1476. 104. Bantz, M. R.; Brantley, E. L.; Weinstein, R. D.; et al. J. Phys. Chem. B 2004,

108, 9787. 105. Brantley, E. L.; Holmes, T. C.; Jennings, G. K. J. Phys. Chem. B 2004,

108, 16077. 106. Jennings, G. K.; Brantley, E. L. Adv. Mater. 2004, 16, 1983. 107. Brantley, E. L.; Holmes, T. C.; Jennings, G. K. Macromolecules 2005, 38, 9730. 108. Arifuzzaman, S.; Özcam, A. E.; Efimenko, K.; et al. Biointerphases 2009, 4, FA33. 109. Schmitt, J.; Decher, G.; Dressick, W. J.; et al. Adv. Mater. 1997, 9, 61. 110. Currie, E. P. K.; van der Gucht, J.; Borisov, O. V.; Stuart, M. A. C. Langmuir 1998,

14, 5740. 111. Currie, E. P. K.; Fleer, G. J.; Stuart, M. A. C.; Borisov, O. V. Eur. Phys. J. E 2000,

1, 27. 112. Liu, Z.; Pappacena, K.; Cerise, J.; et al. Nano Lett. 2002, 2, 219. 113. Bhat, R. R.; Genzer, J.; Chaney, B. N.; et al. Nanotechnology 2003, 14, 1145. 114. Currie, E. P. K.; Norde, W.; Stuart, C. M. A. Adv. Colloid Interface Sci. 2003,

100–102, 205. 115. Bhat, R. R.; Genzer, J. Appl. Surf. Sci. 2005, 252, 2549. 116. Luzinov, I.; Minko, S.; Tsukruk, V. V. Soft Matter 2008, 4, 714. 117. Diamanti, S.; Arifuzzaman, S.; Elsen, A.; et al. Polymer 2008, 49, 3770. 118. Diamanti, S.; Arifuzzaman, S.; Genzer, J.; Vaia, R. A. ACS Nano 2009, 3, 807. 119. Kim, J. U.; O’Shaughnessy, B. Phys. Rev. Lett. 2002, 89, 238301. 120. Zheligovskaya, E. A.; Khalatur, P. G.; Khokhlov, A. R. Phys. Rev. E 1999,

59, 3071.

121. Starovoitova, N. Y.; Khalatur, P. G.; Khokhlov, A. R. In Forces, Growth and Form in Soft Condensed Matter: At the Interface between Physics and Biology; Skjeltorp, A. T., Belushkin, A. V. Eds.; NATO Science Series II: Mathematics, Physics and Chemistry; Kluwer: Dordrecht, 2004; Vol. 160, 253 pp.

122. Zhulina, E. B.; Skvortsov, A. M.; Birshtein, T. M. Vysokomol. Soed. A 1981, 23, 304.

123. Velichko, Y. S.; Khalatur, P. G.; Khokhlov, A. R. Macromolecules 2003, 36, 5047. 124. Thurmond, K. B.; Kowalewski, T.; Wooley, K. L. J. Am. Chem. Soc. 1997,

119, 6656. 125. Huang, H.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. J. Am. Chem. Soc. 1999,

121, 3805. 126. Sanji, T.; Nakatsuka, Y.; Ohnishi, S.; Sakurai, H. Macromolecules 2000,

33, 8524. 127. Stewart, S.; Liu, G. Chem. Mater. 1999, 11, 1048. 128. Discher, B. M.; Won, Y.-Y.; Ege, D. S.; et al. Science 1999, 284, 1143. 129. Nardin, C.; Hirt, T.; Leukel, J.; Meier, W. Langmuir 2000, 16, 1035. 130. Trefonas, P.; West, R.; Miller, R. D. J. Am. Chem. Soc. 1985, 107, 2737. 131. Wullf, G.; Sarhan, A. Angew. Chem. Int. Ed. 1972, 11, 341. 132. Polowinski, S. Template Polymerization; ChemTec Publishing: Toronto-

Scarborough, 1997. 133. de Gennes, P.-G. Scaling Concepts in Polymer Physics; Cornell University Press:

Ithaca, NY, 1979. 134. Berezkin, A. V.; Khalatur, P. G.; Khokhlov, A. R. J. Chem. Phys. 2003, 118, 8049. 135. Berezkin, A. V.; Khalatur, P. G.; Khokhlov, A. R.; Reineker, P. New J. Phys. 2004,

6, 44. 136. Flory, P. J. Principles of Polymer Chemistry; Cornell University Press: Ithaca, NY,

1953. 137. Dotson, N. A.; Galvan, R.; Laurence, R. L.; Tirrell, M. Polymerization Process

Modeling; VCH Publishers Inc.: New York, NY, 1995. 138. Starovoitova, N. Y.; Khalatur, P. G.; Khokhlov, A. R. Dokl. Chem. 2003, 392, 242. 139. Starovoitova, N. Y.; Khalatur, P. G.; Khokhlov, A. R. In Forces, Growth and Form in

Soft Condensed Matter: At the Interface between Physics and Biology; Skjeltorp, A. T., Belushkin, A. V., Eds.; NATO Science Series II: Mathematics, Physics and Chemistry; Kluwer: Dordrecht, 2004; Vol. 160, 253 pp.

140. Starovoitova, N. Y.; Berezkin, A. V.; Kriksin, Y. A.; et al. Macromolecules 2005, 38, 2419.

141. Berezkin, A. V.; Khalatur, P. G.; Khokhlov, A. R. Macromolecules 2006, 39, 8808–8815.

142. Aksimentiev, A.; Holyst, R. J. Chem. Phys. 1999, 111, 2329. 143. Lefebvre, M. D.; Cruz, M. O.; Shull, K. R. Macromolecules 2004, 37, 1118. 144. Pakula, T.; Matyjaszewski, K. Macromol. Theory Simul. 1996, 5, 987. 145. Shull, K. R. Macromolecules 2002, 35, 8631. 146. Pickett, G. T. J. Chem. Phys. 2003, 118, 3898. 147. Matyjaszewski, K.; Ziegler, M. J.; Arehart, S. V.; et al. J. Phys. Org. Chem. 2000,

13, 775. 148. Matyjaszewski, K.; Xia, J. Chem. Rev. 2001, 101, 2921. 149. Lutz, J.-F.; Pakula, T.; Matyjaszewski, K. Advances in Controlled/Living Radical

Polymerization; American Chemical Society: Washington, DC, 2003. 150. Dettmer, C. M.; Gray, M. K.; Torkelson, J. M.; Nguyen, S. T. Macromolecules

2004, 37, 5504. 151. Berezkin, A. V.; Solov’ev, M. A.; Khalatur, P. G.; Khokhlov, A. R. J. Chem. Phys.

2004, 121, 6011. 152. Berezkin, A. V.; Solov’ev, M. A.; Khalatur, P. G.; Khokhlov, A. R. Polym. Sci. A

2005, 47, 622. 153. Smith, J. M. On Evolution; Edinburgh University Press: Edinburgh, 1972. 154. Volkenstein, M. V. General Biophysics; Academic Press: New York, NY, 1983. 155. Leninger, A. L.; Nelson, D. L.; Cox, M. M. Principles of Biochemistry, 2nd ed.;

Worth Publishers: New York, NY, 1993. 156. Grosberg, A. Y.; Khokhlov, A. R. Giant Molecules: Here and There and Everywhere;

Academic Press: New York, NY, 1997. 157. Chertovich, A. V.; Govorun, E. N.; Ivanov, V. A.; et al. Eur. Phys. J. E 2004, 13, 15. 158. Grosberg, A. Y. Biofizika 1984, 29, 569. 159. Leibler, L. Macromolecules 1980, 13, 1602. 160. Fredrickson, G. H. The Equilibrium Theory of Inhomogeneous Polymers; Oxford

University Press: New York, NY, 2006. 161. Erukhimovich, I. Ya. In Nanostructured Soft Matter. Experiment, Theory,

Simulation and Perspectives; Zvelindovsky, A. V., Ed.; Springer: Dordrecht, 2007; pp 327–370.

162. Khalatur, P. G.; Khokhlov, A. R.; Krotova, M. K. Macromol. Symp. 2007, 252, 36–46.

163. Chandler, D.; Andersen, H. C. J. Chem. Phys. 1972, 57, 1930. 164. Schweizer, K. S.; Curro, J. G. Adv. Chem. Phys. 1997, 98, 1. 165. Virtanen, J.; Tenhu, H. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 3716. 166. Virtanen, J.; Baron, C.; Tenhu, H. Macromolecules 2000, 33, 336.

Page 35: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copy722 Shape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers

167. Virtanen, J.; Tenhu, H. Macromolecules 2000, 33, 5970. 168. Virtanen, J.; Lemmetyinen, H.; Tenhu, H. Polymer 2001, 42, 9487. 169. Lozinsky, V. I.; Simenel, I. A.; Kurskaya, E. A.; et al. Dokl. Chem. 2000, 375, 273. 170. Lozinsky, V. I.; Simenel, I. A.; Kulakova, V. K.; et al. Macromolecules 2003,

36, 7308. 171. Lozinsky, V. I.; Simenel, I. A.; Semenova, M. G.; et al. Polym. Sci. A 2006,

48, 435. 172. Lozinskii, V. I.; Simenel, I. A.; Khokhlov, A. R. Dokl. Chem. 2006, 410, 170. 173. Siu, M.; Zhang, G.; Wu, C. Macromolecules 2002, 35, 2723. 174. Wahlund, P.-O.; Galaev, I. Yu.; Kazakov, S. A.; et al. Macromol. Biosci. 2002,

2, 33. 175. Kumar, M.; Kamihira, M.; Galaev, I. Yu.; et al. Langmuir 2003, 19, 865. 176. Lozinsky, V. I. Adv. Polym. Sci. 2006, 196, 87. 177. Laukkanen, A.; Winnik, F. M.; Tenhu, H. Macromolecules 2005, 38, 2439. 178. Dai, S.; Ravi, P.; Tam, K. C. Soft Matter 2009, 5, 2513. 179. Sun, B.; Lin, Y.; Wu, P. Appl. Spectrosc. 2007, 61, 765. 180. Rzaev, Z. M. O.; Pskin, S. D. E. Prog. Polym. Sci. 2007, 32, 534. 181. Zhou, K.; Lu, Y.; Li, J.; et al. Macromolecules 2008, 41, 8927. 182. Burova, T. V.; Grinberg, N. V.; Grinberg, V. Ya.; et al. Macromolecules 2005,

38, 1292. 183. Weng, Y.; Ding, Y.; Zhang, G. J. Phys. Chem. B 2006, 110, 11813. 184. Burova, T. V.; Grinberg, N. V.; Grinberg, V. Ya.; et al. Macromolecules 2008,

41, 5981. 185. Zhang, G.; Wu, C. Adv. Polym. Sci. 2006, 195, 101. 186. Zhang, Q.; Ye, J.; Lu, Y.; et al. Macromolecules 2008, 41, 2228. 187. Siu, M.; He, C.; Wu, C. Macromolecules 2003, 36, 6588. 188. Chen, H.; Li, J.; Ding, Y.; et al. Macromolecules 2005, 38, 4403. 189. Cheng, H.; Liu, G.; Wang, C.; et al. J. Polym. Sci., Part B: Polym. Phys. 2006,

44, 770. 190. Tang, L.; Jin, J. K.; Qin, A.; et al. Chem. Commun. 2009, 4974. 191. Liu, S.; Liu, X.; Li, F.; et al. J. Appl. Polym. Sci. 2008, 109, 4036. 192. Shiomori, K.; Ivanov, A. E.; Galaev, I. Y.; et al. Macromol. Chem. Phys. 2004,

205, 27. 193. Zhou, Y.; Jiang, K.; Song, Q.; Liu, S. Langmuir 2007, 23, 13076. 194. Manokruang, K.; Manias, E. Mater. Lett. 2009, 63, 1144–1147. 195. Kambour, R.; Bendler, J.; Bopp, R. Macromolecules 1983, 16, 753. 196. Jhon, Y. K.; Semler, J. J.; Genzer, J. Macromolecules 2008, 41, 6719.

197. Jhon, J. K.; Semler, J. J.; Genzer, J.; et al. Macromolecules 2009, 42, 2843. 198. Jhon, Y.K.; Krishnamoorti, R.; Genzer, J. 2010. 199. Han, J.; Jeon, B. H.; Ryu, C. Y.; et al. Macromol. Rapid Commun. 2009, 30, 1543. 200. Kriksin, Y. A.; Khalatur, P. G.; Khokhlov, A. R. Macromol. Theory Simul. 2002,

11, 213. 201. Khalatur, P. G.; Khokhlov, A. R.; Mologin, D. A.; Reineker, P. J. Chem. Phys.

2003, 119, 1232. 202. Govorun, E. N.; Khokhlov, A. R.; Semenov, A. N. Eur. Phys. J. E 2003, 12, 255. 203. Semenov, A. N. Macromolecules 2004, 31, 226. 204. Vasilevskaya, V. V.; Klochkov, A. A.; Lazutin, A. A.; et al. Macromolecules 2004,

37, 5444. 205. Mologin, D. A.; Khalatur, P. G.; Khokhlov, A. R.; Reineker, P. New J. Phys. 2004,

6, 133. 206. Khokhlov, A. R.; Khalatur, P. G. Curr. Opin. Colloid Interface Sci. 2005,

10, 22. 207. Manning, G. S. Macromolecules 2007, 40, 8071–8081. 208. Yoshinaga, N.; Bicout, D. J.; Kats, E. I.; Halperin, A. Macromolecules 2007,

40, 2201–2209. 209. Ashbaugh, H. S. J. Phys. Chem. B 2009, 113, 14043–14046. 210. Malik, R.; Hall, C. K.; Genzer, J. Macromolecules 2010, 43, 5149–5157. 211. Genzer, J.; John, Y.K.; Krishnamoorti, R. 212. Galvin, C.J.; Genzer, J. 213. Tonelli, A. E. Macromolecules 2009, 42, 3830. 214. Wattice, W. L.; Suter, U. W. Conformational Theory of Large Molecules. The

Rotational Isomeric State Model in Macromolecular Systems; Wiley: New York, NY, 1994.

215. Rehahn, M.; Mattice, W. L.; Suter, U. W. Adv. Polym. Sci. 1997, 131/132, 1, and references therein.

216. Flory, P. J. Statistical Mechanics of Chain Molecules; Wiley: New York, NY, 1969. 217. Tonelli, A. E. Macromolecules 1977, 10, 153. 218. Kelly, K. M.; Patterson, G. D.; Tonelli, A. E. Macromolecules 1977, 10, 859. 219. Khanarian, G.; Tonelli, A. E. Macromolecules 1982, 15, 145. 220. Khanarian, G.; Cais, R. E.; Kometani, J.; Tonelli, A. E. Macromolecules 1982,

15, 866. 221. Khanarian, G.; Schilling, F. C.; Cais, R. E.; Tonelli, A. E. Macromolecules 1982,

15, 876. 222. Tonelli, A. E.; Khanarian, G.; Cais, R. E. Macromolecules 1985, 18, 2324.

Page 36: Provided for non-commercial research and educational use. Not … · 2012-08-26 · 6.18.3.2 Conformation-Dependent Synthesis: Direct Copolymerization 714 6.18.3.3 Polymer-Analogous

Author's personal copyShape-Controlled Polymers and Nano-Objects | Conformation-Dependent Design of Synthetic Functional Copolymers 723

Biographical Sketches

Jan Genzer received his ‘Diploma-engineer’ degree (Dipl.-Ing.) in Chemical and Materials Engineering from the Institute of Chemical Technology in Prague, Czech Republic, in 1989. In 1991, he moved to the United States to pursue graduate studies at the University of Pennsylvania under the direction of Professor Russ Composto, receiving his PhD degree in Materials Science and Engineering in 1996. After two postdoctoral stints with Professor Ed Kramer first at Cornell University (1996–97) and later at University of California at Santa Barbara (1997–98), Genzer joined the faculty of chemical engineering at the NC State University as an assistant professor in fall 1998. He is currently the Celanese Professor of Chemical and Biomolecular Engineering at NC State University. His honors include Camille Dreyfus Teacher-Scholar Award, National Science Foundation CAREER award, John H. Dillon Award of the American Physical Society, National Science Foundation Award for Special Creativity, and others. He is a Fellow of the American Physical Society. His group at NC State University is actively involved in research related to the behavior of polymers at interfaces and in confined geometries, with particular emphasis on self-assembly and forced assembly and combinatorial methods.

Pavel G. Khalatur received a Ph.D. degree in chemistry from the Tver State University and a Dr.Sc. degree in physics and mathematics from the Institute of Macromolecular Compounds of the Russian Academy of Sciences. In 1990, he became a full professor of the Department of Physical Chemistry of Tver State University. In 2002, he moved to the University of Ulm to pursue his research with A. R. Khokhlov, dedicated to computer-aided molecular design of bioinspired polymers. He has been awarded four Distinguished University Professor Awards from the International Science Foundation for excellence in science and teaching. Prof. Khalatur has authored and coauthored over 300 articles and four books in the field of polymer science and advised more than twenty Ph.D. candidates. The major foci of his research are computer modeling of polymer systems, computer-assisted molecular design, nanoscience and the integral equation theory of simple and polymeric liquids.

Alexei R. Khokhlov is the head of a scientific school, and the author of textbooks widely known both in Russia and abroad. The range of his interests covers polymer science, statistical physics of macromolecules, physical chemistry of polyelec-trolytes and ionomers, microphase separation in polymer systems, polymer liquid crystals, polyelectrolyte responsive gels, topological restrictions in polymer systems, dynamics of concentrated polymer solutions and melts, coil–globule transi-tions, associating polymers, computer simulation of polymer systems, biomimetic polymers, and proton-conducting polymer membranes. His publications include works on both theoretical and experimental science, some of the former being classical in the field. Among others, he has proposed methods to synthesize ‘protein-like’ copolymer sequences as well as macromolecules whose functional properties can be preset. The theoretical principles he developed have been successfully implemented in the field of oil extraction, in the production of durable biocompatible polymers and new efficient catalysts to synthesize vitamins.