IV. Ionomer Phenomena

Similar documents
Material Chemistry KJM 3100/4100. Synthetic Polymers (e.g., Polystyrene, Poly(vinyl chloride), Poly(ethylene oxide))

`1AP Biology Study Guide Chapter 2 v Atomic structure is the basis of life s chemistry Ø Living and non- living things are composed of atoms Ø

Colloidal dispersion

Lecture 5: Macromolecules, polymers and DNA

Electrostatic Self-assembly : A New Route Towards Nanostructures

Lecture 8. Polymers and Gels

Chapter 7. Pickering Stabilisation ABSTRACT

Colloidal Suspension Rheology Chapter 1 Study Questions

Module 4: "Surface Thermodynamics" Lecture 22: "" The Lecture Contains: Examples on Effect of surfactant on interfacial tension. Objectives_template

*blood and bones contain colloids. *milk is a good example of a colloidal dispersion.

Monolayers. Factors affecting the adsorption from solution. Adsorption of amphiphilic molecules on solid support

Aqueous solutions. Solubility of different compounds in water

Lecture 8 Polymers and Gels

Intermolecular forces

Lec.1 Chemistry Of Water

Applied Surfactants: Principles and Applications

LAYER BY LAYER (LbL) SELF-ASSEMBLY STRATEGY AND ITS APPLICATIONS

Miami Dade College CHM Second Semester General Chemistry

Chap. 2. Polymers Introduction. - Polymers: synthetic materials <--> natural materials

Chapter 2 - Water 9/8/2014. Water exists as a H-bonded network with an average of 4 H-bonds per molecule in ice and 3.4 in liquid. 104.

2. Amorphous or Crystalline Structurally, polymers in the solid state may be amorphous or crystalline. When polymers are cooled from the molten state

Model Solutions Spring 2003

Uniform properties throughout! SOLUTE(S) - component(s) of a solution present in small amounts.

Layer-by-Layer (LBL) Self-Assembly

16 years ago TODAY (9/11) at 8:46, the first tower was hit at 9:03, the second tower was hit. Lecture 2 (9/11/17)

One Q partial negative, the other partial negative Ø H- bonding particularly strong. Abby Carroll 2

Amorphous Polymers: Polymer Conformation Laboratory 1: Module 1

Physical Chemistry of Polymers (4)

CHEM J-9 June 2012

Liquid Crystal. Liquid Crystal. Liquid Crystal Polymers. Liquid Crystal. Orientation of molecules in the mesophase

The Chemistry and Energy of Life

Stability of colloidal systems

Enduring Understandings & Essential Knowledge for AP Chemistry

Chemistry: A Molecular Approach, 1 st Ed. Nivaldo Tro

EXAM I COURSE TFY4310 MOLECULAR BIOPHYSICS December Suggested resolution

INTERMOLECULAR AND SURFACE FORCES

6 Hydrophobic interactions

Model Solutions Spring 2003

Chapter 6 Chemistry of Water; Chemistry in Water

Chapter 10: Liquids and Solids

EFFECTS OF ADDED ELECTROLYTES ON THE STRUCTURE OF CHARGED POLYMERIC MICELLES

Liquid Chromatography

COLLOIDAL SOLUTIONS. Department of Medical Chemistry Pomeranian Medical University

Properties of Solutions

Name Biology Chapter 2 Note-taking worksheet

Colloid stability. Lyophobic sols. Stabilization of colloids.

Protein separation and characterization

Ionic polysaccharides, solubility, interactions with surfactants, particle formation, and deposition

Due in class on Thursday Sept. 8 th

Solids, liquids and gases

- intermolecular forces forces that exist between molecules

Life is a chemical process

Microbiology with Diseases by Taxonomy, 5e (Bauman) Chapter 2 The Chemistry of Microbiology. 2.1 Multiple Choice Questions

Anionic Polymerization - Initiation and Propagation

Lecture Presentation. Chapter 12. Solutions. Sherril Soman, Grand Valley State University Pearson Education, Inc.

A dispersion (system) Colloidal solutions High molecular mass compounds

Chapter 11. Liquids and Intermolecular Forces

Solutions and Non-Covalent Binding Forces

1.8 Thermodynamics. N Goalby chemrevise.org. Definitions of enthalpy changes

Chapter 02 The Chemical Basis of Life I: Atoms, Molecules, and Water

Swelling and Collapse of Single Polymer Molecules and Gels.

Foundations in Microbiology Seventh Edition

2 Structure. 2.1 Coulomb interactions

Chapter 002 The Chemistry of Biology

Full file at Chapter 2 Water: The Solvent for Biochemical Reactions

Hydrogel thermodynamics (continued) Physical hydrogels

12A Entropy. Entropy change ( S) N Goalby chemrevise.org 1. System and Surroundings

Chapter 10 Liquids, Solids, and Intermolecular Forces

Proteins polymer molecules, folded in complex structures. Konstantin Popov Department of Biochemistry and Biophysics

UNIT 1: BIOCHEMISTRY

Unit 10: Part 1: Polarity and Intermolecular Forces

Chapters 11 and 12: Intermolecular Forces of Liquids and Solids

Intermolecular forces Liquids and Solids

2) Matter composed of a single type of atom is known as a(n) 2) A) element. B) mineral. C) electron. D) compound. E) molecule.

Polymers in Modified Asphalt Robert Q. Kluttz KRATON Polymers

BIOLOGY 101. CHAPTER 3: Water and Life: The Molecule that supports all Live

CHAPTER 10. Characteristics of the Surfaces of Biomaterials

Chapter 13 States of Matter Forces of Attraction 13.3 Liquids and Solids 13.4 Phase Changes

SPECIALTY MONOMERS FOR ENHANCED FUNCTIONALITY IN EMULSION POLYMERIZATION

Chapter 9 Generation of (Nano)Particles by Growth

General Chemistry A


Solutions and Intermolecular Forces

Biophysics II. Hydrophobic Bio-molecules. Key points to be covered. Molecular Interactions in Bio-molecular Structures - van der Waals Interaction

Chapter 11 Properties of Solutions

CH1810 Lecture #1 Solutions of Ionic Compounds

General Chemistry A

Paper No. 1: ORGANIC CHEMISTRY- I (Nature of Bonding and Stereochemistry)

Full file at

Lecture 4. Donnan Potential

Tunable Nanoparticle Arrays at Charged Interfaces

H 2 O WHAT PROPERTIES OF WATER MAKE IT ESSENTIAL TO LIFE OF EARTH? Good solvent High Surface tension Low vapor pressure High boiling point

CHEMISTRY PHYSICAL. of FOODS INTRODUCTION TO THE. CRC Press. Translated by Jonathan Rhoades. Taylor & Francis Croup

Water. 2.1 Weak Interactions in Aqueous Sy stems Ionization of Water, Weak Acids, and Weak Bases 58

Self-assembly Structures of Block Copolymers in Selective Solvents and of Polysaccharide- Surfactant Mixtures

Liquids and Solids. H fus (Heat of fusion) H vap (Heat of vaporization) H sub (Heat of sublimation)

TECHNOLOGIES THAT TRANSFORM POLLUTANTS TO INNOCUOUS COMPONENTS: CHEMICAL AND PHYSICOCHEMICAL METHODS

It is the size of the

BIOCHEMISTRY GUIDED NOTES - AP BIOLOGY-

Supplementary Information for Blocky Sulfonation of Syndiotactic Polystyrene: A Facile Route Toward Tailored Ionomer

Transcription:

IV. Ionomer Phenomena (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998) The modulus, glass transition temperature, viscosity, melt strength, fatigue, and barrier properties are all strongly affected by the presence of ionizable units in ionomers. Multiplets effectively create physical crosslinks that reduce the mobility of polymer chains, causing these properties changes. Glass Transition Temperature The next figure shows the storage modulus and tanδ for random styrene-co-sodium methacrylate copolymer ionomers of different ion content (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998). These plots show a striking influence of ion content on thermal properties, with the modulus raised near and above the glass transition temperature T g, which itself rises from ~125ºC to ~325ºC, a rather dramatic change given than all the ion contents are less than 21.6%. While many other factors may be important, counterion content, degree of neutralization, and counterion identity all affect T g dramatically. Universal trends are hard to extract. The next figure shows the T g for ethyl acylate-co-methacrylic acid ionomers vs. methacrylic acid content, the ionizable units present in the acid (H + ) form as well as 1

neutralized by monovalent (Cs + ) or divalent cation (Ca +2 ). Trends with the three counterions follow the expected order, indicating that T g rises with the strength of attraction between ionizable unit and counterion (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998). Hypothesizing that shifts of T g reflect the strength and number of these attractions, the same data are replotted in the next figure as a function of cz/a, where c is ion content (mol %), z is the counterion valence, and a is the counterion size (nm). A rather remarkable superposition of data occurs(eisenberg and Kim, Introduction to Ionomers, Wiley, 1998). 2

In contrast, when the same plot is made for sulfonated styrene ionomers, no superposition occurs. Instead, T g correlates well with sulfonation level (c) and not with counterion valence (z) or size (a). Reasons for the difference are unclear. One possibility is that limited dissociation of ionizable units provides the segmental mobility needed at T g for ethyl acylate-comethacrylic acid ionomers while similar dissociation is not associated with T g for sulfonated styrene ionomers. Consistent with this reasoning, the latter has much lower ε. As expected, increasing the degree of neutralization also increases T g, although the effect is weaker than the one illustrated in the previous figures. At higher ion contents, there is evidence for a second T g at higher T. A second T g would indicate sample heterogeneity, perhaps the clustering or microphase separation of multiplets. The weakly sigmoidal nature of the collapsed T g curve for ethyl acylate-comethacrylic acid ionomers may reflect a shift from sensitivity to one T g to the other T g at high values of cz/a. Mechanical Properties Stress-strain curves of model urethane-based ionomers with pendent sulfonate groups are displayed next. (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998) 3

These polyurethanes lack a chain extender, and so crosslinking relies entirely upon association of undissociated ionizable groups into multiplets. The figure shows a significant rise of modulus with neutralization accompanied by little loss of strain at break; these are very tough elastomeric materials, with toughness improved with increase of neutralization. Somewhat as the sulfonated styrene ionomers of the last section, T g for these urethane ionomers is not sensitive to ion type or content. However, stress-strain curves for the material are quite sensitive to the same variables. The next figure shows the influence of ion type on stress-strain curves for a similar (not identical) set of urethane elastmers. (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998) Finally, since ionizable units of an ionomer act as crosslinks, they dramatically increase the plateau modulus, here shown for the same type of polyurethane ionomers (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998) 4

Miscibility Because of the strength of electrostatic interactions, the usual state of incompatibility between two chemically distinct polymers can often be evaded in ionomers. Although strong attractive electrostatic interactions tip the balance toward miscibility in many cases, and provide a homogeneous material at relatively low ion content, these same blends may present processing difficult due to high viscosity and/or irreproducible (kinetically controlled) mixing. Mixing may be achieved in solution, with subsequent solvent removal, or by compounding of the two polymer solids. Electrostatic interactions enhancing miscibility are illustrated in the following sketches: (Eisenberg and Kim, Introduction to Ionomers, Wiley, 1998) 5

V. Electrostatic Complexation Complexation of charged polymers forms the most rapidly growing research endeavor of the charged polymer field, so only the key features of a few specific examples can be mentioned. Fundamental concepts are just emerging, and there remain many mysteries. The compatibility of ionomer blends, discussed briefly in the preceding section, offers a first insight into this sort of complexation. And complexation is a much bigger issue in the context of polyelectrolytes than of ionomers. In general, complexation involves multiple, strong, and localized charge-charge, charge-dipole, or dipole-dipole interactions among charge polymers and/or entities possessing complementary electrostatic functionalities. Because electrostatic interactions are rather insensitive to functional group chemistry, specific interactions are the exception rather than the norm. In much on-going research, electrostatic interactions of charged polymers are being tailored to guide the formation of specific complex structures. Organized, rather than disordered, structures are the goal. Complexes spontaneously form when dissolved polyelectrolytes are added to systems containing certain multivalent counterions, organic dyes with ionic groups, charged surfactants, oppositely charged polyelectrolytes, charged dendrimers, and charged colloids. In addition, there are many examples of polyelectrolytes forming complexes with species such as neutral polymers, neutral surfactants, monovalent counterions, as well as proteins, viruses and other biological structures. Cooperative interactions among chain segments, enabled by the long range of electrostatic interactions, distinguish polyelectrolyte complexes from the complexes formed by low molecular weight compounds. As polyelectrolyte molecular weight decreases, cooperativity is lost, and eventually, complexation itself may disappear. Much inspiration for study of polyelectrolyte complexation is found in the intricate polyelectrolyte complexes found in nature. Shown next is a sequences of images extracted from a movie posted on the web (www.wni.edu.au) that animates the complexation of negatively charged ds-dna with positively charged proteins (histones), to form first nucleosomal particles (appearing at about the third image) and then higher order complex structures. The crude sketch just below the image sequence shows an idealization of a string of nucleosomal core particles, each of which consists of a short cylinder of 8 self-assembled histones around which is wrapped ~2 turns of ds-dna. 6

Need for such a DNA packaging scheme in eukaryotic cells is clear once the random coil size of DNA is compared to the size of the cell nucleus. As a random coil, ds-dna is orders-of-magnitude too large to fit inside. Replicating this sort of assembly with synthetic polyelectrolytes is the long-range goal. Fundamental Physics of Complexation Direct charge-charge interactions between polyelectrolyte host and oppositely charged guest, interactions that provide a negative enthalpy change, favor polyelectrolyte complexation, but the stronger driving force for complexation is the entropy gain associated with the release of small counterions. Each release of a previously condensed counterion generates an entropy gain of approximately kt. For this reason host-guest complexes characterized by 1:1 charge ratio vastly predominate. Many of these stoichiometric complexes are effectively insoluble, although at extremely low c, solubility may be recovered. Nonstoichiometric complexes, on the other hand, are usually soluble or at least dispersible as charged aggregates or assemblies. Trade-off between enthalpy and entropy has been quantified by Muthukumar in simulations complexation between positive and negative polyelectrolytes, as shown below. 7

Enthalpy Free Energy Entropy While ξ and c s play key roles in the final properties of soluble complexes, their greatest impacts are felt during complex formation and dissociation. Newly formed complexes are rarely ordered, and if mobilities of components are low, will remain disordered forever. There seems to be no universal approach for describing kinetic processes within a complex or predicting whether order will eventually appear. Complexes with smaller, less charged, and more mobile species, such as counterions, organic dyes, and surfactants, more often organize into well-defined structures, while polyelectrolytepolyelectrolyte complexes usually possess a permanent scrambled egg morphology. Complexation at higher concentrations often leads to structures reminiscent of thermoreversible gels. Such systems can disrupted or dissociated by altering parameters such as c s so as to weaken electrostatic interactions. Polyelectrolyte-Surfactant Complexes 0.0 0.1 1.0 10.0 ξ When charged surfactants are added to a solution of oppositely charged polyelectrolyte, the surfactant molecules begin to form micelles at a well-defined surfactant concentration that can be several orders-of-magnitude below the usual critical micelle concentration (CMC); this new concentration threshold is usually termed the critical aggregation concentration (CAC). Micelles formed under this condition bind to the polyelectrolyte chains to form soluble complexes that have sometimes been described as possessing the necklace of beads morphology; in other instances, spherical or cylindrical surfactant micelles apparently cluster into large, dense aggregates surrounded or interpenetrated by polyelectrolyte. 8

These polyelectrolye-bound surfactant micelles are smaller than the free micelles formed above the CMC in the absence of polyelectrolyte. Details concerning the local organization of soluble polyelectrolyte-surfactant complexes remain uncertain, but some have speculated that the polyelectrolyte chain stabilizes such smaller micelles by adsorbing to their surfaces. At surfactant concentrations below the CMC, stabilization of polyelectrolyte-surfactant complexes comes from both electrostatic and hydrophobic interactions. As the surfactant concentration increases, reaching a level at which enough micelles bind on the polyelectrolyte to raise the complex stoichiometry to approximately 1:1, solubility is lost and an ordered polymer-micelle precipitate appear. The unique solid-state structures so-formed are being intensely investigated for applications ranging from non-wetting coatings to drug delivery. Only as more surfactant is added to the phase separated mixture do free surfactant micelles eventually begin to appear. In some instances, even greater levels of added surfactant can resolubilize the precipitated complex. All of these behaviors depend strongly on c s. The following sketch shows the evolution of complex morphology for the system cetytrimethylammonium chloride poly(styrene sulfonate) as the bulk charge ratio [s]/[p] (surfactant-to-polymer charge of entire system) rises from well below unity to well above unity. This sequence was determined by x-ray and neutron scattering in collaboration with Strey and Nause. Some of the x-ray patterns are displayed lower on the page. Soluble 40-48Å 46-50Å Spherical Micelles Cylindrical Micelles 120-180Å ~ 46Å Insoluble ~ Increasing surfactantto-polymer charge ratio HCP-C Pm3n Cubic 9

Pm3n cubic hexagonal Polyelectrolyte-Polyelectrolyte Complexes Polyelectrolytes of opposite charge tend to form insoluble, stoichiometric (1:1 charge ratio) complexes (PECs) when mixed together at bulk 1:1 charge ratio, especially if the mixing is performed under salt-free conditions at polymer concentrations that are neither extremely low nor high. Complex formation is fast and nearly molecular weight independent beyond the oligomer range. Solutions of soluble, nonstoichiometric complexes can be formed in two ways, by mixing oppositely charged polyelectrolytes of highly different molecular weights at nonstoichiometric bulk charge mixing ratio (ratio>10:1 in favor of the high M polymer) or by mixing oppositely charged polyelectrolytes at very low polymer concentrations. As opposed to polyelectrolyte-surfactant complexes, polyelectrolyte-polyelectrolyte complexes have little if any order, a feature probably reflecting the low mobility of their constituent chains and the strength of mutual attractions. The driving forces for complexation is principally counterion release into the polyelectrolyte-depleted solution phase, a release that raises the entropy of the complexed system. Depending on levels of added electrolyte, existing polyelectrolyte-polyelectrolyte complexes can undergo swelling, flocculation, and dissolution. 10

At low and moderate concentrations, complexation by titration of one component into the second follows a two-step process, as shown in the following figure. Much below 1:1 bulk charge stoichiometry, complexes form primary aggregates, which contain 10-1000 chains stabilized electrostatically by an excess of the majority polyelectrolyte at their surfaces. Approaching or exceeding 1:1 bulk charge stoichiometry, the primary aggregates phase separate (flocculate) and then sediment. When a weakly charged polyelectrolyte is mixed with a strongly or weakly charged polyelectrolyte, the mixture may sometimes phase separate into two liquid phases varying strongly in polyelectrolyte concentration. Such phase separation is called coacervation ; the denser of the two phases often has the properties of a weak physical gel. Coacervation is often observed when a charged globular protein solution is mixed with a solution of an oppositely charged polyelecrolyte of high charge density: the polyelectrolyte chains are presumed to wrap around the proteins, which effectively act as weak physical crosslinks since their overall surface charge is so low. Changing ph so as to weaken the polyelectrolyte binding to protein reversible changes system association. Two variants of bulk polyelectrolyte complexation, one old and one new, have been reported to create ordered structures. In the old approach, termed matrix polymerization, a template polyelectrolyte is mixed with a monomer of opposite charge, which is then radically polymerized in presence of the template. The polymerized polyelectrolyte, if its charge density is comparable to the template polyelectrolyte and polymerization is rapid, has a chain length nearly identical to the template. The resulting insoluble complex has well-organized local order as inferred by x-ray diffraction. In the 11

new approach, a block polyelectrolyte, i.e., a chain with blocks of neutral and ionizable units, is mixed with a polyelectrolyte of opposite charge. The polyelectrolyte complex microphase separates from the neutral component upon complexation. This approach has been reported to produce nearly monodisperse nanoparticles dispersed in water when the neutral sequence is comprised of hydrophilic poly(ethylene oxide), which forms a stabilizing brush on the nanoparticle surface. Layer-by-Layer Deposition The layer-by-layer method capitalizes on overcompensation of surface charge when polyelectrolytes are exposed to an oppositely charged surface. Thus, a bare surface with inherently negative charge, such as a silicon oxide in contact with water, will adsorb sufficient cationic polyelectrolyte for the surface to become positively charged. This positive charge, in turn, can be overcompensated by subsequent adsorption of an anionic polyelectrolyte. Pursuing this dipping process through alternating, sequential exposures of surface to solutions of anionic and cationic polyelectrolyte builds a film possibly containing 100s or even 1000s of layers, as first discovered by Decher in 1991. This process is electrostatically analogous to the formation of PECs in a bulk solution, except for the constraint of a flat growth surface and time-separated addition of complexing components. These small differences, however, dramatically change the level of order that can be achieved. As shown below, film thickness grows with layer number linearly after a brief nonlinear onset of ~5 layers. In this onset regime, the bare substrate charge affects the amount of polymer deposited, but subsequently, the substrate itself has no effect on the amount deposited. [Schlenoff and Dubas, Macromolecules 34 (2001) 592] 12

There are numerous factors affecting the layer growth process, and indeed, the method has recently been extended to nonelectrostatic growth mechanisms such as hydrogen bonding. An enormous number of charged components have been applied to this technique, including many synthetic polyelectrolytes, biopolymers (proteins, polysaccharides, nucleic acids), clays, silica, viruses, and dendrimers. The same charged components do not have to been used in each dipping cycle throughout multilayer assembly one can start with one negatively charged polyelectrolyte, for example, and then switch to another later in assembly. Although often sketched as discrete alternating layers of two charged components, many experiments have shown that two components are interpenetrated. Flexibly polyelectrolytes extend across multiple adjacent layers. The mass of an individual layer can exceed several monolayer coverages of polyelectrolyte. While most studies have been performed on flat substrates, the method has been successfully employed on spherical colloids, although the achievable number of layers and their perfection may be lesser. Desirable properties of the final films included high uniformity, good stability in aqueous environments, and precise control of final thickness. Among the electrostatic issues of importance to the layer-by-layer method include (i) control of layer thickness; (ii) presence (or not) of counterions; (iii) influence of the polymer ionizable units and added salt on layer growth and properties; and (iv) cause of the charge overcompensation needed to make the growth method work. The following plot shows that c s has a huge effect on the amount of polyelectrolyte deposited, a key feature that an electrostatic model must explain. [Dubas and Schlenoff, Macromolecules 32 (1999) 8153] 13

Several investigators have confirmed that counterions are not present in the bulk of the multilayer assembly; there is an intrinsic 1:1 compensation of polyelectrolyte charges except near the surface, where counterions are present to compensate the excess surface charge of the last deposited polyelectrolyte layer. Lack of counterions in the bulk of the film mandates that the two polymers be fully ionpaired with each other. If so, stratification as discrete layers of one polyelectrolyte after another is impossible, as at each height above the surface, there must be equal numbers of charged segments from the two polyelectrolytes to achieve charge balance. The most important electrostatics issue for layer-by-layer assembly is surface charge overcompensation, which turns out to be a subtle issue. One feature of this overcompensation has already been revealed by the persistence of the onset regime over ~5 layers the surface charge seen by an adsorbing polyelectrolyte has contributions from approximately this number of previous layers. Naively, one might have instead expected that only the topmost polyelectrolyte controlled the surface charge of the assembly. Several methods have revealed that the topmost polyelectrolyte layer is fuzzy with many loops and tails. At high enough c s, small ions effectively can compete for polyelectrolyte charges, disrupting polyelectrolyte-polyelectrolyte ion-pairing. When this happens, the assembly is disrupted. Values of c s for total disruption are typically greater than ~ 1 M. When one of the polyelectrolytes has long neutral sequences (e.g., it is a polyelectrolyte block copolymer), after or during assembly this neutral component can microphase separate, creating well-defined laterally ordered domains within the film. According to a growth model by Schlenoff, the key to charge overcompensation is the presence of electrolyte ions in solution that can penetrate and interfere with polyelectrolyte-polyelectrolyte ion-pairing in ~5 layers nearest the assembly surface. These layers are therefore swollen with solvent, allowing interpenetration of polyelectrolyte from the overlying solution. This bath polyelectrolyte is first captured electrostatically near the surface, after which it is never released. However, the mobilizing effect of added electrolyte allows segments of the captured polyelectrolyte to gradually burrow down into the assembly, the reason for substantial spreading of chains across layers. Whether it is thermodynamic or kinetic factors that eventually limit burrowing and the number of attached chains is unclear; At high c s, the topmost region is more swollen with solvent, explaining why more polyelectrolyte can add and layer increments are thicker. 14