The lattice model of polymer solutions

Similar documents
Rubber elasticity. Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge. February 21, 2009

Lattice protein models

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

University of Washington Department of Chemistry Chemistry 452/456 Summer Quarter 2014

Chemistry 2000 Lecture 11: Chemical equilibrium

PHASE TRANSITIONS IN SOFT MATTER SYSTEMS

V(φ) CH 3 CH 2 CH 2 CH 3. High energy states. Low energy states. Views along the C2-C3 bond

Physical Chemistry of Polymers (4)

Crystalline Solids have atoms arranged in an orderly repeating pattern. Amorphous Solids lack the order found in crystalline solids

Solids / Crystal Structure

3.091 Introduction to Solid State Chemistry. Lecture Notes No. 9a BONDING AND SOLUTIONS

+ kt φ P N lnφ + φ lnφ

Chapter 10. Lesson Starter. Why did you not smell the odor of the vapor immediately? Explain this event in terms of the motion of molecules.

Chapter 4 Polymer solutions

Gases and States of Matter: Unit 8

CH.8 Polymers: Solutions, Blends, Membranes, and Gels

Lecture 8. Polymers and Gels

Liquids, Solids and Phase Changes

Effect of crystallinity on properties. Melting temperature. Melting temperature. Melting temperature. Why?

CH 2 = CH - CH =CH 2

2.1 Traditional and modern applications of polymers. Soft and light materials good heat and electrical insulators

Physical Science Exam 3 Study Guide. Dr. Karoline Rostamiani. Chapter 3

Solutions. Solutions. How Does a Solution Form? Solutions. Energy Changes in Solution. How Does a Solution Form

Thermodynamic of polymer blends Assoc.Prof.Dr. Jatyuphorn Wootthikanokkhan

Critical Temperature - the temperature above which the liquid state of a substance no longer exists regardless of the pressure.

Chapter 13. Properties of Solutions. Lecture Presentation. John D. Bookstaver St. Charles Community College Cottleville, MO

Liquids and Solids Chapter 10

- intermolecular forces forces that exist between molecules

Chapter 8. Chapter 8. Preview. Bellringer. Chapter 8. Particles of Matter. Objectives. Chapter 8. Particles of Matter, continued

Intermolecular Forces and States of Matter AP Chemistry Lecture Outline

Random walks. Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge. March 18, 2009

Chemistry 2000 Lecture 9: Entropy and the second law of thermodynamics

Chapter 19 Chemical Thermodynamics Entropy and free energy

Chapter 17. Free Energy and Thermodynamics. Chapter 17 Lecture Lecture Presentation. Sherril Soman Grand Valley State University

Chemical Engineering 160/260 Polymer Science and Engineering. Lecture 14: Amorphous State February 14, 2001

Lecture 8 Polymers and Gels

Liquids and Solutions

Entropy, free energy and equilibrium. Spontaneity Entropy Free energy and equilibrium

Colligative properties CH102 General Chemistry, Spring 2011, Boston University

2. As gas P increases and/or T is lowered, intermolecular forces become significant, and deviations from ideal gas laws occur (van der Waal equation).

Matter: Properties & Change

Olle Inganäs: Polymers structure and dynamics. Polymer physics

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

Chemistry C : Polymers Section. Dr. Edie Sevick, Research School of Chemistry, ANU. 5.0 Thermodynamics of Polymer Solutions

P a g e What is the algebraic sign for enthalpy of solution? A. positive B. negative C. not enough information is given

Physical Biochemistry. Kwan Hee Lee, Ph.D. Handong Global University

2. As gas P increases and/or T is lowered, intermolecular forces become significant, and deviations from ideal gas laws occur (van der Waal equation).

They are similar to each other. Intermolecular forces

Chapter 12: Solutions. Mrs. Brayfield

Liquids and Solutions Crib Sheet

compared to gases. They are incompressible. Their density doesn t change with temperature. These similarities are due

Chapter 2: INTERMOLECULAR BONDING (4rd session)

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

CHEM. Ch. 12 Notes ~ STATES OF MATTER

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

VII. Porous Media Lecture 32: Percolation

CH.7 Fugacities in Liquid Mixtures: Models and Theories of Solutions

Swelling and Collapse of Single Polymer Molecules and Gels.

Kinetic Theory (Kinetikos - Moving ) Based on the idea that particles of matter are always in motion

Intermolecular Forces: Liquids, and Solids. Chapter 11

1051-3rd Chem Exam_ (A)

CHAPTER 10. States of Matter

CHAPTER 10. Kinetic Molecular Theory. Five Assumptions of the KMT. Atmospheric Pressure

1051-3rd Chem Exam_ (B)

1051-3rd Chem Exam_ (C)

Chapter 13 Properties of Solutions

What is the volume of the unit cell of Ni in ml?

Chapter 12 Intermolecular Forces and Liquids

Chapter 19 Chemical Thermodynamics Entropy and free energy

Chapter 13 Properties of Solutions

States of Matter; Liquids and Solids. Condensation - change of a gas to either the solid or liquid state

Classification of Matter. States of Matter Physical and Chemical Properties Physical and Chemical Changes

Chapter 19. Chemical Thermodynamics

2. Derive ideal mixing and the Flory-Huggins models from the van der Waals mixture partition function.

Chapter 19. Chemical Thermodynamics

Liquids are collections of particles that are held together but they can flow (intermolecular forces)

Chapter 13 Properties of Solutions

School of Chemical & Biological Engineering, Konkuk University

Phase Diagrams: Conditions for Equilibrium (CfE)

Chapter 3. Preview. Section 1 Three States of Matter. Section 2 Behavior of Gases. Section 3 Changes of State. States of Matter.

Intermolecular Forces, Liquids, Solids. IM Forces and Physical Properties

Chapter 19. Chemical Thermodynamics

Chemistry, The Central Science, 11th edition Theodore L. Brown, H. Eugene LeMay, Jr., and Bruce E. Bursten Chapter 13 Properties of Solutions

Chapter 10. Liquids and Solids

Solutions. Experiment 11. Various Types of Solutions. Solution: A homogenous mixture consisting of ions or molecules

Chapter 19. Chemical Thermodynamics

Lattice Theories for Polymer/Small-Molecule Mixtures and the Conformational Entropy Description of the Glass Transition Temperature

CHAPTER 11: INTERMOLECULAR FORCES AND LIQUIDS AND SOLIDS. Chemistry 1411 Joanna Sabey

Chem 30A. Ch 12. Liquids, Solids, and Intermolecular Forces

Lab Week 4 Module α 1. Polymer chains as entropy springs: Rubber stretching experiment and XRD study. Instructor: Gretchen DeVries

They are similar to each other

POGIL: Principles of Solubility

Chemistry 2000 Lecture 12: Temperature dependence of the equilibrium constant

Week 11/Th: Lecture Units 28 & 29

CHEMISTRY Topic #2: Thermochemistry and Electrochemistry What Makes Reactions Go? Fall 2018 Dr. Susan Findlay See Exercises in Topic 8

Department of Chemistry University of Texas at Austin

Lecture 1 Chapter 12 Sections 1-2. Solutions Solubility

Copyright 2018 Dan Dill 1

Chapter 12. Insert picture from First page of chapter. Intermolecular Forces and the Physical Properties of Liquids and Solids

Rationale: Phase diagrams are standard in all high school chemistry textbooks and therefore are considered prior knowledge.

Transcription:

The lattice model of polymer solutions Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge February 25, 2009 1 The lattice model of polymer solutions In the last note, we looked at the lattice model of small-molecule solutions. We now turn to polymers, using the same machinery. The lattice theory of polymer solutions is known as Flory-Huggins theory. In many ways, you will find this theory similar to the smallsolute case, except that the statistics are a little more complicated given that a polymer is a connected entity. Figure 1 shows a lattice model of a polymer. Each filled circle represents a chain segment, a piece of the polymer that is about the same size as a solvent molecule. A chain segment is not necessarily a monomer. In fact, for many polymeric materials, a chain segment will be smaller than a monomer. A polymer is taken to be made up of ν s chain segments (on average). The value of ν s can be estimated as the ratio of the molar volumes of the polymer and solvent. As in the small-molecule case, the solvent is assumed to have a coordination number z. Recall again that the flat illustration of figure 1 is only an aid to visualization and that, in general, the lattice will be three-dimensional, and may have a non-cubic geometry (e.g. it could be a hexagonal close-packed lattice). Neglecting the ends of a chain, which, for long polymers, represent a tiny minority of segments, a linear polymer segment makes z 2 contacts with lattice sites to which the segment is not bonded. Another key assumption in the present note is that all the monomers are the same. There is therefore no energetic difference between placing one monomer or another in contact with the solvent. That being the case, there are only two significant possibilities to consider: Either the polymer is well solvated, i.e. it assumes a more-or-less random conformation in the solvent, or it doesn t dissolve at all. In the next lecture, we will talk about a class of polymers for which this is not true, namely proteins. 1

Figure 1: Lattice model of a polymer. The open circles represent solvent molecules. The filled circles connected by bonds represent the polymer. Each filled circle is a chain segment, a piece of the polymer (not necessarily a monomer) whose size is about the same as that of a solvent molecule. 2 Enthalpy of solution The derivation of the enthalpy of solution for a polymer is extremely similar to that for a small-molecule solution. There is an energy difference of ɛ = ɛ 12 1 2 (ɛ 11 + ɛ 22 ) per solute-solvent (i.e. polymer-solvent) contact made. The derivation of this equation is about the same as before, except that we have to consider one polymer segment and its neighbors rather than a small solute molecule. We now ask how many solute-solvent contacts there are in the solution on average. A given segment has z 2 neighbors that could be occupied either by solvent or by a nonbonded polymer segment. If we have N 2 polymer molecules, each with an average of ν s segments, and N 1 solvent molecules, then the probability that a given lattice point that does not contain a segment bonded to the segment of interest, is occupied by a solvent molecule is N 1 /(N 1 + ν s N 2 3) N 1 /(N 1 + ν s N 2 ). The 3 in the first part of this equation accounts for the fact that the segment of interest and its two directly bonded segments are excluded. Of course, for macroscopic systems, this term is negligible. Since there are z 2 non-bonded sites adjacent to any segment (neglecting the ends of the polymer) and there are ν s N 2 segments in total, the total number of solute-solvent contacts should be (z 2)ν s N 2 N 1 /(N 1 + ν s N 2 ) = (z 2)N 1 ν s N 2 /(N 1 + ν s N 2 ) = (z 2)N 1 φ 2, where φ 2 is the 2

volume fraction of the polymer: φ 2 = V 2 V 1 + V 2 = ν s N 2 N 1 + ν s N 2 in the lattice model since each polymer segment or solvent molecule occupies the same volume (one lattice site). The enthalpy of solution is then simply the number of solute-solvent contacts times the enthalpy change per contact, or sol H = (z 2)N 1 φ 2 ɛ. This is essentially identical (give or take the factor of z 2 instead of z) to the van Laar equation. The same factors which affect the enthalpy of solution for small molecules are thus important to the enthalpy of solution of polymers. One qualitative difference is the size of φ 2 : It takes a lot of small solute molecules to make the solute volume fraction significant, but relatively few polymer molecules due to the large sizes of polymer molecules. If we compare two solutes with similar positive values of ɛ, one of which is a small molecule and one a polymer, the polymer may be only sparingly soluble while the small molecule has moderate solubility due to this difference. 3 Entropy of solution The entropy of solution can have several components. One component, perhaps the obvious one and the one we will focus on here, is the conformational entropy, i.e. the entropy associated with the different possible arrangements of the polymer in solution. There can also be entropy associated with the movements of parts of the polymer other than the backbone if the polymer has significant side chains, and these entropy terms can be different between the pure and solvated polymer. If, for instance, the side chains are locked in position in the pure polymer and are free to move in solution, the theory presented here will underestimate the entropy of solution, and thus the solubility of a polymer. In many cases though, these effects are relatively small compared to the conformational entropy. There is another issue of significance in polymer science which doesn t arise in the smallmolecule theory. When we calculated the entropy of solution in the small-molecule case, because the lattice model only allows one microstate for the segregated state, we had sol S = S mixture S segregated = S mixture. If the solid polymer is crystalline, this relationship can be approximately true for polymer solutions as well. Even crystalline polymers are generally not perfectly crystalline and will usually have amorphous regions (figure 2). In fact, we often characterize polymers in terms of their percent crystallinity, which is rarely more than 90% and can be substantially lower. Some solid polymers in fact exist in an essentially entirely amorphous state, which can be either a glass (hard) or a rubber (soft). 1 An amorphous polymer would have a significant entropy, which would then reduce the entropy of solution 1 Most amorphous polymers go through a phase transition between a glassy and a rubbery phase at some specific temperature, so these are not categories of polymers but categories of behavior. 3

crystalline amorphous crystalline Figure 2: Schematic representation of a polymer with a high percent of crystallinity. The crystalline regions consist of orderly rows of polymer strands. These strands can be joined by turns or by amorphous strands. and make these polymers less soluble than crystalline equivalents. The Flory-Huggins theory describes the solubility of an amorphous polymer. In a lattice model, there is no distinction between an amorphous solid polymer and a liquid polymer, so the theory also applies to solutions made from polymer melts. We want to count the number of different ways we can put N 2 polymer molecules, each of which has ν s chain segments, in a lattice with = N 1 + ν s N 2 sites. The remaining sites will of course be filled with the solvent molecules. Suppose that we have already placed i polymer molecules in the lattice. Let ξ i+1 be the number of different ways we can place the (i + 1)st polymer molecule in the lattice. We calculate ξ i+1 by considering each segment in turn. Since the first i polymers occupy ν s i sites, there are ν s i ways we can place the first segment. The probability that any given neighboring site to the first segment is unoccupied is p 0 = ( ν s i 1)/ ( ν s i)/. Since there are z neighboring sites, on average there are p 0 z different ways we can place the second segment. The second segment has z 1 neighbors, excluding the neighboring site occupied by the first segment. The probability that any given one of these z 1 neighboring sites is unoccupied is p 1 = ( ν s i 2)/ p 0, so the number of different ways to place the third segment is approximately p 0 (z 1). This same reasoning, with the same probability p 0 that neighboring sites are empty, applies to any subsequent segment in our polymer, provided ν s. 2 The number of different ways 2 Quite apart from the approximation used explicitly above, p 0 is an overestimate of the probability that an adjacent site is unoccupied since it takes no account of the fact that adjacent sites are more likely to be occupied by segments from the same polymer than by segments from other polymers because of the connectivity of the polymer molecules. This can be fixed, but it s a relatively small correction which doesn t improve the fit of the theory to experimental data. 4

we can place the ν s segments of the (i + 1)st polymer in the lattice is therefore ξ i+1 ( ν s i)p νs 1 0 z(z 1) νs 2 ( ν s i)p νs 1 0 (z 1) νs 1 = ( ν s i)(z 1) νs 1 ( N0 ν s i ( ) z 1 νs 1 = ( ν s i) νs. ) νs 1 In the second step, we approximated z(z 1) νs 2 by (z 1) νs 1, which is reasonable provided z is not too small and ν s is reasonably large. Now note that [ ν s i]! [ ν s (i + 1)]! = [ ν s i][ ν s i 1]... [ ν s (i + 1) + 1] ( ν s i) νs to the same level of approximation as we used to replace p 1, p 2,... by p 0. Thus we have ( ) [ ν s i]! z 1 νs 1 ξ i+1, [ ν s (i + 1)]! or ξ i [N ( ) 0 ν s (i 1)]! z 1 νs 1. (1) [ ν s i]! The number of ways of placing all N 2 polymer molecules in the lattice is W = 1 N 2! N 2 ξ i. The factor of 1/N 2! arises because our expression for ξ i assumes that we are placing the polymers into the lattice one after the other. Since the solution process is not ordered, we have to divide by the number of permutations of the N 2 molecules. If we substitute equation 1 into W, we can separate the product into two terms: W = 1 N 2! = 1 N 2! N 2 ( z 1 [ ν s (i 1)]! [ ν s i]! ) N2 (ν s 1) N 2 N 2 ( ) z 1 νs 1 [ ν s (i 1)]!. [ ν s i]! The remaining product can be evaluated by expanding the factorials. In the equation that follows, the terms collected together in braces each represent one term in the product: N 2 [ ν s (i 1)]! [ ν s i]! = { ( 1)... ( ν s + 1)} = {( ν s )( ν s 1)... ( 2ν s + 1)}... {... ( N 2 ν s + 1)}! ( N 2 ν s )!. 5

W is therefore W =! N 2!( N 2 ν s )! ( z 1 ) N2 (ν s 1) = (N 1 + N 2 ν s )! N 1!N 2! ( z 1 N 1 + N 2 ν s ) N2 (ν s 1). The conformational entropy is S c = k ln W. Using our expression for W and simplifying, we get { ( ) ( ) } S c = k N 1 ln + N 2 ln + N 2 (ν s 1) [ln(z 1) 1]. (2) N 1 The entropy of solution is given by N 2 sol S = S c S p, where S p is the entropy of the pure polymer, since the entropy of the pure solvent is zero in the lattice model. We can easily get the entropy of an amorphous polymer (or of a molten polymer) from equation 2 by setting N 1 = 0: S p = kn 2 {ln ν s + (ν s 1) [ln(z 1) 1]}. The entropy of solution is therefore, after a small rearrangement, { ( ) ( )} sol S = k N 1 ln + N 2 ln. N 1 N 2 ν s This equation can be rewritten in terms of the volume fractions of the solvent and polymer: of, if you prefer to use moles, sol S = k (N 1 ln φ 1 + N 2 ln φ 2 ), sol S = R (n 1 ln φ 1 + n 2 ln φ 2 ). Although this equation is identical in form to that derived for small molecules, there is an important difference. For small molecules, φ i X i. For polymers, this is of course not the case. This leads to large deviations from ideal behavior for polymer solutions, as observed for instance in vapor pressure measurements where we find large deviations from Raoult s law. Exercises 1. A classic demonstration involves dissolving packing peanuts, which are made from expanded polystyrene, into acetone. A huge amount of polystyrene will dissolve in acetone. Acetone has a density of 0.79 g/cm 3 and a molar mass of 58.08 g/mol. Polystyrene has a density of 1.05 g/cm 3. A particular sample of polystyrene has an average molar mass of 3.5 kg/mol. Suppose that we dissolve 100 g of polystyrene in 100 g of acetone. What is the entropy of solution? 2. Suppose that we have a solvent for which z = 8 and V 1,m = 200 ml/mol, and a polymer with V 2,m = 5000 ml/mol and an average molar mass of 20 kg/mol. The interaction energy per contact is ɛ = 500 J/mol. What is the solubility, in g/l, of the polymer in this solvent at 25 C? 6