112
A Thermodynamic Investigation of the PVT, Solubility and Surface Tension of Polylactic Acid (PLA)/CO 2 Mixtures by Syed Hassan Mahmood A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy, Graduate Department of Mechanical & Industrial Engineering, University of Toronto Copyright by Syed Hassan Mahmood 2012

A Thermodynamic Investigation of the PVT, Solubility and ... · A Thermodynamic Investigation of the PVT, Solubility, and Surface Tension of Polylactic Acid (PLA)/CO 2 Mixtures Syed

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

  • A Thermodynamic Investigation of the PVT, Solubility and Surface Tension of Polylactic Acid (PLA)/CO2

    Mixtures

    by

    Syed Hassan Mahmood

    A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy,

    Graduate Department of Mechanical & Industrial Engineering, University of Toronto

    Copyright by Syed Hassan Mahmood 2012

  • ii

    A Thermodynamic Investigation of the PVT, Solubility, and Surface Tension of Polylactic Acid (PLA)/CO2 Mixtures

    Syed Hassan Mahmood

    Degree of Master of Science, 2012

    Department of Mechanical and Industrial Engineering

    University of Toronto

    ABSTRACT

    This thesis illustrates a comprehensive study on the PVT, solubility and surface tension

    properties of polylactic acid (PLA) with dissolved CO2 based on thermodynamic models. The

    solubility of CO2 in PLA melt was calculated by means of a gravimetric method, using a

    Magnetic Suspension Balance (MSB). The swelling volume of the polymer/gas mixture due to

    dissolution of gas was compensated for by direct measurement through a view cell or by

    theoretical models such as Simha Somcynsky (SS) - Equation of State (EOS) and Sanchez

    Lacombe (SL) - Equation of State (EOS). Three grades of PLA (i.e., PLA3001D, PLA8051D,

    and PLA4060D) were chosen. It was observed that the pressure, temperature, D-content and

    Molecular weight variance had an effect on the swelling and solubility.

    The surface Tension of PLA/CO₂ mixture was also calculated from the captured image

    using the Axsymmetric Drop Shape Analysis (ADSA). The effects of varying the pressure,

    temperature, and molecular weight on surface tension were investigated.

  • iii

    To the one above for providing me

    with an opportunity

    To my parents For endless love, support

    and encouragement

  • iv

    ACKNOWLEDGEMENTS

    I would like to acknowledge the support of my supervisor Prof. Chul B. Park, whose

    encouragement and support was second only to that of my parents, without which this research

    would not have been possible.

    I would also like to thank the members of my Thesis Committee, Prof. Sanjeev

    Chandra and Prof. Kamran Behdinan for their invaluable advice and guidance.

    My gratitude is extended to the Department of Mechanical and Industrial Engineering

    at the University of Toronto for providing the University of Toronto Fellowship. Also, I would

    like to thank the members of the Consortium for Cellular and Micro-Cellular Plastics

    (CCMCP) for their funding and support in this research.

    I would like to thank all my colleagues and fellow researchers formerly and presently

    working in the Microcellular Plastics Manufacturing Laboratory for their friendship,

    cooperation and support. I am especially thankful to Dr. Gary Li, Dr. Takashi Kuboki, Dr. Nan

    Chen, Raymond Chu, Peter Jung, Anson Wong, Lun Howe Mark, Nemat Hossiney, Reza

    Nofar and Mohammad M. Hassan for their collaboration on our common research publications

    and for the thought-provoking discussions we had during the work. I also want to acknowledge

    the contribution of Mohammad M. Hassan and Dr. Guangming Li in the theoretical and

    experimental work.

    I would also like to thank my parents and all my family members for their endless love,

    support and encouragement. Finally, I would like to thank God for providing me with an

    opportunity and the patience to grasp that opportunity.

  • v

    TABLE OF CONTENTS

    Abstract ........................................................................................................................................ii

    Acknowledgements ..................................................................................................................... iv

    List of Figures .......................................................................................................................... viii

    Nomenclature ............................................................................................................................... x

    CHAPTER 1. INTRODUCTION .......................................................................................... 1

    1.1 Preamble ............................................................................................................................. 1

    1.2 Foam Processing ................................................................................................................. 3

    1.2.1 Microcellular Processing ........................................................................................... 3

    1.2.2 Physical Blowing Agents........................................................................................... 4

    1.2.3 Thesis Objectives and Scope of Research ................................................................. 6

    1.2.4 Thesis Structure ......................................................................................................... 7

    CHAPTER 2. LITERATURE SURVEY AND THEORETICAL BACKGROUND ........... 8

    2.1 Solubility ............................................................................................................................. 8

    2.2 Solubility Measurement Methods ..................................................................................... 10

    2.2.1 Gravimetric Sorption Technique ............................................................................. 11

    2.2.2 Pressure Decay Sorption Technique ........................................................................ 13

    2.2.2.1 Previous Research Using Pressure Decay Sorption Technique ....................... 15

    2.2.3 Volume Decay Sorption Technique ........................................................................ 16

    2.2.4 Piezoelectric Quartz Sorption .................................................................................. 17

    2.2.5 In-Line Measurement of Gas Solubility .................................................................. 19

    2.2.5.1 In-Line Monitoring ........................................................................................... 19

    2.2.5.2 In-Line infrared Sensors ................................................................................... 20

    2.2.5.3 Ultrasound ........................................................................................................ 21

    2.2.6 Modified Magnetic Suspension Balance Theoretical Treatments ........................... 22

    2.3 Theoretical Treatments ..................................................................................................... 26

    CHAPTER 3. RESEARCH METHODOLOGY FOR PVT AND SOLUBILITY STUDY 29

    3.1 Introduction ....................................................................................................................... 29

    3.2 Theoretical Background .................................................................................................... 29

    3.3 Methodology and Approach ............................................................................................. 32

    3.4 Summary ........................................................................................................................... 34

    CHAPTER 4. SOLUBILITY AND SWELLING BEHAVIOR OF PLA IN PRESENCE

    OF CO₂………………………………………………………………………………………...35

  • vi

    4.1 Introduction ....................................................................................................................... 35

    4.2 Experimental ..................................................................................................................... 36

    4.2.1 Materials .................................................................................................................. 36

    4.2.2 PVT Data for PLA ................................................................................................... 36

    4.3 Solubility of CO₂ in PLA (Binary System)....................................................................... 37

    4.3.1 Swelling Behavior of PLA in Presence of CO2 ....................................................... 37

    4.3.1.1 Experimental Setup .......................................................................................... 37

    4.3.1.2 Experimentation ............................................................................................... 37

    4.3.1.3 Pressure and Temperature Effect on Volume Swelling ................................... 38

    4.3.1.4 Comparison of the experimentally measured data and Theoretically Predicted

    Data ………………………………………………………………………………..40

    4.3.1.5 Effect of ‘D’ content/Molecular Weight on volume swelling .......................... 44

    4.3.2 Solubility of CO2 in PLA......................................................................................... 45

    4.3.2.1 Pressure and Temperature effect on solubility of CO₂ in PLA. ....................... 45

    4.3.2.2 Comparison of Theoretical and Experimental Solubility ................................. 48

    4.3.2.3 Effect of D-content/Molecular weight on Solubility of CO₂ ........................... 51

    4.4 Summary ........................................................................................................................... 53

    CHAPTER 5. Surface Tension OF PLA/CO₂ MELT ......................................................... 54

    5.1 Introduction ....................................................................................................................... 54

    5.2 Experimental Materials ..................................................................................................... 57

    5.3 Measurement of Surface Tension of PLA/CO₂ mixture ................................................... 57

    5.3.1 Surface Tension of PLA/CO₂ melt .......................................................................... 57

    5.3.2 Density determination.............................................................................................. 59

    5.3.3 Effects of pressure and temperature variance on surface tension ............................ 62

    5.3.4 Effect of Molecular weight on surface tension ........................................................ 65

    5.4 Summary ........................................................................................................................... 66

    CHAPTER 6. CONCLUSIONS AND FUTURE WORK ................................................... 68

    6.1 Summary ........................................................................................................................... 68

    6.2 Recommendations and Future Work ................................................................................ 69

    References .................................................................................................................................. 70

    Appendix .................................................................................................................................... 86

  • vii

    LIST OF TABLES

    Table 1: The scaling Parameters for SS-EOS are as following ................................................. 96

    Table 2: The scaling Parameters for SL-EOS are as following ................................................. 96

    Table 3: Solubility of PLA3001D at 180 °C .............................................................................. 96

    Table 4: Solubility of PLA8051D at 180 °C .............................................................................. 97

    Table 5: Solubility of PLA4060D at 180 °C .............................................................................. 97

    Table 6: Solubility of PLA3001D at 200 °C .............................................................................. 97

    Table 7: Solubility of PLA8051D at 200 °C .............................................................................. 98

    Table 8: Solubility of PLA8051D at 200 °C .............................................................................. 98

    Table 9: Swelling ratio of PLA3001D/CO₂ melt at 180 °C ....................................................... 98

    Table 10: Swelling ratio of PLA8051D/CO₂ melt at 180 °C ..................................................... 99

    Table 11: Swelling ratio of PLA4060D/CO₂ melt at 180 °C ..................................................... 99

    Table 12: Swelling ratio of PLA3001D/CO₂ melt at 200 °C ..................................................... 99

    Table 13: Swelling ratio of PLA8051D/CO₂ melt at 200 °C ..................................................... 99

    Table 14: Swelling ratio of PLA4060D/CO₂ melt at 200 °C ................................................... 100

  • viii

    LIST OF FIGURES

    Figure 2-1 Bubble nucleation inside a die ................................................................................... 9

    Figure 2-2: Bubble nucleation inside a mold in injection molding machine: (a) nucleation starts

    at gate (generates high cell density); (b) nucleation starts in the mold cavity (generates low cell

    density). [14] .............................................................................................................................. 10

    Figure 2-3 Details of MSB for density measurement [65] ........................................................ 23

    Figure 4-1: Effect of Temperature variance on Swelling Ratio for PLA3001D ........................ 39

    Figure 4-2: Effect of Temperature variance on Swelling Ratio for PLA8051D ........................ 39

    Figure 4-3: Effect of Temperature variance on Swelling Ratio for PLA4060D ........................ 40

    Figure 4-4: Swelling Ratio of PLA3001D at 180 °C with varying pressure ............................. 41

    Figure 4-5: Swelling Ratio of PLA8051D at 180 °C with varying pressure ............................. 41

    Figure 4-6: Swelling Ratio of PLA4060D at 180 °C with varying pressure ............................. 42

    Figure 4-7: Swelling Ratio of PLA3001D at 200 °C with varying pressure ............................. 42

    Figure 4-8: Swelling Ratio of PLA8051D at 200 °C with varying pressure ............................. 43

    Figure 4-9: Swelling Ratio of PLA4060D at 180 °C with varying pressure ............................. 43

    Figure 4-10: Effect of varying D-content/Mw on swelling ratio at 180 °C ............................... 44

    Figure 4-11: Effect of varying D-content/Mw on swelling ratio at 200 °C ............................... 45

    Figure 4-12: Effect of varying temperature on solubility of PLA3001D .................................. 46

    Figure 4-13: Effect of varying temperature on solubility of PLA8051D .................................. 47

    Figure 4-14: Effect of varying temperature on solubility of PLA4060D .................................. 47

    Figure 4-15: Solubility of Carbon Dioxide in PLA3001D at 180 °C with varying pressure .... 48

    Figure 4-16: Solubility of Carbon Dioxide in PLA8051D at 180 °C with varying pressure .... 49

    Figure 4-17: Solubility of Carbon Dioxide in PLA4060D at 180 °C with varying pressure .... 49

  • ix

    Figure 4-18: Solubility of Carbon Dioxide in PLA3001D at 200 °C with varying pressure .... 50

    Figure 4-19: Solubility of Carbon Dioxide in PLA8051D at 200 °C with varying pressure .... 50

    Figure 4-20: Solubility of Carbon Dioxide in PLA4060D at 200 °C with varying pressure .... 51

    Figure 4-21: Effect of varying D-content/Mw on Solubility of Carbon Dioxide at 180 °C ...... 52

    Figure 4-22: Effect of varying D-content/Mw on Solubility of Carbon Dioxide at 200 °C ...... 52

    Figure 5-1: Density difference with pressure change at 180 °C ................................................ 60

    Figure 5-2: Density difference with pressure change at 200 °C ................................................ 61

    Figure 5-3: Change in surface tension with pressure at 180 °C ................................................. 63

    Figure 5-4: Surface tension at various temperature and pressure .............................................. 64

    Figure 5-5: Relationship between surface Tension and Solubility ............................................ 65

    Figure 5-6: Effect of molecular weight on Surface Tension ...................................................... 66

  • x

    NOMENCLATURE

    Ai = Helmholtz energy [J]

    CS = Costas and Sanctuary

    ci = Chain (molecule) flexibility; 3c is total external degrees o f freedom attributed to

    a chain (molecule)

    d = Bond length in PCM EOS

    EOS = Equation of state

    = Molar Gibbs free energy for polymer/gas mixture [J/mol]

    H = Enthalpy [J]

    h = Plancks constant 6.6260755×10-34

    [J.s]

    , Bk k = Boltzmann constant 1.380658×10

    -23 [J/K]

    im = Molar mass of segment (mer) of “i” component [g/mol]

    M = Molecular weight of per molecule [g/mol], M = misi

    = Avogadro’s number 6.0221367×1023

    = Pressure [Pa]

    = Characteristic pressure of component “i” [Pa],

    ** i

    i *

    i

    qzP

    = Reduced pressure */ PP

    PCM = Prigogine cell model

    PC-

    SAFT

    = Perturbed Chain Statistical Associating Fluid Theory

    = The number of nearest neighbor sites per chain-like molecule (s-mer), si(z - 2)+2.

    = Dimensionless identity for SS-EOS

    = The combinatorial factor for PCM

    r = Number of mer per molecule, **

    *

    ρRT

    MPr

    = Gas constant 8.3143 [J/(mol·K)]

    S = Entropy [J/K]

  • xi

    SAFT = Statistical Associating Fluid Theory

    = Number of mers per molecule of component “i”

    SL = Sanchez–Lacombe

    SS = Simha–Somcynsky

    = Volume swelling ratio

    T = Temperature [K]

    *

    iT = Characteristic temperature of component “i” [K], ii

    B

    qzT

    ck

    = Reduced temperature */TT

    *

    iυ = Characteristic volume per mer of component “i” [m3/mer]

    *

    iV = Characteristic volume of component “i” (m3/mol)

    = Reduced volume,

    ix = Mole fraction of “i” component in mixture system

    X = Solubility (g-gas/g-polymer)

    y = Occupied lattice site fraction

    Z = 12, the lattice coordination number

    i * = Characteristic energy per mer of component “i” (J/mer)

    = dimensionless number for SS-EOS

    1 = Volume fraction of gas in mixture system

    2 = Volume fraction of polymer in mixture system

    G

    1 = Chemical potential of gas in vapor phase [J/mol]

    P

    1 = Chemical potential gas in the polymer melt [J/mol]

    * = Characteristic density of bulk material [g/cm3]

    Λ

    = de Broglie wavelength

    = Flexibility parameter

    = Lattice-site volume

    Γ = PCM geometrical constant

  • xii

    ~ = Reduced density V

    ~/1

    Subscripts

    1 = Indicates gas

    2 = Indicates polymer

    i = Indicates component number or component in x-direction

    j = Indicates component number in y-direction

    F = Indicates foam

    G, g = Indicates gas

    Superscripts

    G, g = Indicates gas or vapor

    P, p = Indicates polymer

  • 1

    CHAPTER 1. INTRODUCTION

    1.1 Preamble

    Foamed plastics surround us in the shape of different commodities every day. This is a

    testament to the extent that foamed polymeric materials have pervaded our everyday life.

    Foamed polymeric material can be classified as either a thermoplastic or thermoset foam.

    Thermoplastic foams are recyclable, whereas thermoset foams cannot be reprocessed because

    it’s extensive crosslinking. The applications of the foamed product dictate the density of the

    final product; consequently, several types of polymeric materials are foamed to various low

    densities for applications that require attributes such as weight reduction, insulation, buoyancy,

    energy dissipation, convenience, and comfort.

    Thermoplastic and thermoset foam products differ from solid plastic products and are

    identifiable by virtue of a unique cellular structure. The latter is obtained by the atmospheric

    expansion of a blowing agent in the polymer matrix. This cellular structure is normally

    characterized in terms of the cell density and cell size. For conventional foams, typical cell

    population densities are within the range of 10³-10⁶ cells/cm³, with cell sizes of the order of

    100 μm or larger. However, the cell size and distribution can be inconsistent, hence

    compromising the mechanical properties of the foam. Conversely, microcellular plastics are

    characterized by cell densities greater than 10⁹ cells/cm³ and cell sizes smaller than 10 μm.

    This group of plastic foams was conceived by Masrtini et al. [1], and is based on the notion

    that if a large number of bubbles smaller than the pre-existing flaws in the polymer are created,

    the material cost can be reduced without compromising mechanical properties. It has been

  • 2

    observed that microcellular plastics possess superior impact strength [1], toughness [2], and

    fatigue life [3] compared to solid polymers.

    Furthermore, solid polymeric foams may possess either closed or open cells; closed

    cell foams possess a cellular structure in which neighboring air bubbles are entrapped in a

    continuous macromolecular phase. Open cell foams, on the other hand, have a cellular network

    in which continuous channels are available throughout the solid macromolecular phase.

    The gaseous phase in any polymeric foam material is obtained using blowing agents in the

    foam manufacturing process. There are generally two types of blowing agents used in foam

    production: chemical blowing agents and physical blowing agents. Chemical blowing agents

    are chemical compounds which evolve gases under foam processing conditions through

    thermal degradation or chemical reactions. Physical blowing agents, on the other hand, are

    inert gases, such as nitrogen and carbon dioxide; volatile hydrocarbons such as propane, n-

    butane, i-pentane; and low boiling point chlorofluorocarbons (CFCs), hydrofluorocarbons

    (HFCs), and hydrochlorofluoro-carbons (HCFCs).

    Due to the environmental hazard posed by CFCs and HCFCs, there has been a drive to

    replace these blowing agents with more environmentally friendly substitutes. The properties of

    blowing agents and their impact on the processing variables in foam production are essential

    issues that must be considered in order to find effective replacements. The specific

    thermodynamic and kinetic properties that have the greatest influence on the ability to produce

    microcellular foams are the solubility and diffusivity of the blowing agent in the polymers.

    The solubility determines the amount of blowing agent that can be absorbed by the polymer at

    any given temperature and pressure, while the diffusivity determines both the rate at which the

    blowing agent will penetrate into the polymer matrix to form a homogeneous solution, as well

  • 3

    as the rate at which the blowing agent escapes to the atmosphere during cell nucleation and

    growth processes.

    1.2 Foam Processing

    1.2.1 Microcellular Processing

    The foamed plastics industry is continuously developing, which encompasses various

    methods of production for many product applications. The choice of polymer, blowing agent,

    and production method dictates the foam formation and its morphology; therefore, its

    properties, stemming from an industrial challenge to enhance polymer value by improving

    performance/weight characteristics. Earlier work focused on batch processes; the concepts

    resulted in good foam structures with cell sizes less than 5 μm using supercritical carbon

    dioxide [1]. The batch process was further developed and improved by Cha et al. [4]; whereas

    Kumar [5] and Kumar and Suh [6, 7] developed a semi-continuous process derived from a

    modified thermoforming process. Further investigation by Park [8], Park et al. [9], and Park

    and Suh [10] led to the development of an extrusion-based process for a microcellular filament

    as the first step in the development of a continuous process.

    The fundamental principle involved in the formation of microcellular polymer foams

    consists of three basics steps: 1) polymer/blowing agent solution formation; 2) microcellular

    nucleation; and 3) cell growth and density reduction. The single-phase polymer/blowing agent

    solution in the first step are formed by saturating the polymer with the blowing agent under

    high pressure. The saturation point is determined by the solubility limit of the blowing agent in

    the polymer, while the time required for the solution formation is determined by the rate of

    diffusion of the blowing agent into the polymer matrix.

  • 4

    Microcellular nucleation is achieved by inducing a thermodynamic instability in the

    single-phase solution. This is usually accompanied by drastically reducing the solubility of the

    gas in solution by controlling the pressure and/or temperature of the solution [11-13]. Since the

    separation of the polymer and gas phases is thermodynamically more favorable, the resulting

    supersaturated mixture becomes the driving force for the nucleation of numerous microcells

    [1]. Continuous microcellular processing typically utilizes a rapid pressure drop to nucleate

    bubbles. This stage is very crucial to the overall process, because it dictates the cell

    morphology of the material and its resulting properties. Therefore, solubility as a function of

    pressure is important for the development of the process.

    The final stage in the production of microcellular plastics is cell growth. After cell

    nucleation has occurred, any available gas diffuses into the cell and increases the cell size,

    thereby reducing the density of the polymer matrix. Generally, cell growth is affected by the

    time allowed for the cells to grow, the system temperature, the amount of gas available (state

    of super saturation), the processing pressure, and the viscoelastic properties of the polymer/gas

    solution [14].

    1.2.2 Physical Blowing Agents

    An important aspect of the creation of a cellular structure is the use of a physical

    blowing agent. Historically, CGCs and HDCs, such as CFC-11 and HCFC-141b, were used

    primarily for low density foam production mainly because of their solubility, volatility, and

    non-toxic nature; however, their stability and reactivity with ozone in the atmosphere raised

    substantial concern about ozone depletion. In an effort to address the environmental impact of

    these man-made compounds, the Montreal Protocol [15] was signed in 1989 by 29 countries

    and amended in subsequent years. The Montreal Protocol mandates the gradual phase-out of

  • 5

    the production of CFCs by 2010 and HCFCs by 2030 undeveloped countries. As a result, CFC

    production fell from 980 metric tons in 1986 to 95 metric tons in 1996 [16]. Since the early

    1990s, global warning or the greenhouse effect has become another major issue. The

    Intergovernmental Panel on Climate Change (IPCC) reported a scientific assessment on the

    warming potential of various compounds relative to carbon dioxide [17]. It was reported that

    the global warming potential of CFCs and HXFXs is 5,000-10,000 times greater than that of

    carbon dioxide with stratospheric life cycles of 60-130 years. In addition, the Kyoto Protocol

    on Climate Change [18] adopted by 160 nations in 1997, sets binding limits on greenhouse gas

    emissions for developed countries. It strengthens the framework established by the Montreal

    Protocol with new policies and measures.

    The combined global warming and ozone depleting potential of these substances, as

    well as the resulting policies, have created a void in the foam processing industry; the need

    has, therefore, arisen to replace CFCs and HCFCs with environmentally friendly blowing

    agents possessing good foam-blowing properties. The concern over the ozone layer and global

    warming represent just a few of the issues facing the foam processing industry. Disposal,

    waste stream control, and usage of recycled plastics still require a deep understanding of

    foaming technology.

    Microcellular plastics do not utilize the conventional CFC foam blowing agents;

    instead, carbon dioxide and nitrogen are typically used. Although these blowing agents

    represent a dramatic improvement in terms of environmental hazard, they, too, pose their own

    difficulties. Carbon dioxide has a high solubility in the plastic melt, approximately 10% at 200

    °C and 27.6 MPa [19], and produces a very uniform cell structure. Conversely, nitrogen has a

    lower solubility, about 2-3% at 200 °C and 27.6 MPa, and processing requires a much higher

    pressure to dissolve sufficient nitrogen to create a uniform structure as explained in Section

  • 6

    1.2.1. Higher processing, however, requires more robust processing equipment and leads to

    higher equipment and processing costs. Other alternative blowing agents used are liquid

    blowing agents such as butane and iso-pentane. To date, however, liquid blowing agents have

    not been successfully used to achieve a microcellular structure. One possible barrier that exists

    with the use of these liquid blowing agents is the lack of information about their solubility and

    diffusivity.

    1.2.3 Thesis Objectives and Scope of Research

    Since the solubility of blowing agents in polymer melts plays a key role in the plastic

    foaming processing, this research was focused on the following aspects:

    i. To propose technically sound experimental approaches and thermodynamic

    models for the PVT, solubility and surface tension investigation of polymer/gas

    mixture (binary).

    ii. To obtain reliable (more accurate) PVT, solubility data, surface tension and

    thermodynamic properties of polymer/blowing agents by systematic

    investigation

    iii. To verify the accuracy of solubility data determined by using various EOSs

    and correct them with the help of a visualization system.

    iv. To determine the effect of D-content/Mw on the solubility, surface tension and

    swelling of the PLA/gas mixture.

  • 7

    1.2.4 Thesis Structure

    This section provides a brief overview of the present thesis.

    Chapter 2 explores the literature review on the thermodynamic study of the phase equilibrium

    in a polymer system and gas solubility in polymer melts. The various methods available to

    measure the solubility and diffusivity of blowing agents in polymer are examined. Theoretical

    methods to predict the solubility is discussed as well.

    Chapter 3 introduces the general research methodologies for the study of PVT and solubility

    behavior of polymers. It was found that there are some deviations existing among the

    thermodynamic models in terms of theoretical solubility and swollen volume prediction. As a

    result, further investigation was done to verify the equation of states (EOSs).

    Chapter 4 introduces the study of PVT and the solubility behavior of polymers/gas (binary

    system). The solubility of CO2 in polymer was determined by using experimental data. The

    visualization system was used to obtain the swelling behavior of the polymer/gas mixture. The

    obtained data is then compared with the theoretical data obtained by means of EOS, namely

    SS-EOS and SL-EOS.

    Chapter 5 encompasses the measurement of the surface tension of the PLA/CO₂. Effect on the

    interfacial tension with the variance in pressure, temperature, molecular weight, and D-content

    is investigated. SS-EOS is used to obtain theoretical density of the PLA/CO₂ and was

    compared to the experimentally obtained data.

    Chapter 6 provides a summary, as well as conclusions of the research. Recommendations for

    future work are also presented in Chapter 6.

  • 8

    CHAPTER 2. LITERATURE SURVEY AND THEORETICAL

    BACKGROUND

    2.1 Solubility

    Solubility is defined as the amount of physical blowing agent (PBA) that can be

    dissolved into a unit mass of polymer at a particular temperature and pressure (where the unit

    is g-PBA/g-Polymer). The solubility is not only a critically important parameter for fabricating

    plastic foams, but also an important property in developing blowing agents (BAs) and

    evaluating their performances.

    For effective process design, the system pressure must be high enough to dissolve all of

    the injected gas into the polymer melt. When the polymer/gas solution exits the extrusion die,

    the pressure will drop dramatically; this will initiate the bubble nucleation. The bubbles’

    nucleation stage is crucial in the plastic foaming process due to the formation of a

    microcellular structure. Theoretically, cell nucleation occurs when the pressure of the

    polymer/gas mixture drops below the solubility pressure (or threshold pressure [20] to be

    exact), as shown in Fig. 2.1 The cell nucleation mechanism has been described in detail by lot

    of researchers [21-25], where the nucleation rate was governed by the degree of super

    saturation, i.e., a metastable state determined by the saturation or solubility information.

  • 9

    Figure 2-1 Bubble nucleation inside a die

    Blander and Katz [21-23] have reviewed the classical nucleation theory to estimate the

    rates of bubble nucleation in pure liquids. The work of formation, W, for a spherical bubble of

    radius, R, is shown in equation 1-2:

    2-1

    where is the surface tension; is the pressure of the bubble at the moment it is formed,

    which is typically determined as a saturation pressure in a pure component system or solubility

    pressure corresponding to the amount of dissolved gas in the mixture system; is the pressure

    of the system; n is the number of bubbles; and are the chemical potentials of the

    new and old phase, respectively. Hence, the solubility pressure information is required for the

    calculation of the nucleation point and nucleation rate. Also, to study surface tension of the

    polymer/gas mixture, solubility information is a prerequisite. For this reason, the solubility of

    gases in a polymer melt during the plastic foam processing condition has been of great interest

    to foaming manufacturers and researchers.

    Solubility information also plays a very important role in foaming with an injection

    molding machine. In our group, Lee et al. [26] found the cavity pressure of a foaming mold

  • 10

    has a significant influence on cell nucleation (Fig. 2.2). He claimed that if the cavity pressure

    is lower than the solubility pressure (or the threshold pressure [20]) of the injected gas and if

    the pressure before the gate is high enough, cell nucleation occurs at the gate with a high

    pressure drop rate. In such cases, the cell density will be high. However, if the cavity pressure

    is higher than the solubility pressure (or the threshold pressure), cell nucleation occurs along

    the mold cavity with a low pressure drop rate, resulting in a low cell density.

    (a) (b)

    Figure 2-2: Bubble nucleation inside a mold in injection molding machine: (a) nucleation

    starts at gate (generates high cell density); (b) nucleation starts in the mold cavity (generates

    low cell density). [14]

    2.2 Solubility Measurement Methods

    There are two experimental techniques, known as permeation and sorption that can be

    used to determine the solubility of a gas in a polymer. Permeation experiments involve

    measurements of the steady state mass flow of a gas flowing through a thin membrane;

  • 11

    whereas in the sorption kinetics techniques, the mass uptake of gas by the polymer sample is

    measured. The major difference between the two techniques is in the method by which

    solubility is determined. Permeation experiments rely primarily on a mathematical expression

    to determine the permeability, diffusivity, and hence the equilibrium solubility of the gas in the

    polymer indirectly when steady state flow has been attained. The sorption method, on the other

    hand, directly measures the mass gain of the polymer due to gas dissolution and, therefore,

    represents a more direct approach to determining the equilibrium solubility. The sorption

    method will be employed in this thesis to determine the solubility information for polymer-

    blowing agent combinations, and hence this section will focus only on a review of sorption

    techniques.

    2.2.1 Gravimetric Sorption Technique

    The gravimetric sorption method measures the solubility by simply measuring the mass

    gain of a polymer sample due to gas dissolution. One of the earliest gravimetric techniques

    utilized the quartz spring measuring system known as the McBain Balance [27]. The balance

    was operated by suspending the polymer from a quartz spring in a low pressure gas

    environment. As the polymer gained weight due to gas dissolution, the spring elongated.

    Utilizing Hooke’s Law, the mass of the sample can be determined as a function of elongation

    by calibrating the spring with known weight increments. The quartz spring method was used,

    for example, to determine the solubility data of ethylbenzene [28] and toluene [29] in

    polystyrene at various vapor pressures.

    Batch process, a high pressure gravimetric technique, was utilized by Baldwin et al.

    [30] for the measurement of carbon dioxide solubility and diffusivity in thermoplastic

    polyesters. This process utilizes multiple samples, which are exposed to the gas for various

  • 12

    time periods, and compiles the mass uptake curve by normalizing the time axis for sample

    thickness.

    An in situ gravimetric sorption method directly measures the solubility by measuring

    the mass gain of the polymer with a high-precision electro balance capable of measurements at

    high temperatures and pressures. The sensitivity of the instrumentation, which can attain

    values of 1 ppm (part per million), makes the technique desirable for solubility measurement

    involving low solubility gases such as inert gases. This method is also suitable for the

    measurement of solubility for polymers in either the rubbery or glassy state. The apparatus is

    mounted on a vibration-free surface with the weighing unit contained in a constant temperature

    environment.

    The solubility of the blowing agent is determined from measurements of the increased

    mass of the sample with increasing blowing agent pressure. Wong et al. [31] reported on the

    use of an electronic microbalance to measure the gas solubility and diffusivity of carbon

    dioxide and HFC134a in PS, filled poly (vinyl chloride) -FPVC, and unplasticized poly (vinyl

    chloride) UPVC.

    An alternative method for measuring gas solubility in polymers was presented by

    Chaudhary and Johns [32]. It involved using a magnetic suspension device similar to the

    electro balance. The most significant difference is that the weighing mechanism is physically

    decoupled from the high temperature and high pressure environments through a magnetic

    suspension coupling. This equipment was used to measure the solubilities of nitrogen,

    isobutane, and carbon dioxide in polyethylene. More recently, Sato et al. [33] reported on the

    use of a magnetic suspension balance to measure the solubilities of carbon dioxide in

    poly(vinyl acetate) (PVAc) and polystyrene.

  • 13

    The high sensitivity of these types of balances dictates that mass measurements must

    be corrected to accommodate for the change in buoyancy of the sample. Therefore, knowledge

    of the dilation of the polymer with blowing agent uptake is also required for the solubility

    calculations.

    2.2.2 Pressure Decay Sorption Technique

    The pressure decay technique is used to determine the solubility of gas in a closed

    system by measuring the pressure decrease due to gas dissolution in the polymer sample. This

    method relies on the assumption that all changes in the gas pressure are due to mass sorption

    of the polymer. Consequently, the mass uptake of the polymer is determined indirectly by

    measuring the pressure decay of the fixed volume system. By measuring the apparatus volume

    accurately and recording the temperature and pressure of the system, the mass of the gas in

    closed system is determined as a function of time using its equation of state. The solubility is

    then determined from the overall experimental pressure change. This technique requires

    careful calibrations and can be used only for gases whose equations of state are accurately

    known. The three methods used for employing the pressure decay technique are single-, dual-,

    and three-chamber systems.

    Single-Chamber Sorption: The single-chamber system [34-37, 11-12] consists of a single

    chamber containing the polymer sample. The chamber is subjected to a rapid pressure

    increase, and the resulting pressure decay due to sorption is recorded as a function of time.

    Due to rapid mass gain in initial stages and the thermodynamics of the gas system, a stable

    reading is often not recorded until the pressure reading (needed to determine the initial mass of

    gas in the system) is extrapolated from the pressure decay curve. The extrapolation, however,

  • 14

    can cause a significant error in determining the initial mass of gas present in the system, and

    the corresponding total mass change due to gas sorption in the polymer.

    Dual-Chamber Sorption: The dual-chamber sorption system [13, 38] uses a reservoir chamber

    of known volume, filled with gas at a known pressure, while another chamber contains the

    polymer sample. By opening a valve separating both chambers, the gas is allowed to flow into

    the second chamber and, therefore, into the polymer. The valve is then closed and the sorption

    chamber is observed for pressure decay. The reservoir chamber pressure is also measured. The

    mass absorbed by the polymer sample is then determined based on the difference of the initial

    mass of gas in the reservoir chamber and the final mass of gas in both the reservoir and

    sorption chambers.

    Sorption experiments are usually performed in a stepwise manner in order to make sure

    the pressure drop for each experimental pressure step is relatively small. A typical pressure

    decay sorption experiment begins with a low reservoir and corresponding low sorption

    pressure. The resulting pressure decay due to sorption is observed until the equilibrium

    pressure indicates that equilibrium mass gain has occurred. New gas is introduced into the

    system without evacuating the chambers so that the pressure is increased by a step amount.

    The pressure decay is then monitored, and the process is repeated. The solubility is determined

    as a function of pressure by successively adding the mass gain of each pressure.

    Three-Chamber Sorption: The three-chamber system [39,40] uses the measurement principle

    identical to the dual-chamber configuration described above; that is, the mass uptake of the

    polymer is determined from the equation of state of the gas, using measurements of chamber

    volume, gas pressure, and gas temperature. The three-chamber system consists of two

    reservoir chambers: the first reservoir is used as a pressure source for the sorption chamber;

    while the other is used as a source for the first reservoir. This configuration allows for multiple

  • 15

    measurements at different pressures without introducing new gas and a new sample. This

    method minimizes the temperature shock to the system caused by the introduction of gas.

    2.2.2.1 Previous Research Using Pressure Decay Sorption Technique

    Nevitt and Weale [13] were responsible for some of the earliest measurements of gas

    solubilities in polymers using the dual-chamber system. They reported on the solubility of

    hydrogen and nitrogen in polystyrene over the pressure of 8.1-30.4 MPa, and at elevated

    temperatures up to 190 °C. High pressure was achieved in the reservoir chamber by using a

    mercury pump. The pressure in the sorption chamber was measured one minute after first

    subjecting the sample to the high-pressure gas. This delay in measurement was a result of the

    pressure instability produced by the initial expansion of gas into the sorption chamber; this

    was further compounded since the gas was not pre-heated to match the temperature of the

    sorption chamber.

    The unstable pressure observed initially contributed to the difficulty experienced in

    determining the initial pressure reading required to calculate the equilibrium solubility. To

    reduce the magnitude of this error, the researchers employed a large sample of 40-100 grams,

    which was cut into thin strips to increase the mass diffusion rate (or reduce the time required to

    obtain equilibrium stability), and thus increase the magnitude of the pressure drop. Utilizing

    the stepwise sorption technique described earlier, the solubility was then calculated as a

    function of pressure.

    Lundberg et al. [34,35] and Lundberg [36] used a single-chamber sorption apparatus to

    determine the solubility of gases in polymers at pressures between 3 and 71 MPa, and

    temperatures between 102 and 188 °C. The stepwise sorption experiment was used to estimate

    the solubility and diffusivity of a gas in a molten polymer.

  • 16

    Durril [37] and Durril and Griskey [11,12] employed a pressure decay method with a

    single-chamber apparatus to investigate the solubility and diffusivity coefficients of nitrogen,

    helium, carbon dioxide, and argon in molten polyethylene, polyisobutylene, and

    polypropylene, at pressures up to 2 MPa. Before coming into contact with the sample, the test

    gas was preheated in a thermostatted air environment. The first pressure reading, however, was

    not reduced until 100 seconds after the gas first contacted the polymer sample. A stepwise

    sorption methodology was used to calculate the solubility as a function of pressure.

    Other researchers have utilized the pressure decay method to only measure the

    solubility characteristics of gases in polymers at pressures up to 2 MPa [41-43] and 8.3 MPa

    [44]. Stern and De Meringo [38] used a dual-chamber system to measure the solubility of

    carbon dioxide in cellulose acetate at pressures up to 4.6 MPa.

    Sato et al. [39] employed a three-chamber sorption apparatus to measure the

    solubilities of carbon dioxide and nitrogen in polystyrene for pressures up to 20 MPa, and

    temperatures in the range of 100-180 °C. The sorption chambers were controlled to within

    0.05 K by a constant temperature air bath. In a later publication, Sato et al. [40] reported on the

    solubility of carbon dioxide and nitrogen in polypropylene, high-density polyethylene, and

    polystyrene. PVT measurements of the polymer at high temperatures and pressures were

    conducted to provide the volume of polymer necessary for the solubility calculations, while

    the swollen polymer volumes, caused by gas dissolution at different pressures and

    temperatures, were predicted using the Sanchez Lacombe Equation of State [45-47].

    2.2.3 Volume Decay Sorption Technique

    As the name implies, volume decay sorption techniques measure the volume change of

    the gas due to polymer sorption in a closed system at constant pressure and temperature. The

  • 17

    mass uptake of the polymer, or essentially the solubility, is indirectly determined from

    measurements of the volume decay.

    A volume decay sorption apparatus was utilized by Rosen [48] to measure the

    solubility and diffusivity of acetone in cellulose acetate, methyl chloride vapor in polystyrene,

    and water vapor in neoprene, at sub-atmospheric pressures. The system was designed as an

    alternative to the quartz spring apparatus.

    Mulrooney [49] used a constant pressure sorption concept based on the volume decay

    method to investigate the solubility and diffusivity of liquid blowing agents such as isopentane

    in polystyrene at elevated pressures. A positive displacement syringe pump capable of

    operating in a constant pressure mode was used as the constant pressure source, while the

    entire assembly was operated in a thermostatted air bath for constant temperature control. The

    reasoning was that since the system was closed, any volume changes occurring in the blowing

    agent were correlated to the piston movement of the pump and electronically recorded.

    However, the swelling effect of the polymer could not be accounted for. If the volume

    increased equally as the volume of isopentane decreased, then no net change in volume would

    be observed. However, if the volume change of the isopentane was less than the volume

    change of the sample, the net measured volume change would be underestimated.

    2.2.4 Piezoelectric Quartz Sorption

    Piezoelectric quartz sorption is a technique which measures solubility based on the

    principle that the vibration frequency of a quartz crystal changes in response to a change in the

    mass deposited on the crystal surface. This technique is usually applied to organic solvents.

    There are two main components in this experimental set up: a sorption cell containing

    the polymer coated with the piezoelectric crystal oscillator, and a solvent cell containing the

  • 18

    gas. When gas is introduced into the sorption cell, it is adsorbed onto the polymer. This, in

    turn, changes the frequency of the crystal oscillator, which is measured with a frequency

    detector, recorded, and indicated on a frequency counter. Also, this experiment incorporates a

    few other variables that could lead to a frequency change, which include the following: the

    sorption of gas into the polymer, adsorption of gas onto the crystal, coating of polymer film,

    hydrostatic pressure of ideal gas, and viscous resistance of the gas.

    Bonner and Cheng [50, 51] experimentally determined that the frequency of a quartz

    crystal oscillator do vary with temperature and pressure. Hence, in order to account for the

    pressure dependence of the frequency in their sorption measurements, two crystal oscillators

    with similar pressure dependencies are used. One of the sorption crystals is coated with the

    polymer, while the other uncoated crystal oscillator is used as a reference crystal. In a situation

    when a reference crystal is not used, an accurate estimate of the pressure dependence of the

    crystal oscillator at the experimental temperature would be needed. Such an estimate has been

    reported by Stockbridge [52] for pressures below 0.13MPa.

    By using the experimental technique with the uncoated reference crystal, Masuoka et

    al. [53] investigated the solubilities of benzene, cyclohexane, n-hexane, toluene, and

    ethylbenzene in polyisobtylene at low temperatures up to 65 °C and low pressures. The effect

    of polymer coating thickness (in the range of about 0.2-1.4 µm) on the solubility of the solvent

    in the polymer is tested, and they found that the experimental results were not affected within

    this range. On top of that, they also concluded that the molecular weight had no definitive

    effect on the solubility for polymer molecular weights of 50,000 and 100,000. In an

    experiment dispensing with the reference crystal, Wang et al. [54] experimentally determined

    the pressure dependence of the crystal oscillator without a polymer coating at pressures up to

    10 MPa in an atmosphere of helium.

  • 19

    2.2.5 In-Line Measurement of Gas Solubility

    In-line measurement techniques were developed from an interest in determining phase

    equilibria during the actual foaming process. Usually, these techniques would incorporate the

    measurement devices in-line with the foaming process. Phase separation, or the solubility

    limit, is then detected by means of sensitive instrumentation or visually.

    2.2.5.1 In-Line Monitoring

    Dey et al. [55] and Zhang et al. [56, 57] reported an in-line technique for measuring the

    gas solubility in various polymers during the foam extrusion process. The apparatus for this

    technique consisted of an extruder with a specially designed optical window, and the flow

    restrictor valve positioned between the die and the end of the extruder. Through this window,

    the occurrence of bubble formation could be observed using a microscope-CCD camera-

    monitor/ recorder system.

    In order to detect the appearance or disappearance of bubbles during phase separation,

    a two-phase, polymer-gas mixture was created by initially using a low pressure in the optical

    window. The pressure was then gradually increased so that the polymer and gas became a

    single-phase solution. This pressure was taken to be the lowest pressure required to keep the

    gas in the solution under specified conditions. Lastly, the solubility was calculated by

    combining this information with the gas flow rate and the melt throughput.

    The parameters affecting the in-line measurement of gas solubility was found to be the

    degree of mixing (single- or twin-screw extruder), the screw rotational speed, and polymer

    throughput. One reported advantage of this in-line technique was that the solubility data could

    be recorded in real-time, and therefore, could account for the dynamic nature of the extrusion

  • 20

    process, the possible role played by the extrusion process in gas dissolution, and bubble

    nucleation in the melt.

    2.2.5.2 In-Line Infrared Sensors

    Near infrared (NIR) spectroscopy is a technique for monitoring the polymer/blowing

    agent mixture during polymeric extrusion foaming processes. Through the use of dual-

    transmission infrared sensors or probes, which transmit NIR light through the polymer running

    in a flow cell, infrared monitoring of the process is achieved. The probes are linked with fibre-

    optic cables to a Fourier transform near-infrared spectrometer (FT-NIR), which records the

    absorption spectra of the melt. The flow cell for NIR measurements is located at the exit of the

    extruder on a side stream of polymer flow taken from the main flow stream. Downstream of

    the flow cell, a gear pump is installed to realize a steady flow rate.

    NIR spectroscopy has been reported to have plenty of advantages, such as remote data

    collection and ease of sample handling. It has also been used for the online measurements of

    polymer composition [58], polymer viscosity [59], and concentration of HCFC in polystyrene

    [60].

    For instance, on-line NIR spectroscopy was utilized by Nagata et al. [61] in measuring

    the Carbon Dioxide (CO2) concentration in molten propylene for CO2 extrusion foaming

    processes. Three different CO2 concentrations and three separate flow rates were used

    experimentally. In order to remove the baseline of the obtained NIR spectra, the wavelet

    transform was employed (the given signals were represented by the linear combinations of

    known functions). They claimed that experiments demonstrated a strong correlation between

    the NIR spectrum and the CO2 raw NIR spectrum; the effects of temperature and flow rate

    were erased. This technique, however, is limited in practice, since the incident light from the

  • 21

    probes would be scattered out if any dispersed material is present in relatively large quantities

    in the melt. The absorbance may then become too weak to be analyzed precisely. Furthermore,

    the calibration curve must be developed whenever the polymer and/or the processing

    conditions are changed.

    Thomas et al. [60] also investigated the ability of NIR spectroscopy to detect bubble

    formation in the die as a function of blowing agent concentration and pressure for the

    PS/HCFC 142b system. When the die pressure was gradually decreased, they observed that

    NIR sensors could detect degassing of the melt. The appearance of bubbles caused scattering

    of the light, which induced a large increase in attenuation at the level of baseline absorbance of

    infrared waves. They also investigated the effect of talc on the performance of NIR

    spectroscopy, and found that NIR analyses were still possible for talc contents

  • 22

    (A1, A2, A3, …) that are detected by the receiving transducer. From the thickness, e (m), and

    the time delay between successive echoes, ∆t (s), the sound velocity, v (ms-1

    ), is determined

    using the following relation:

    2-2

    On the other hand, the attenuation, a (dB/cm), is obtained through the relative amplitude of

    successive echoes:

    2-3

    Sahnoune et al. employed these techniques [63] to measure the thermodynamic

    properties of polystyrene/HCFC 142b mixtures. For phase separation measurements, they

    observed that the velocity of sound decreased by as much as 4.5% from a steady state value as

    the pressure was decreased. This was explained to be due to the phase separation process. The

    attenuation, on the other hand, exhibited a different trend in relation to the thermodynamics

    state of the blowing agent.

    2.2.6 Modified Magnetic Suspension Balance Theoretical Treatments

    Masahiro Ohshima and his coworkers [65] tried to measure the solubility of gas in

    polymer by modifying the MSB. The densities of two polymer/CO2 single-phase solutions,

    poly(ethylene glycol) (PEG)/CO2 and polyethylene (PE)/CO2, were measured at temperatures

    higher than the melting temperature of a polymer under CO2 pressures in the range of 0-15

    MPa using a newly-proposed gravimetric method. A magnetic suspension balance (MSB) was

    used for the density measurement under the high pressure CO2. A thin, disc-shaped platinum

  • 23

    plate was submerged in the polymer/CO2 single-phase solution in the MSB high-pressure cell.

    The weight of the plate was measured while keeping CO2 pressure and temperature in the

    sorption cell at a specified level. Since the buoyancy force exerted on the plate by the

    polymer/CO2 solution reduced the apparent weight of the plate, the density of the

    polymer/CO2 solution could be calculated by subtracting the true weight of the plate from its

    measured weight. Experimental results showed that the density of PE/CO2 solution increased

    with the increase of CO2 pressure; and the density of PEG/CO2 solution decreased with the

    increase of CO2 pressure. To differentiate the effect of CO2 dissolution in polymer from that of

    mechanical pressure, the density of polymer/CO2 solution was compared with the density of

    neat polymer under the given mechanical pressure, which was calculated using the Sanchez-

    Lacombe equation of state and Pressure-Volume-Temperature (PVT) data of the polymer. The

    comparison could elucidate that the dissolution of CO2 in the polymer-reduced densities of

    both PEG/CO2 and PE/CO2 systems. However, this was not the case; the degree of CO2

    induced-density reduction was different between the two polymer/CO2 systems.

    Figure 2-3 Details of MSB for density measurement [65]

  • 24

    When the platinum plate is submerged in polymer/CO2 solution, the measured weight

    of the plate becomes smaller than the true weight of the plate due to a buoyancy force exerted

    on the plate by the polymer/CO2 solution. The buoyancy force is equal to the weight of

    polymer/CO2 solution displaced by the plate, and it is calculated by multiplying the plate

    volume by the density of polymer/CO2 solution. Therefore, knowing the volume and mass of

    the plate a priori, the density of polymer/CO2 solution can be calculated from the buoyancy

    force or the apparent weight of the platinum plate.

    The force balance equation around the plate and the wire is expressed by

    2-4

    where and are the density of polymer/CO2 solution and CO2, respectively;

    is the readout value of the apparent total weight of the plate, wire, and

    measuring load hook at the experimental temperature, T, and CO2 pressure, P, condition;

    is the apparent total weight of the plate, wire and measuring load hook at a reference

    temperature and pressure condition; and are the volume of the platinum plate and that of

    the wire, respectively; is the volume of measuring load hook; is the volume fraction of

    the wire submerged in the solution; d is diameter of the wire connecting the platinum plate to

    the measuring load hook; c is surface tension of polymer/CO2; h is contact angle between the

    wire and the polymer/CO2 solution as shown in Figure 2-3(b); g is the gravitational constant.

    The subscript i, for example di and Vw,i in Figure 2-3(b), indicates that it is the value in the

    case of using the i-th wire.

    Considering that the plate and wire were both made of platinum, the temperature and

    pressure corrections of the volumes, V and Vw, were made using Eq. (2.5):

  • 25

    2-5

    where Vref and Vw,ref are reference volumes of platinum plate and wire; m and E are Poisson’s

    ratio and Young’s modulus of the platinum, respectively. They are given by 0.38 and 1.68

    MPa, respectively. ζ is the coefficient of thermal expansion, which is 9.1x 10-6

    K-1

    .

    The surface tension of polymer/CO2, g, and contact angle, u, were unknown and no

    literature value was available. To eliminate g and u from the balance equation, two wires in

    different diameter, d1 and d2, were used. The density measurements were conducted using each

    wire individually at the same temperature and pressure.

    Assuming that the two wires have the same surface tension, c, and contact angle, h, against the

    polymer, we get the following:

    2-6

    Thus, the density of polymer/CO2, is given by the following:

    2

    -

    7

  • 26

    2.3 Theoretical Treatments

    There are theoretical approaches for explaining and predicting the solubility of gas in a

    polymer. These theories are initially devised from the prediction of the pressure-volume-

    temperature relationship for a pure component. They are then expanded to polymer/solute

    systems. The models presented are based on a lattice fluid model in which each molecule

    occupies r sites (an r-mer) with vacant sites present. It is assumed that there are random

    mixings of r-mers with each other and with the vacant sites. These theoretical models, which

    are summarized and presented in this thesis, are the Flory-Huggins theory [66, 67], Sanchez-

    Lacombe Equation of State (SL-EOS) [45-47], and the Simha-Somcynsky Equation of State

    (SS-EOS) [68].

    The Flory-Huggins (F-H) theory [66, 67] was derived from considering the polymer

    solution as a lattice in which a solvent molecule occupies the same lattice position as the

    polymer segment. It gives information about the solubility and phase relationships, and

    assumes that the volume and enthalpy of mixing are zero. This introduces a reduced Gibbs

    energy parameter, x, to correct the energetic effect of mixing. The x-parameter is taken to be

    independent of composition and temperature. The original F-H theory was modified by Blanks

    and Prausnitz [69], who introduced an entropic contribution to the x-parameter. Nevertheless,

    even though the theory is modified, the F-H theory is still considered inadequate for describing

    polymer solutions. This is because it ignores the equation of state properties of pure

    components and the effect of polymer chain architecture on intermolecular packing.

    The Sanchez and Lacombe Equation of State (S-L EOS) [45-47] is a lattice fluid model

    for pure fluids and mixtures. It requires three pure component parameters to characterize a

    pure fluid and one adjustable binary interaction parameter. When the PVT properties of the

    components at the solubility pressure are acquired, the equilibrium solubility of gas dissolved

  • 27

    into a polymer can also be determined. However, one complication that might arise is that

    there is scarce information on the PVT properties of polymers and interaction parameters. If

    the solubility data is available in a limited range, one can use non-linear regression analysis to

    determine the parameters.

    Panayiotou and Verra (P-V) [70] also obtained an equation of state based on a lattice

    hole theory. The difference between this theory and the S-L EOS is that a constant site volume

    for all r-mers is used, and non-random mixing arising from the molecular interaction is

    introduced. The adjustable binary interaction parameter in the P-V EOS is incorporated as a

    correction for the binary interaction energy. In the S-L EOS, the binary interaction term

    modifies the characteristic pressures. Thus, it has a different physical meaning.

    With much similarity to the F-H or S-L EOS theories, the Simha-Somcynsky (SS)

    model [68] originates from treating molecules as segments on a lattice. In the case of a

    mixture, a lattice contains both species, which are divided into approximately equal-sized

    segments. Nonetheless, unlike the other theories, the SS theory allows for a pressure- and

    temperature-dependent fraction of vacancies or holes that express free-volume within the

    lattice. This accounts for molecular disorder in the lattice model. The equations derived from

    the SS theory incorporate the temperature- and pressure-independent parameters that account

    for intra- and intermolecular interactions within the mixture's components.

    Based on the lattice fluid model, Rodgers and Sanchez [71] determined that adding an

    empirical correlation for the interaction parameter would improve the predictive scope of the

    LF model. With the addition of this correlation using Hansen's three-dimensional solubility

    parameters [72], the LF model was reported to be able to predict solubilities in all types of

    gas/polymer systems without the use of adjustable parameters. In other words, only the pure

    component equation-of-state and solubility parameters are required.

  • 28

    Curro et al. used the Polymer Reference Interaction Site (PRIS) theory [73, 74] to

    compute the sorption of a monatomic gas in a polymer liquid. This theory describes the

    intermolecular packing between polymer chains and solute using the integral equation theory

    of molecular liquids. Also, the chemical potentials of the solute species in the polymer must be

    obtained in order to calculate the sorption of a gas in a polymer liquid.

  • 29

    CHAPTER 3. RESEARCH METHODOLOGY FOR PVT AND

    SOLUBILITY STUDY

    3.1 Introduction

    Gas solubility in polymers can be measured using different techniques, that is,

    gravimetric techniques, including vibrating or oscillating techniques; PVT techniques with the

    pressure decay method; and gas-flow techniques. A brief review of the technique which was

    implemented is given below, followed by the description of a technique we recently developed

    that couples a new gravimetric technique with a PVT visualization technique.

    3.2 Theoretical Background

    Theoretical models to determine the solubility and swelling of the polymer/gas mixture

    such as the Sanchez and Lacombe (SL) [45-47] and the Simha and Somocynsky (SS) EOS

    [68] are all based on the statistical thermodynamic theory.

    The equation of state of the pure component system is written in the following manner for the

    SL EOS:

    3.1

    or the SS EOS:

  • 30

    3.2

    3.3

    In the above relations, ~,~

    ,~

    PT and V~

    are reduced parameters. They are calculated from the

    characteristic reducing parameters P*, T*, V* and as follows:

    3.4

    where y is the fraction of occupied lattice sites, s is the number of segments per chain of molar

    mass, c is the number of external degrees of freedom per chain, and finally V~

    and

    are dimensionless quantities.

    The solubility of gas in the polymer (binary system) can be calculated theoretically using the

    phase equilibrium theory:

    3.5

  • 31

    where is the chemical potential of gas in the gas phase and is the chemical potential of

    gas in polymer/gas mixture phase. Under the phase equilibrium condition, the mass fraction of

    gas in the polymer/gas mixture phase, i.e., the theoretical solubility , can be obtained

    by solving Eq. (3.5).

    In the case of SS-EOS, Eq. (3.6) was used to solve [75]:

    3-6

    And Eq. (3.7) was used to calculate [36, Gli paper]:

    3-7

    where is the molar free energy of the polymer/gas mixture [36 Gli paper].

    3-8

    In the case of SL-EOS, Eq. (3.9) and Eq. (3.10) were used to determine and :

    3-9

  • 32

    3-10

    The swollen volume can be obtained from the following relation:

    3-11

    where S is the gas solubility (g-gas/g-polymer) in the polymer melt which is calculated from

    the EOS, m (g) is the initial weight of the polymer sample, vp,pure (cm3/g) is the specific volume

    of pure polymer.

    3.3 Methodology and Approach

    The details of the measuring procedure using MSB can found in our previous literature

    [76]. The amount of gas which is absorbed by the polymer can be determined by the following

    equation:

    3-12

    where W(P,T) is the weight of the sample at temperature T and pressure P; W(0,T) is the

    weight of the sample at temperature T and vacuum; ρgas is the density of the gas inside the

  • 33

    chamber at temperature T and pressure P; VB, VP, and VS are the volume of the sample holder,

    the volume of pure polymer, and the swollen volume of polymer, respectively, due to gas

    dissolution at temperature T and pressure P.

    By ignoring the polymer’s swollen volume (VS) in Eq. 1, the measured weight gain Wg

    in Eq. (3.12) can be transformed to the apparent solubility (Eq. 3.13), Xapparent, which is less

    than the actual solubility:

    3-13

    As shown in Eqs. 3.12 and 3.13, it is impossible to measure the accurate solubility of

    the gas in the polymer melt by ignoring the swollen volume (VS). However, presently, there is

    a lack of reliable and accurate PVT data for polymer/gas mixtures which are measured

    experimentally. Hence, the swollen volume is typically estimated by an equation of state

    (EOS) which can be applied to a two-component mixture system under equilibrium.

    A general approach that combines the experimental solubility measurement and the

    thermodynamic models was proposed by Li et al. [77] Firstly, a gravimetric method is carried

    out to experimentally measure the gas sorption in a polymer melt (apparent solubility, Xapparent).

    Secondly, the SS-EOS or SL-EOS is applied to calculate the phase equilibrium (theoretical

    solubility, Xtheory) and the swollen volume of polymer, Vs. Thirdly, the theoretically predicted

    swollen volume, Vs can be used to complete the correction on the apparent solubility, Xapparent,

    and then to obtain the actual solubility or corrected solubility, Xcorrected. Meanwhile, the PVT

  • 34

    apparatus is also used to determine the swollen volume [78]. The swollen volume of the

    polymer/gas mixture may also be obtained from the following relation:

    3-14

    where is obtained from measuring the volume of the polymer/gas drop mixture,

    where as is obtained from PVT equation. The corrected solubility, Xcorrected, with the

    buoyancy effect compensation can be obtained using Eq. 3.14:

    3-15

    3.4 Summary

    Based on the magnetic suspension balance, a robust general research approach was

    established for the calculation of solubility, which has been described in detail in the previous

    chapter. In order to obtain accurate solubility data, buoyancy effect must be accounted for,

    which is generated due to the swelling behavior of the polymer in the presence of a gas.

    Experimental and theoretical approaches have been implemented in order to account for the

    swelling behavior of the polymer/gas mixture melt. The phase equilibrium and the PVT

    behavior of the polymer/supercritical fluid are studied in detail with the theoretical approach

    proposed by Dr. Guangming Li and the experimental approach put forward by Dr. Yao Gai

    Gary Li.

    The results showed that the SS-EOS predicted the swelling and hence the solubility of

    carbon dioxide in PLA in close proximity of the theoretical values in comparison to SL-EOS.

  • 35

    CHAPTER 4. SOLUBILITY AND SWELLING BEHAVIOR OF

    PLA IN PRESENCE OF CO₂

    4.1 Introduction

    Based on the magnetic suspension balance, a general approach was established to

    measure the solubility of carbon dioxide. In order to obtain accurate solubility data, inclusion

    of the swelling volume is essential, which is generated from the dissolution of the blowing

    agent in the polymer.

    The theoretical approach for the determination of swollen volume and phase

    equilibrium was built on a variety of technically sound thermodynamic models, such as SS-

    EOS and SL-EOS. Previous work [75] has illustrated some deviation among the

    thermodynamic models in terms of theoretical solubility and swollen volume prediction. In

    this chapter, a systematic investigation is illustrated to investigate the factors that govern gas

    solubility in a polymer. The two models, namely SS-EOS and SL-EOS, are compared with the

    experimental results in terms of their ability to predict theoretical solubility and volume

    swelling.

    Three different grades of PLA are utilized to investigate the effect of D-content on

    solubility and swelling. The effect of varying molecular weight and D-content on the gas

    solubility and swelling volume is also investigated.

  • 36

    4.2 Experimental

    4.2.1 Materials

    Three different grades of Polylactide (PLA) from were used in the

    experiments: PLA 3001D (1.4% D-content), ; PLA 8051D (4.6% D-

    content), ; and PLA 4060D (12% D-content), . The

    PLA was received in the form of pellets from LLC. Carbon dioxide

    (Coleman grade, 99.99% purity) was obtained from BOC Canada.

    4.2.2 PVT Data for PLA

    The PVT data published by Sato for PLA was used to obtain the Tait’s equation [79].

    The same Tait’s equation was used for the three different grades of PLA. Since A Tait’s

    equation represents the PVT behavior of a polymer, therefore, generalizing a Tait’s equation

    for different grades of PLA is not preferable. This makes it harder to observe the effect that

    different variables have on the solubility, swelling volume, and surface tension. The latter

    notion is discussed in depth later in the Chapter.

    4-1

    In the above equation, the temperature, T, is in (°C) and pressure, P, is in (bars). The PVT data

    obtained was also used to derive the characteristic parameters for both SS-EOS and SL-EOS:

    P*, V*, T*. All the characteristic parameters for the PLA grades are listed in Appendix 3.

  • 37

    4.3 Solubility of CO₂ in PLA (Binary System)

    4.3.1 Swelling Behavior of PLA in Presence of CO2

    4.3.1.1 Experimental Setup

    The experimental setup consisted of the following components: a high-pressure

    chamber with a sapphire window for the purpose of visualization; a 2024 x 2024 resolution

    JAI Pulnix TM4100 CL camera with control software; Schneider 4/80 lens and extension

    tubes; a temperature controller (Omega CN132) with thermocouple (Omega RTD); two

    cartridge heaters; an automatic high-precision XY stage with Galil motion controller and

    control board; a manual 1 in. XYZ stage to adjust the position of the light source; a syringe

    pump connected to the gas tank; and a backlight source with a light equalizer/diffuser.

    4.3.1.2 Experimentation

    PLA samples were sliced from a strip obtained from a micro compounder and weighed

    using a precision microbalance. The selected PLA samples were attached to the droplet rod to

    form the sessile drop for each experiment. Swelling measurements for mixtures

    were conducted at two different temperatures, 453 K and 473 K. At each temperature, the

    pressure of CO2 inside the chamber was varied from 6.894 MPa (1000 psi) to 20.684 MPa

    (3000 psi) in 3.447 MPa (500 psi) increments. Each pressure level was maintained for 1.5

    hours to ensure that the equilibrium conditions were established for the polymer/gas solution.

    Equilibrium was considered to be achieved when the total volume of the polymer/gas solution

    no longer changed.

  • 38

    In order to compare the experimental data with the theoretical, SS-EOS and SL-EOS

    were used to determine the theoretical volume swelling ratio. The parameters used to

    determine the theoretical swelling ratio ( ) are shown in Appendix 3.

    4.3.1.3 Pressure and Temperature Effect on Volume Swelling

    It was observed that the volume of the PLA3001D/CO2 mixture increased with an

    increase in pressure as illustrated in Fig. 4.1 and Fig. 4.2. Due to the increase in pressure inside

    the chamber, the density of the CO2 gas increases, hence more CO2 molecular will penetrate

    into the PLA polymer matrix causing more dilation until it reaches the saturation point.

    With a fixed temperature, an increase in pressure causes an increase in the volume of

    the polymer/gas mixture, as well as the volume swelling ratio. Since the solubility of CO2

    increases [80] as the pressure is increased, more CO2 gas is permitted to enter the PLA melt

    matrix, hence an increase in swelling volume was observed. The hydraulic pressure effect due

    to the CO2 was accounted for by using Tait’s equation. The Tait’s equation obtained from the

    PVT apparatus was compared with the Tait’s equation obtained through parameters based on

    the PVT data of PLA by Sato et al. for SS-EOS and SL-EOS [79].

    At isobaric conditions, the volume swelling of the PLA/CO2 mixture tends to decrease

    as illustrated in Figures 4.1-4.3 with an increase in the temperature. As the temperature

    increased, the polymer chains became softer which increased the free volume as well as

    specific volume. The solubility of CO2 in PLA is known to decrease as the temperature

    increases [81]. This means that the diffusion of CO2 out of the polymer increased; hence more

    CO2 is forced out of the polymer/gas mixture compared to what would enter due to the free

    volume at an elevated temperature. In other words, despite the increase in the free volume

  • 39

    within the PLA/ CO2 matrix, CO2 will escape out of the matrix. For instance, at the 13.79 MPa

    pressure level and 453 K; PLA-3001D/CO2 has a volume swelling ratio of 9.32%; whereas at

    473 K, the swelling ratio is 8.88%.

    180 185 190 195 200

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14S

    we

    llin

    g R

    atio

    (%

    )

    Temperature (Celcius)

    1000 psi

    1500 psi

    2000 psi

    2500 psi

    3000 psi

    PLA 3001D

    Figure 4-1: Effect of Temperature variance on Swelling Ratio for PLA3001D

    180 185 190 195 2001.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    Sw

    elli

    ng

    Ra

    tio

    (%

    )

    Temperature (Celcius)

    1000 psi

    1500 psi

    2000 psi

    2500 psi

    3000 psi

    PLA 8051D

    Figure 4-2: Effect of Temperature variance on Swelling Ratio for PLA8051D

  • 40

    180 185 190 195 2001.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    Sw

    elli

    ng

    Ra

    tio

    (%

    )

    Temperature (Celcius)

    1000 psi

    1500 psi

    2000 psi

    2500 psi

    3000 psi

    PLA 4060D

    Figure 4-3: Effect of Temperature variance on Swelling Ratio for PLA4060D

    4.3.1.4 Comparison of the Experimentally Measured Data and

    Theoretically Predicted Data

    The experimental data obtained using the in-house PVT apparatus [82] was compared

    with the swelling volume ratio obtained via EOS. SS-EOS and SL-EOS were implemented in

    order to calculate the theoretical swelling volume ratio [83]. It was evident from Figure 4.4-4.9

    that the SS-EOS provides a more realistic prediction of the swelling volume ratio; whereas the

    SL-EOS exaggerated the swelling volume ratio with respect to the experimental result. This

    has been illustrated in our previous work in detail [82, 84, 85].

  • 41

    1000 1500 2000 2500 3000

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    1.18

    1.20

    1.22

    Sw

    ell

    ing

    Ra

    tio

    Pressure (psi)

    3001D EXP

    3001D SS-EOS

    3001D SL-EOS

    180 oC

    Figure 4-4: Swelling Ratio of PLA3001D at 180 °C with varying pressure

    1000 1500 2000 2500 30001.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    1.18

    1.20

    1.22

    Sw

    ell

    ing

    Ra

    tio

    Pressure (psi)

    8051D EXP

    8051D SS-EOS

    8051D SL-EOS

    180 oC

    Figure 4-5: Swelling Ratio of PLA8051D at 180 °C with varying pressure

  • 42

    1000 1500 2000 2500 30001.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    1.18

    1.20

    1.22

    Sw

    ell

    ing

    Ra

    tio

    Pressure (psi)

    4060D EXP

    4060D SS-EOS

    4060D SL-EOS

    180 oC

    Figure 4-6: Swelling Ratio of PLA4060D at 180 °C with varying pressure

    1000 1500 2000 2500 30001.00

    1.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    1.18

    1.20

    Sw

    ell

    ing

    Ra

    tio

    Pressure (psi)

    3001D EXP

    3001D SS-EOS

    3001D SL-EOS

    200 oC

    Figure 4-7: Swelling Ratio of PLA3001D at 200 °C with varying pressure

  • 43

    1000 1500 2000 2500 30001.00

    1.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    1.18

    1.20

    Sw

    ell

    ing

    Ra

    tio

    Pressure (psi)

    8051D EXP

    8051D SS-EOS

    8051D SL-EOS

    200 oC

    Figure 4-8: Swelling Ratio of PLA8051D at 200 °C with varying pressure

    1000 1500 2000 2500 30001.00

    1.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    1.18

    Sw

    ell

    ing

    Ra

    tio

    Pressure (psi)

    4060D EXP

    4060D SS-EOS

    4060D SL-EOS

    200 oC

    Figure 4-9: Swelling Ratio of PLA4060D at 180 °C with varying pressure

  • 44

    4.3.1.5 Effect of ‘D’ Content/Molecular Weight on Volume Swelling

    It was observed that at 453 K, PLA3001D has a higher volume swelling ratio than PLA

    8051D. For example, at 17.24 MPa, PLA 8051D (D content of 4.6%) has a volume swelling of

    11.29%; whereas PLA 3001D (D content of 1.4%) has a volume swelling of 12.21%. Similarly,

    the swelling ratio of PLA 4060D at 17.24 MPa is 10.81%, which was less than that of PLA

    8051D. At this instant, we cannot conclusively state the D-content’s effect on the volume

    swelling. This is due to the presence of two variables, the D-content and molecular weight. In

    order for us to state any concrete effect of the D-content, we need to experiment with two PLA

    samples with similar, if not the same, molecular weight and different D-content.

    1000 1500 2000 2500 30001.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    Sw

    ell

    ing

    Ra

    tio

    (%

    )

    Pressure (psi)

    3001D EXP

    8051D EXP

    4060D EXP

    180 oC

    Figure 4-10: Effect of varying D-content/Mw on swelling ratio at 180 °C

  • 45

    1000 1500 2000 2500 3000

    1.02

    1.04

    1.06

    1.08

    1.10

    1.12

    1.14

    1.16

    Sw

    ell

    ing

    Ra

    tio

    (%

    )

    Pressure (psi)

    3001D EXP

    8051D EXP

    4060D EXP

    200 oC

    Figure 4-11: Effect of varying D-content/Mw on swelling ratio at 200 °C

    4.3.2 Solubility of CO2 in PLA

    With the prediction of swollen volume from SS-EOS and SL-EOS and the

    experimental data (mentioned in Sec. 4.3.1), the solubility of in PLA at 180 °C and 200

    °C was obtained by utilizing the apparent solubility obtained from the MSB. Methodology

    using the MSB has been discussed at length in Chapter 3.

    4.3.2.1 Pressure and Temperature effect on Solubility of CO₂ in PLA.

    The effect of