Thermodynamics of Polymerization - PowerPoint PPT Presentation

1 / 32
About This Presentation
Title:

Thermodynamics of Polymerization

Description:

Thermodynamics of polymerization determines the position of the equilibrium ... yield a wide variety of polymers including proteins, nylons, and polyesters. ... – PowerPoint PPT presentation

Number of Views:860
Avg rating:3.0/5.0
Slides: 33
Provided by: drrober9
Category:

less

Transcript and Presenter's Notes

Title: Thermodynamics of Polymerization


1
Thermodynamics of Polymerization
  • Thermodynamics of polymerization determines the
    position of the equilibrium between polymer and
    monomer(s). Thus, it impacts both
    polymerization, depolymerization, and
    degradation. The thermodynamics of
    polymerization of most olefins is favorable due
    to the exothermic nature of converting ? bonds
    into ? bonds. For cyclic compounds, the driving
    force for polymerization can vary over a much
    wider range, and one observes a variety of
    behaviors ranging from completely unreactive to
    spontaneously polymerizable under all conditions.
  • The well known thermodynamic expression
  • ?G ?H - T?S
  • yields the basis for understanding
    polymerization/depolymerization behavior.
  • For polymerization to occur (i.e., to be
    thermodynamically feasible), the Gibbs free
    energy of polymerization ?Gp
  • If ?Gp O, then depolymerization will be
    favored.
  • Thus Any factor that lowers the enthalpy, H
    (i.e., makes ?Hp more negative), or raises the
    entropy, S (i.e., makes ?S more positive), will
    shift the equilibrium towards polymerization.
  • Standard enthalpy and entropy changes, ?Hop and
    ?Sop are reported for reactants and products in
    their appropriate standard states. Generally
  • Temperature 25oC 298K
  • Monomer pure, bulk monomer or 1 M solution
  • Polymer solid amorphous or slightly crystalline
  • Polymerization is an association reaction such
    that many monomers associate to form the polymer
  • Regardless of mechanism, there is a large loss in
    the total number of rotational and translation
    degrees of freedom in the total system as the
    monomers associate.
  • This occurrence thus yields a major loss in
    entropy upon polymerization.
  • Thus ?Sp processes.
  • Since depolymerization is almost always
    entropically favored, the ?Hp must then be
    sufficiently negative to composite for the
    unfavorable entropic term. Only then will
    polymerization be thermodynamically favored by
    the resulting negative ?Gp.

2
Thermodynamics of Polymerization (continued)
  • Since most polymerizations are characterized by
    an exothermic propagation reaction and an
    endothermic depropagation reaction, the
    activation energy for the depropagation reaction
    is higher, and its rate increases more with
    increasing temperature compared to the
    propagation reaction. As shown below, this
    results in a ceiling temperature, defined as the
    temperature at which the propagation and
    depropagation reaction rates are exactly equal at
    a given monomer concentration.
  • At long chain lengths, the chain propagation
    reaction
  • is characterized by the following equilibrium
    expression
  • The standard-state enthalpy and entropy of
    polymerization are related to the standard-state
    monomer concentration, Mo (usually neat liquid
    or 1 M solution) as follows

3
Thermodynamics of Polymerization (continued)
  • At equilibrium, ?G 0, and T Tc (assuming that
    ?Hpo and ?Spo are independent of temperature).
  • Or
  • Or
  • At Mc Mo, Tc ?Hpo/?Spo

4
Thermodynamics of Polymerization(continued)
  • Other possible effects on ?Hp
  • loss of resonance stabilization upon
    polymerization
  • changes in bond hybridization
  • changes in H-bonding between M and P states
  • Notice the small changes in the ?Sp values. This
    small variation is attributed to the loss of
    translational entropy which is about constant
    from system to system.
  • For the systems in the table above, the
    equilibrium at 25 oC (i.e., at the standard state
    condition) favors the formation of polymer. This
    may be verified using the equation we examined
    previously.
  • ?Go -RT lnKeq
  • As the temperature increases, the equilibrium
    constant decreases (characteristic of an
    exothermic reaction). When Tc is exceeded, Keq
    becomes less than 1, and thus, depolymerization
    becomes the dominant process.
  • It is very important to note that the Tc concept
    applies only to closed systems at equilibrium.
    For open systems, monomer may volatilize away,
    and thus, depolymerization may occur well below
    the predicted Tc. In fact, few polymers actually
    match their thermal stability as predicted from
    the Tc approach.

5
Experimental Determination of ?Hop and ?Sop
  • ?Hop - by direct calorimetric measurement of
    amount of heat evolved when known amount of the
    monomer is converted to a known amount of
    polymer.
  • or
  • by heats of combustion of M and P which yields
    ?Hof (enthalpy of formation) of M and P. The
    ?Hop is thus obtained by the relationship
  • ?Sop - from the absolute entropies of M and P,
    such that
  • The absolute entropies may be obtained from
    calorimetric measurements of heat capacities of M
    and P over a wide T range, as given by

6
Floor Temperature Behavior
  • Although the vast majority of all polymerizations
    possess negative ?H and ?S, and hence display
    ceiling temperature behavior, four distinct
    possibilities exist as outlined in the table
  • As stated earlier, -?S for polymerization is
    almost universal.
  • Therefore, for olefins and small cyclics,
    polymerization is possible at low temperatures.
  • However, many compounds are never spontaneous
    toward polymer due to ?H (e.g. cyclohexane,
    tetrasubstituted olefins)
  • ?S for polymerization is rare, but known
    examples exist (see below).
  • This rare behavior leads to floor temperature
    behavior or entropy-driven polymerizations.
  • Floor temperature monomers are invariably large
    cyclics containing large atoms from the third row
    and below of the periodic table, that yield
    polymers with highly flexible chains.
  • Examples of monomers possessing a floor
    temperature

7
The Reactivity of Large Molecules
  • In general, when considering growing polymer
    chains (i.e., regardless of the type of
    polymerization mechanism), the reactivity of the
    chain ends will be the primary focus in studying
    the kinetics of the polymerization reaction.
  • Thus, investigations of the kinetics of
    polymerization may be simplified by assuming that
    the rate constant of the chain growth reaction is
    independent of the size of the molecule to which
    the reactive functional group is attached.
  • The validity of the assumption that the rate of
    polymerization is independent of changes in
    molecular size of the reactants may be
    rationalized by observing the behavior of several
    small molecule reactions.
  • For reactions involving homologous series of
    reactants, the rate constant levels off and
    becomes independent of molecular size when n 2.
  • Note that this behavior is quite analogous to
    step-wise polymerization.
  • Further physical rationalizations for the
    underlying assumption include
  • 1. The larger and heavier the molecule, the
    greater the distance between the center of mass
    of the molecule and the reactive chain end.
    Thus, the mobility of the reactive end group in
    solution is much greater than the mobility of the
    molecular center of mass (i.e., the average
    mobility of the total chain). This enhanced
    mobility of the reactive sites yields an
    "encounter rate" which is much greater than that
    predicted by the total molecular mass and is
    approximately independent of the molecular size.
  • 2. In most polymerization reactions, the
    diffusion rate of reactants (i.e., the reactive
    chain ends and monomers) is much more rapid than
    the chemical reaction.

8
Dependence of kp on Molecular Size
9
The Reactivity of Large Molecules(continued)
  • Consider the following kinetic scheme
  • where A is the reactive site, M is a monomer,
    (AM) represents the pair of reactants trapped in
    the "liquid cage", and P is the product polymer.
  • The rate constants k1 and k-1 represent diffusion
    rate constants into and out of the liquid cage,
    while k2 is the rate constant for the chemical
    reaction.
  • Assuming a steady-state concentration of the
    trapped reactants, the rate of polymer formation
    is given by
  • If the diffusion is much more rapid than the
    chemical reaction, such that k-1k2, then
  • Since diffusion into the cage is affected by
    molecular size in the same way as diffusion out
    of the cage, the effect of molecular size cancels
    out of the rate expression.

10
Kinetics of Condensation (Step-Growth)
Polymerization
  • Step-Growth polymerization occurs by consecutive
    reactions in which the degree of polymerization
    and average molecular weight of the polymer
    increase as the reaction proceeds. Usually
    (although not always), the reactions involve the
    elimination of a small molecule (e.g., water).
    Condensation polymerization may be represented by
    the following reactions
  • Monomer Monomer Dimer H2O
  • Monomer Dimer Trimer H2O
  • Monomer Trimer Tetramer H2O
  • Dimer Dimer Tetramer H2O
  • Dimer Trimer Pentamer H2O
  • Trimer Trimer Hexamer H2O
  • Generally, the reactions are reversible, thus the
    eliminated water must be removed if a high
    molecular weight polymer is to be formed.
  • Based on the assumption that the polymerization
    kinetics are independent of molecular size, the
    condensation reactions may all be simplified to
  • COOH HO ? COO H2O
  • Note that there are many types of step-growth
    polymerization reactions which yield a wide
    variety of polymers including proteins, nylons,
    and polyesters. Although similar treatments
    apply to all step-growth polymerizations, this
    section will focus on the kinetics of
    polyesterification.

11
Kinetics of Condensation (Step-Growth)
Polymerization
  • Polyesterification reactions are catalyzed by
    acid and the mechanism is given by
  • Step 1 Fast Equilibrium
  • Step 2 Nucleophilic attack slow, rate
    determining step
  • Step 3 Loss of water
  • Step 4 Regeneration of catalyst
  • In this mechanism, step 1 is a fast equilibrium
    and step 2 is the slow, rate-determining step,
    which follows the rate law
  • By applying the fast equilibrium assumption, the
    rate law becomes

12
Polyesterification Without Acidic Catalyst
  • In this case, the carboxylic acid groups must
    themselves function as the catalyst such that
    H ? COOH and thus,
  • where kexp includes k2, Keq1, and other
    constants of the acid-base equilibrium of the
    carboxylic acid.
  • For a stoichiometric feed ratio of the reactants
    at the beginning of the reaction (t 0),
  • such that COOH OH at all times, and the
    rate equation becomes
  • which upon integration yields

RCOOHo R'OHo 2HOOC-R-COOHo 2HOR'OHo
13
Polyesterification Without Acidic Catalyst
(continued)
  • Consider the fractional conversion of the
    polymerization reaction, P, expressed in terms of
    the fraction of COOH groups (or OH groups) that
    have reacted
  • Substitution into the integrated rate expression
    yields
  • Note that experimental data for esterification
    reactions show that plots of 1/(1-p)2 vs. time
    are linear only after ca. 80 conversion.
  • This behavior has been attributed to the reaction
    medium changing from one of pure reactants to one
    in which the ester product is the solvent.
  • Thus, the true rate constants for condensation
    polymerizations should only be obtained from the
    linear portions of the plots (i.e., the latter
    stages of polymerization).
  • For example, the kinetic plots for the
    polymerization of adipic acid and
    1,10-decamethylene glycol show that at 202oC, the
    third-order rate constant for the uncatalyzed
    polyesterification is k 1.75 x 10-2 (kg/equiv)2
    min-1.

14
Uncatalyzed Polyesterification
15
Acid-Catalyzed Polyesterification
  • Recall that the rate law from the acid catalyzed
    polyesterification is given by
  • If acid is added to the system, then by
    definition of a catalyst, the acid concentraion
    remains constant.
  • Furthermore, at the stoichiometric feed, RCOOH
    OH, the rate expression becomes
  • and in terms of their fractional conversion of
    the reactive groups,
  • Thus a second-order plot of 1/(1-p) vs. time
    yields a linear relationship.
  • Note that experimental data are usually linear
    only beyond ca. 80 conversion.
  • The polyesterification of adipic acid catalyzed
    by p-toluene sulfonic acid shows the the rate
    constant for reaction with 1,10-decamethylene
    glycol at 161 oC and 0.4 p-toluene sulfonic acid
    is k 9.7 x 10-2 (kg/equiv) min-1.
  • Note that this rate constant is significantly
    larger than the noncatalyzed rate constant.

16
Catalyzed Polyesterification
17
Time Dependence of the Degree of Polymerization
  • Consider a polyesterification of bifunctional
    monomers, at a stoichiometric feed ratio.
  • In general, a polymer of (AB)n may be formed in
    the reaction
  • HO-(CO)-R-(CO)-OH HO-R'-OH ?
    HO-(CO)-R-(CO)-O-R'-OH H2O
  • or
  • HO-A-OH H-B-H ? HO-A-B-H H2O
  • where A and B are the structural units
    -(CO)-R-(CO)- and -O-R-O-, respectively.
  • If water is efficiently removed during the
    reaction (which must be done to obtain high
    polymer), then the number of COOH groups present
    is equal to the number of molecules present, at
    all times.
  • where N is the total number of molecules in the
    system and V is the volume.
  • Since the structural units A and B are never
    removed during the reaction, the total number of
    structural units present at all times is constant
    and equal to the number of initial molecules.

18
The Number Average Molecular Weight in
Polycondensation
  • By defining the average degree of polymerization
    of the system, Xn, as the average number of
    structural units per molecule, the relationship
    becomes
  • This relationship is a special case of the
    Carother's Equation.
  • Note that for condensation polymers prepared from
    two reactants, the average number of repeating
    units per molecule is one-half the average degree
    of polymerization.
  • If Mo is the average molecular weight of the
    structural units, then the number average
    molecular weight, Mn may be defined as
  • where Nx is the moles of x-mer of mass Mx, and 18
    is added to account for the unreacted (HOH)
    groups at the ends of each polyester chain.
  • The following figure demonstrates the dependence
    of the number average molecular weight on the
    fractional conversion.
  • Clearly, very high conversions are required in
    order to obtain useful polymers of molecular
    weights greater than 10,000.

19
Mn as a Function of Conversion
20
The Number Average Molecular Weight (continued)
  • Using the kinetic relationships derived earlier,
    a dependence of the molecular weight on reaction
    time may be given by
  • For large reactions times (i.e., for conversions
    greater than 80) the following approximations
    are reasonable.

(uncatalyzed)
(catalyzed)
(uncatalyzed)
(catalyzed)
21
Molecular Weight Distributions of Linear
Condensation Polymers
  • While the average degree of polymerization may be
    determined at any time t using the above
    relationships, it is equally important to know
    the distribution of molecular weights and the
    dependence of this distribution on reaction time.
  • Given a reacting system composed of an A-B type
    monomer, we wish to define the number fraction of
    molecules, at a given conversion, p, which
    contain exactly x structural units. A key
    question becomes
  • What is the probability that a molecule selected
    randomly from the polymerization mixture will
    contain exactly x structural units?
  • p conversion fraction of COOH groups that
    have reacted at time t, and
  • (1-p) fraction of COOH groups remaining at time
    t
  • Thus, the probability of obtaining the molecule
    shown above is given by
  • Prob(x) px-1(1-p)
  • The chance that a randomly selected molecule
    contains exactly x structural units is equal to
    the fraction of molecules composed of x-mers,
    such that

(1)
(2)
22
Molecular Weight Distributions of (continued)
  • where Nx is the number of x-mers in a system of N
    molecules. Thus, the relationship becomes
  • Therefore, we can see that Prob(x) is the mole
    fraction of molecules containing x structural
    units
  • If the evolved water is completely removed during
    the polymerization, then
  • NCOOH N No(1-p)
  • where No is the initial number of molecules.
    Combining eqs. (3) and (4) yields
  • Nx No (1-p)2 px-1
  • As shown in the following Figure, for any given
    conversion, p, low molecular weight polymers
    (i.e., the low values of x) have the highest
    probability of being formed in the total
    distribution.
  • However, the distribution becomes broader and the
    average molecular weight increases as the
    conversion increases.

(3)
(4)
(5)
23
Effect of Conversion on the Number Distribution
of Structural Units
Numerical distribution of the number of
structural units in a condensation polymer for
various conversions.
24
Molecular Weight Distributions of (continued)
  • The number average molecular weight is obtained
    from Prob(x) and the definition of an average.
    Neglecting the weight of water on the terminal
    groups of the condensation polymer, the molecular
    weight of an x-mer is given by
  • Mx-mer xMo
  • where Mo is the average molecular weight of the
    structural units.
  • Thus, we have
  • Now, it can be shown that for p 1,
  • Combining eqs. (8) and (9) yields

(6)
(7)
(8)
(9)
(10)
25
Molecular Weight Distributions of (continued)
  • The weight fraction of x-mers, Wx, may be defined
    as the total weight of molecules containing
    exactly x structural units divided by the total
    weight of polymer
  • The following is true for p 1
  • Combination of eqs. (11) and (12) yields the
    simplification
  • Again, the following Figure shows that this
    distribution of Wx favors low molecular weight
    polymer at low conversions.
  • In addition, the weight average molecular weight,
    Mw , may be defined as
  • In view of eq. (11) we have

(11)
(12)
(13)
(14)
(15)
26
Effect of Conversion on the Weight Distribution
of Structural Units
27
Molecular Weight Distributions of (continued)
  • Combination of eqs. (10) and (15) shows that the
    polydispersity is given by

28
Effect of Non-Stoichiometric Reactant Ratios
  • The highest possible molecular weight is achieved
    in polycondensation reactions using equal
    concentrations of reacting groups.
  • However, it is often desirable to produce a
    specific molecular weight in polymerization.
    This is accomplished by designing the system so
    that unreacted or unreactive end groups are
    incorporated into the polymer. Since molecular
    weight is inversely proportional to the number of
    end groups, this offers a means for molecular
    weight control. We will consider three types of
    systems.
  • Type 1 A system of A-A and B-B monomers in
    which the total number of A functional groups,
    NA, is less than (or equal to) the total number
    of B functional groups, NB. We define a
    stoichiometric imbalance parameter, r, where,
  • In this situation, reaction proceeds until the A
    groups are completely consumed and all the chain
    ends possess unreacted B groups. It is obvious
    that the greater the stoichiometric imbalance,
    the more leftover B groups there will be, and the
    lower the molecular weight.
  • Type 2 A system of A-A and B-B monomers in
    which molecular weight control is achieved by the
    addition of small amounts of a monofunctional
    monomer containing either a single A or B group.
  • Type 3 A system of A-B monomers in which
    molecular weight control is achieved by addition
    of small amounts of mono- and/or polyfunctional
    monomers containing only A or only B groups.

29
Effect of Non-Stoichiometric Reactant Ratios
(continued)
  • For all types of systems, the polymerization can
    be designed to yield the desired through the use
    of the Carothers Equation. The key to this
    method is the concept of number average
    functionality, favg. To compute favg, one must
    first identify which is the minority or deficient
    type of group, A or B. Thus for the case in
    which the A groups are deficient in number,
  • where,
  • N(A)s the number of moles of each type of
    monomer carrying an A group
  • f(A)s functionality of each type of monomer
    carrying an A group
  • Ns the number of moles of each type of
    monomer present (A and B)
  • The Carothers Equation is
  • or
  • where,
  • p the fractional conversion of the deficient
    groups
  • the number average degree of
    polymerization

30
Effect of Non-Stoichiometric Reactant Ratios
(continued)
  • For Type 1 systems, the total number of molecules
    at any time is given by
  • With the degree of polymerization defined as the
    average number of structural units per molecule,
    the average degree of polymerization in terms of
    conversion and feed ratio is now given by
  • Note that at a stoichiometric ratio r 1 the
    above relationship reduces to the previous form
  • In addition, the maximum average degree of
    polymerization possible corresponds to a complete
    conversion of the A groups (i.e., p 1), such
    that

31
Effect of Non-Stoichiometric Reactant Ratios
(continued)
  • For Type 2 systems (with NA NB NB)
  • where, NB number of B groups contributed by a
    monofunctional monomer, and
  • For Type 3 systems (with NA NB)
  • where, NBf number of B groups contributed by a
    polyfunctional monomer, and f functionality of
    polyfunctional monomer, and,
  • All other cases should be treated using favg and
    the general Carothers Equation.

32
Branched and Cross-Linked Condensation Polymers
  • Mono and bifunctional monomers yield linear
    polymers however, if one of the reactants is a
    tri- or multifunctional monomer, then a branched
    or crosslinked polymer will result.
  • The general form of the Carothers equation allows
    the possibility of calculating the conditions
    needed to avoid or ensure the reaching of the gel
    point (i.e., the point of extensive
    crosslinking).
  • Since gelation is presumed to occur when the
    average degree of polymerization becomes
    infinitely large, the Carothers equation reduces
    to
  • where pc is the critical conversion.
  • In practice, it is important to note that this
    approach often overestimates the reaction point
    at which gelation occurs.
  • This overestimation is attributed to the broad
    molecular weight distribution in which the high
    molecular weight molecules reach the gelation
    point before those which have the average value
    of the molecular weight.
Write a Comment
User Comments (0)
About PowerShow.com