Sie sind auf Seite 1von 32

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 - TS yields the basis for understanding polymerization/depolymerization behavior. For polymerization to occur (i.e., (i e to be thermodynamically feasible), feasible) the Gibbs free energy of polymerization Gp < O. 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 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

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. In practice:

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 < O for nearly all polymerization processes.

Polymerization is favored at low temperatures: TSp is small Depolymerization is favored at high temperatures: TSp is large

Therefore, thermal instability of polymers results when TSp overrides Hp and thus Gp > O; this thi causes the th system t to t spontaneously t l depolymerize d l i (if kinetic ki ti pathway th exists). it)

PSC 480/740 Kinetics


1

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.
6 5 4
-1

kdp kp[M] kp[M] - kdp Tc


300 350 400 450 500
o

k, sec c

3 2 1 0

550

600

Temperature, K

Each ceiling temperature, Tc, and corresponding equilibrium monomer concentration, [M]c, are related to the enthalpy and entropy of polymerization, defined for monomer at standard 1 M concentration, as follows: G = H o TSo + RT ln Keq At equilibrium, G = 0, and T = Tc.

H o Tc So = RTc ln l Keq

PSC 480/740 Kinetics


2

Thermodynamics of Polymerization
( (continued) )
The equilibrium constant is defined by kp/kdp, or typically, as 1/[M] for standard state defined as 1 M.
H o Tc = So + Rln[M]c

H o So ln[M]c = RTc R At [M]c =1 M, Tc = Hpo/Spo


Specific Examples of Monomer - Polymer Equilibrium kcal/mol Monomer Ethylene Isobutylene Styrene -methyl methyl styrene 2,4,6-trimethyl styrene TFE Hp -21.2 -12.9 -16.7 -8.4 84 -16.7 -37 cal/mol-deg Sp -24 -28 -25.0 -24 24 ---26.8 (H/S) Tc(oC) 610 175 395 66 --1100

Notice the large variation in the -H values.


ethylene th l > isobutylene i b t l - attributed tt ib t d to t steric t i hinderance hi d along l the th polymer l chain, which decreases the exothermicity of the polymerization reaction. ethylene > styrene > -metylstyrene - also due to increasing steric hinderance along the polymer chain. Note, however, that 2,4,6-trimethylstyrene has the same -H value as styrene Clearly, styrene. Clearly the major effect occurs for substituents directly attached to the polymer backbone.

PSC 480/740 Kinetics


3

Thermodynamics of Polymerization
(continued)
Other possible effects on Hp

loss of resonance stabilization upon polymerization changes g in bond hybridization y 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, away and thus, thus depolymerization may occur well below the predicted Tc. In fact, few polymers actually match their thermal stability as predicted from the Tc approach.

PSC 480/740 Kinetics


4

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:
o H p

1 o H o ( P ) H f n f (M ) n

Sop - from the absolute entropies of M and P, such that:

o p

1 o S ( Pn ) S o ( M ) n

The absolute entropies may be obtained from calorimetric measurements of heat capacities p of M and P over a wide T range, g as given by:
T

S (T ) =
o

T dT
0

Cp

PSC 480/740 Kinetics


5

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.

+S for polymerization is rare, but known examples exist (see below).


Therefore for olefins and small cyclics, Therefore, cyclics polymerization is possible at low temperatures. However, many compounds are never spontaneous toward polymer due to +H (e.g. cyclohexane, tetrasubstituted olefins) This rare behavior leads to floor temperature behavior or entropy-driven p y 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.
H + + S + + Behavior Ceiling temperature Always spontaneous toward polymer Floor temperature Never spontaneous toward polymer

Examples of monomers possessing a floor temperature:


H3C H3C H3C Si O O Si CH3 O Si CH3 CH3 S S S S

S S S

Si O H3C CH3

octamethylcyclotetrasiloxane (OMCTS) Hp ~ 0 Sp = 6.7 6 7 J/mole K


o

elemental sulfur Hp = 13.5 kJ/mole Sp = 31 J/mole J/ l oK

PSC 480/740 Kinetics


6

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. 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 g homologous g 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 y of the molecular center of mass ( (i.e., the average g mobility y 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 p than the chemical reaction.

PSC 480/740 Kinetics


7

Dependence of kp on Molecular Size

PSC 480/740 Kinetics


8

The Reactivity of Large Molecules


(continued)
Consider the following kinetic scheme:
k k-1 k2

~~A + M ~~(A+M)

~~(A+M) ( ) P

where A is the reactive site, M is a monomer, ( (A+M) ) represents p the pair p 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. A Assuming i a steady-state d concentration i of f the h trapped d reactants, the h rate of polymer formation is given by:

d [ P] = dt

k1k 2 [ A][ M ] k 1 + k 2

If the diffusion is much more rapid than the chemical reaction, such that k-1>>k2, then:

d [ P] = dt

k1 k 2 [ A][ M ] k 1

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 t of f the th rate t expression. i

PSC 480/740 Kinetics


9

Kinetics of Condensation (Step) Polymerization y Growth)


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. 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 i + H2O Monomer + Dimer Trimer + H2O Monomer + Trimer Tetramer + H2O Dimer + Dimer Tetramer + H2O Di Dimer +T Trimer i Pentamer P t + 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 stepgrowth polymerizations, this section will focus on the kinetics of polyesterification.
10

PSC 480/740 Kinetics

Kinetics of Condensation (Step) Polymerization y Growth)


Polyesterification reactions are catalyzed by acid and the mechanism is given by: Step 1: Fast Equilibrium

R-(C=O)OH +

H+

k1 k-1

R-(C=OH)OH

Step 2: Nucleophilic attack slow, rate determining step

R' OH + R'-OH

R (C OH)OH R-(C=OH
)-R'

k2

R-C(OH) R C(OH)2-(OH (OH )-R' ) R'

Step 3: Loss of water

R-C(OH)2-(OH

k3

R-(C=OH )-O-R' + H2O


k4

Step 4: Regeneration of catalyst

R-(C=OH

)-O-R'

+ H2O

R-(C=O)-O-R' + H+

In this mechanism, mechanism step 1 is a fast equilibrium and step 2 is the slow, slow rate-determining step, which follows the rate law:

-d [COOH] dt

= k2[R(C=OH )OH][R'-OH]

By applying the fast equilibrium assumption, the rate law becomes:

-d[COOH] dt
PSC 480/740 Kinetics

= k2Keq1[R-(C=O)OH][R [R-(C=O)OH][R'-OH][H -OH][H+]

11

Polyesterification Without Acidic y Catalyst


In this case, the carboxylic acid groups must themselves function as the catalyst such that [H+] [COOH] and thus, -d[COOH] = kexp[COOH]2[OH] dt
where h kexp includes i l d k2, Keq1, and d other h constants of f the h acid-base id b equilibrium of the carboxylic acid.

For a stoichiometric feed ratio of the reactants at the beginning of the reaction (t = 0),

[RCOOH]o = [R'OH]o = 2[HOOC-R-COOH]o = 2[HOR'OH]o


such that [COOH] = [OH] at all times, and the rate equation becomes

-d[COOH] dt

= kexp[COOH]3

which upon integration yields:

1 1 = + 2kexpt 2 [COOH]2 [COOH ]o


PSC 480/740 Kinetics
12

Polyesterification Without Acidic y ( (continued) ) Catalyst


Consider the fractional conversion of the polymerization reaction, P, expressed in terms of the fraction of COOH groups (or OH groups) that have reacted:

XCOOH = p = 1 -

[COOH] or [COOH]o

[COOH] = [COOH]o(1-p)

Substitution into the integrated rate expression yields:

1 2 = 1 + 2 [ ] COOH o k exp t (1-p)2

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 d is i the h solvent. l 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 1 75 x 10-2 (kg/equiv) (kg/eq i )2 min-1.

PSC 480/740 Kinetics


13

Uncatalyzed Polyesterification

PSC 480/740 Kinetics


14

Acid-Catalyzed Polyesterification
Recall that the rate law from the acid catalyzed polyesterification is given by:

-d[COOH] dt

= k2Keq1[R-(C=O)OH][R'-OH][H+]

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:

-d[COOH] dt

= kexp[COOH]2

and in terms of their fractional conversion of the reactive groups,

1 1p 1-p

= 1 + kexp[COOH]ot

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. i The polyesterification of adipic acid catalyzed by p-toluene sulfonic acid shows the the rate constant for reaction with 1,10decamethylene 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.
15

PSC 480/740 Kinetics

Catalyzed Polyesterification

PSC 480/740 Kinetics


16

Time Dependence of the Degree of y Polymerization


Consider a polyesterification of bifunctional monomers, at a stoichiometric feed ratio. In g general, ,ap polymer y of (AB)n ( ) may y be formed in the reaction:
HO-(C=O)-R-(C=O)-OH + HO-R'-OH HO-(C=O)-R-(C=O)-O-R'-OH + H2O

or
HO-A-OH HO A OH + H H-B-H B H HO HO-A-B-H A B H + H2O

where A and B are the structural units -(C=O)-R-(C=O)- and -O-R-O-, respectively.

If wate water is s efficiently e c e t y removed e oved during du g the t e reaction eact o (which (w c must ust be done to obtain high polymer), then the number of COOH groups present is equal to the number of molecules present, at all times. N = [COOH ] = [COOH ]o (1 p ) V

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.
N structural units V = [COOH ]o

PSC 480/740 Kinetics


17

The Number Average Molecular g in Polycondensation y Weight


By defining the average degree of polymerization of the system, DP, as the average number of structural units per molecule, the relationship becomes:
[COOH ]o DP = [COOH ] 1 = 1 P

This relationship is known as 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 that make up a repeat unit (i.e., (i e the average monomer molecular weight), then the number average molecular weight, Mn may be defined as:
Mn

n m n
x x

= DP M o + 18 =

Mo + 18 1 P

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 f l polymers l of f molecular l l weights i ht greater t than th 10,000. 10 000

PSC 480/740 Kinetics


18

Mn as a Function of Conversion

PSC 480/740 Kinetics


19

The Number Average Molecular g ( (continued) ) Weight


Using the kinetic relationships derived earlier, a dependence of the molecular weight on reaction time may be given by:

Mn

2 = M o 1 + 2[COOH ]o kt

12

+ 18

(uncatalyzed)

Mn

= M o (1 + [COOH ]o kt ) + 18

(catalyzed)

For large reactions times (i.e., for conversions greater than 80%) the following approximations are reasonable.

Mn

M o [COOH ]o 2kt M o [COOH ]o kt

(uncatalyzed)

Mn

(catalyzed)

PSC 480/740 Kinetics


20

Molecular Weight Distributions of y 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. time A key question becomes:

What is the probability that a molecule selected randomly from the polymerization mixture will contain exactly x structural units? P = fraction of COOH groups that have reacted at time = t, and (1-P) = fraction of COOH groups remaining at time = t P = probability that at time = t a given COOH group will have reacted. (1-P) = probability that at time = t a given COOH group will not have reacted.

Recall that:

In other words:

By considering the linear combinations of x structural units, the probability that the randomly selected molecule will contain exactly x structural units is given by: Prob(x) = Px-1(1-P)

where (1-P) may be considered as the probability that the chain end will not have reacted, and P considered as the probability that the chain will have grown from x reactions (-1 for the end group).

PSC 480/740 Kinetics


21

Molecular Weight Distributions of


(continued)
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

Nx N

= Prob(x)

where Nx is the number of x-mers in a system of N molecules. Thus, the relationship becomes:

Nx

= NP x 1 (1 P )

If the evolved water is completely removed during the polymerization, then NCOOH = N = No(1-P)

where No is the initial number of molecules. Thus, 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. , the distribution becomes broader and the average g However, molecular weight increases as the conversion increases.

PSC 480/740 Kinetics


22

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.

PSC 480/740 Kinetics


23

Molecular Weight Distributions of (continued) ( )


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, the number average molecular weight may be obtained from the probability function and the usual definition of an arithmetic average as given by:
No No

Mn

x 1 xM Prob ( x ) = M ( 1 P ) xP o o x =1 x =1

PSC 480/740 Kinetics


24

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. 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:

Mw

wx mx w
x

WM
x =1

No

x mer

= M o xWx
x =1

No

where, wx is the weight of x-mer of mass mx.

In terms of the probability function, the weight fraction of x xmers and the weight average molecular weight are given by:

Wx

= x(1 P) 2 P x 1

and thus:

Mw
PSC 480/740 Kinetics

= (1 P ) M o x 2 P x 1
2 x =1

No

25

Effect of Conversion on the Weight Distribution of Structural Units

PSC 480/740 Kinetics


26

Molecular Weight Distributions of (continued) ( )


Since P < 1, the summations appearing in the expressions for the number and weight average molecular weights may be evaluated as:

xP
x =1

No

x 1

1 = 1 P =

(for M n )

x 2 P x1

1+ P (1 P )3

(for M w )

Thus, the expressions for the average molecular weights become:

Mn

Mo = 1 P

and

Mw

1+ P = Mo 1 P

Therefore simple combination shows that the weight average molecule weight is always greater than the number average molecular weight, such that:

Mw

= M n (1 + P)

and for linear condensation polymers, the polydispersity is given by: g y Mw = 1+ P Mn


27

PSC 480/740 Kinetics

Effect of Non-Stoichiometric Reactant Ratios


In general, it is only possible to produce high molecular weight condensation polymers when equal concentrations of reactive functional groups are maintained throughout the course of the reaction.

The reaction must be initiated with stoichiometric ratios and the reaction mixture must be free of impurities that contain the same functional groups.

For example, consider the A-A, B-B polymerization reaction: A~~~~A + B~~~~B A~~~~ab~~~~B + ab

where A and B are the reactive functional groups.

Furthermore, consider the feed ratio to contain an excess of monomer B, such that:
r

N Ao N Bo

< 1

Note that NAo = twice the number of initial AA molecules

Since A and B must react on a 1:1 basis, and polymerization will be complete when all of A is gone, consider the conversion P = XA.

PSC 480/740 Kinetics


28

Effect of Non-Stoichiometric (continued) ) Reactant Ratios (


Thus, for bifunctional monomers, the total number of molecules at any time is given by: 1 r N = N Ao 1 P + 2r 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:

DP =

1+ r (1 + r ) 2rP

Note that at a stoichiometric ratio r = 1 the above relationship reduces to the previous form:

DP =

1 1 P

In addition, the maximum average degree of polymerization possible ibl corresponds d to t a complete l t conversion i of f the th A groups (i.e., P = 1), such that:

DPmax

1+ r 1 r

and

(M )

n max

1+ r = Mo 1 r
29

PSC 480/740 Kinetics

Branched and Cross-Linked y Condensation Polymers


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 degree of crosslinking or branching in these systems may be controlled by the relative amounts of multifunctional and difunctional monomers.

In a system containing equivalent numbers of A and B groups, the number of monomer units present initially is No and the corresponding number of functional groups is NoFavg,

where Favg is the average functionality of the monomers as defined by:

Favg

N F N
i i

where Ni is the number of molecules of monomer i with functionality Fi. For example, if a system contains 2 mol of a triol and 3 mol of a diacid, diacid then Favg = [2(3)+3(2)]/(2+3) = 12/5 or 2.4. 24 If N is the number of molecules after the reaction has occurred, then 2(No - N) is the total number of functional groups (e.g., COOH's and OH's) that have reacted.

PSC 480/740 Kinetics


30

Branched and Cross-Linked y (continued) ( ) Condensation Polymers


Therefore, the extent of reaction P may also be defined as the fraction of total functional groups reacted, as given by:

P =

2( N o N ) N o Favg

Since the degree of polymerization is given by:

DP =
combination yields:

No N

DP =

2 2 PFavg

which may be rearranged to yield the familiar (and general) form of the Carothers equation:

P =

2 2 Favg DPFavg

which relates the extent of reaction and degree of polymerization l i ti to t the th average functionality f ti lit of f the th system. t

PSC 480/740 Kinetics


31

Branched and Cross-Linked y ( (continued) ) Condensation Polymers


This 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 above relationship reduces to:

Pc

2 Favg

where h Pc is i the th critical iti l conversion. i 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.

PSC 480/740 Kinetics


32

Das könnte Ihnen auch gefallen