Вы находитесь на странице: 1из 20

The Economic and Technical Importance of PVC Stabilizers

427

PVC Stabilizers

Dr. R. Bacalogulu, Dr. M. H. Fisch, Polymer Additives, Crompton Corp., Tarrytown, NY, USA, Dipl. Chem. J. Kaufhold, Dipl. Chem. H. J. Sander*, Polymer Additives, Witco Vinyl Additives GmbH, Lampertheim, Germany

3.1 The Economic and Technical Importance of PVC Stabilizers


Polyvinyl chloride (PVC) was one of the first thermoplastics developed. It has become worldwide a very important bulk plastic over its almost 70 year history. PVC consumption in different geographic areas and expected demand through 2000 are shown in Fig. 3.1.
30

25

World

PVC consumption (Mio. t)

20

15

10

Asia North America

Western Europe

0 1980 1985 1990 1995 2000

Year
Fig. 3.1 PVC consumption from 1980 to 2000

PVC including the various copolymers of vinyl chloride and chlorinated PVC is expected to remain important among thermoplastics because of its compatibility with a large number of other products (e. g., plasticizers, impact modifiers), in contrast to other plastics. Because PVCs mechanical properties can be adjusted over a wide range, yielding everything from rigid to flexible end products, there are many different processing methods and applications for PVC. The toxicological problems which at one time were major obstacles in the manufacture and processing of PVC were solved satisfactorily many years ago [1, 2].
* Recent address: Baerlocher GmbH, Unterschleissheim, Germany

428

PVC Stabilizers

When PVC was first developed, flexible PVC was dominant, but rigid PVC production has increased continually and is now approximately two-thirds of total consumption in many countries. The low thermal stability of PVC is well known. Despite this fact, processing at elevated temperatures is possible by adding specific heat stabilizers that stop the damage. This is one of the main reasons PVC has become a major bulk plastic. The development and production of suitable heat stabilizers followed the production of PVC from the beginning, and remains a precondition for processing and application in the future. Consumption of heat stabilizers in Western Europe was approximately 150,000 tons in 1995 and is estimated to be 170,000 tons by the year 2000 [3]. The consumption of thermal stabilizers for PVC worldwide is estimated to be 450,000 tons [4].

3.2 Thermal Degradation and Stabilization of PVC


3.2.1 Mechanism of PVC Degradation
When PVC is processed at high temperatures, it is degraded by dehydrochlorination, chain scission, and crosslinking of macromolecules. Free hydrogen chloride (HCl) evolves and discoloration of the resin occurs along with important changes in physical and chemical properties. The evolution of HCl takes place by elimination from the polymer backbone; discoloration results from the formation of conjugated polyene sequences of 5 to 30 double bonds (primary reactions). Subsequent reactions of highly reactive conjugated polyenes crosslink or cleave the polymer chain, and form benzene and condensed and/or alkylated benzenes in trace amounts depending on temperature and available oxygen (secondary reactions). 3.2.1.1 Dehydrochlorination of PVC in the Absence of Air (Primary Degradation) Any mechanism of degradation has to explain a series of experimental facts. Structural irregularities, such as tertiary or allylic chlorine atoms, increase the degradation rates measurably at the beginning of the process by a rapid dehydrochlorination that starts the degradation process (Scheme 3.1). Initial rates of degradation are proportional to the content of these irregularities. However, PVC degrades even if these irregularities are eliminated by special polymerization conditions or treatments because of the dehydrochlorination of normal monomer units (random elimination) (Scheme 3.1). It is estimated that after allowing for the differences in concentrations and reaction rates, the rate of random degradation in commercial PVC because of normal chain secondary chlorine atoms has the same order of magnitude as does degradation that results from structural irregularities [5, 6, 7]. Cis-ketoallylic structures, although very reactive in dehydrochlorination (Scheme 3.1), are not present in commercial PVC but can be generated by thermal oxidative processes [7, 8]. After the reactive irregularities initially present are exhausted, degradation continues because of the elimination initiated from normal monomer units [6, 7, 9]. These findings indicate that thermal degradation in PVC is an intrinsic property of this polymer and that changes in synthesis conditions or special treatments that eliminate structural irregularities improve the stability of PVC, but can not completely eliminate its degradation. Stabilizers must be used.

Thermal Degradation and Stabilization of PVC

429

Dehvdrochlorination of structural Dehydrochlorination of structural irregularities Cl Cl

Cl Cl Allylic chlorides Cl

Cl Tertiary chlorides

Cl

O Cis keto allylic chlorides

Dehvdrochlorinationof normal monomer units Dehydrochlorination of normal monomer


Cl Cl Cl

Scheme 3.1

Not all allylic chlorine atoms preexisting and/or formed in the degradation process accelerate degradation. Single double bonds can be identified in degraded PVC by NMR spectroscopy. Double bond sequences, once formed, do not increase by continuation of degradation [6]. There are allylic chlorides with some forms of alkenic double bonds that are stable under degradation conditions [6]. The conjugated polyene sequences are generated in apparently parallel processes from the first moment of degradation. For relatively low conversions, their concentrations increase linearly with time. Zero order rate constants calculated as slopes of these lines decrease exponentially with the increase of the number of double bonds in the sequence [6, 10]. In the thermal degradation of solid PVC, an induction period is observed, and then for higher conversions, the degradation rate increases with time, indicating an autocatalytic process. Hydrogen chloride formed in the degradation increases both the degradation rate and the mean number of double bonds in the polyene sequence, and consequently plays an essential catalytic role in PVC degradation [11, 12, 13]. Some local configurations and conformations of the polymer chain of PVC, such as the conformation GTTG (G for Gauche T for Trans) at the end of certain isotactic sequences, favor degradation. These conformations exhibit a high local mobility relative to the remaining structures in PVC and possess some chlorine atoms with very high degrees of

430

PVC Stabilizers

freedom. Both features make possible the adoption of the conformation enabling the elimination reaction [14, 15]. It follows that dehydrochlorination is possible only for specific local conformations. Along the same line, PVC molecules at the surface of primary particles in the solid state have a much higher conformational mobility than molecules in the interior. PVC degradation consequently is expected to take place predominantly at the surface of primary particles. It is well known that dehydrochlorination of PVC proceeds violently in the presence of Lewis acids such as FeCl3 [111], ZnCl2, [112],AlCl3, [113] SiCl4, GeCl4, SnCl4, BCl3, and GaCl3 [16, 17]. This process is responsible for the very fast discoloration of PVC in the presence of Zn or Sn carboxylates that act as stabilizers till the corresponding halides are formed and fast dehydrochlorination starts. The reaction mechanism of a complex chemical process such as PVC degradation defines the sequence of elementary reactions leading from reactants to products and describes each of these reactions. The mechanism of PVC degradation should explain the above fundamental observations and should also agree with the observations related to PVC stabilization that are discussed later in this chapter. The dehydrochlorination of PVC is a very specific chemical process because of the existence of a long series of alternating CHCl and CH 2 groups in the polymer backbone that makes possible a chain of multiple consecutive eliminations. However, the parallel formation of conjugated polyene sequences containing 1 to 30 double bonds cannot be explained by a simple consecutive elimination. The chain reaction model from Scheme 3.2 can explain this apparent contradiction [6].

Cl

Cl [

Cl ]n -HCl -HCl k Termination Cl [ ][ ] 2 Cl [ ][ 3 ] n-2 ] ] n-1 [ ]n

PVC
Initiation ki HCl catalyzed

I I

-HCl
1k

Propagation k' HCl catalyzed

Termination -HCl -HCl Termination

2k

Propagation k' -HCl HCl catalyzed ........ Propagation k' -HCl -HCl HCl catalyzed

Cl [ [ ] m ] n-m+1

m-1

k Termination
Scheme 3.2

I- active intermediates

Thermal Degradation and Stabilization of PVC

431

The first elimination from a monomer residue from the chain (-CH2-CHCl-) or a structural irregularity such as a tertiary chlorine atom (-CH2-CCl<) forms an active intermediate (I1) or a stable monoalkene. This active intermediate partitions between a stable sequence of two double bonds and a new intermediate (I2). The fate of the second intermediate is analogous to that of the first one and the process continues in this way, generating all the double bond sequences. The concentration of each intermediate is lower than the concentration of the previous one. A simple steady state approximation shows that all polyene sequences in the distribution are formed simultaneously and the apparent rate constants decrease exponentially in agreement with experiment [6]. There is a general consensus that the intermediates in the degradation process are allylic sequences with progressively increased numbers of conjugated double bonds [7, 18] However, the mechanism of initiation, propagation, and termination steps is controversial. An early mechanism hypothesized that the intermediates were allylic radicals [19, 20] (Scheme 3.3).
Cl CH CH2 Cl CH R Cl CH CH Cl CH2 CH CH2 Cl CH Cl CH CH RH Cl CH CH2 Cl CH Cl CH CH CH CH Cl Cl CH2 CH Cl Cl CH2 CH CH2 Cl CH

Cl CH

CH

CH2

Cl CH CH CH CH

Cl

CH

Cl CH
Scheme 3.3

HCl

CH

CH2

. . . .

S c h e m

The major problem with this mechanism is that the chlorine atom is known to be so reactive as to be non-selective. Data on model compounds showed that the allylic hydrogen atom, (>C=CH-CH2-CCl<), has only slightly higher reactivity toward abstraction by a chlorine atom that is free to diffuse throughout the polymer matrix than does a hydrogen atom from a secondary carbon (-CH 2-CCl<). The above mechanism consequently generates primarily isolated double bonds and not the observed sequences of conjugated polyenes, owing to the much higher concentration of hydrogen on secondary carbon atoms [7]. This radical mechanism also fails to explain the very important catalytic role of HCl. In addition, there are no reliable proofs that radicals are intermediates in PVC degradation in the absence of initiators and/or oxygen [18]. An ion pair mechanism (Scheme 3.4) was considered for the initiation step by ionization of chlorine followed by rapid elimination of a proton. A much faster ionization of the activated allylic chlorine formed was considered responsible for the chain reactions [21, 22].

432

PVC Stabilizers

Cl

Cl

Cl

Cl

Cl

Cl

Cl

Cl _ HCl

Cl

+ _ Cl

Cl

Cl

Cl

Cl

Cl

Cl _

Cl HCl

+ _ Cl

Cl

Cl

Cl

Scheme 3.4

However, this mechanism does not explain the previously presented experimental facts. Moreover, it does not postulate any interruption reactions of the degradation chain. The only product of degradation should be a polyene resulting from elimination of all chlorine atoms from the PVC molecule. Consequently, the ion pair mechanism cannot explain the real distribution of polyene sequences as a function of the number of double bonds. This mechanism also does not explain the catalytic role of HCl. Formation enthalpy of Cl2Hcomplexes of 3 to 4 kcal/mol as an intermediate for the catalyzed process in this mechanism does not compensate for the high activation enthalpy of C-Cl ionization (140180 kcal/ mol). A concerted elimination mechanism postulated by A. R. Amer and J. S. Shapiro [23], modified by M. Fisch and R. Bacaloglu [18], and based on experimental data and molecular orbital calculations [6, 18, 24, 25, 26, 27] and additional experimental data from W. H. Starnes and coworkers [28] best explains the experimental facts (Scheme 3.5). The first step is slow formation of a double bond randomly along the polymer chain via a 1,2-unimolecular elimination of HCl through a four-center transition state (Scheme 3.5) or a

Thermal Degradation and Stabilization of PVC


Trans stable

433

H Cl Cl Cl Cl Elimination Cl Cl (HCl catalysis) Cl Cl


Initiation (HCl catalysis)

Cl Cl Cl

PVC Cl Cl Cl

H Cl H Cl
Propagation

Cis-trans isomerization (HCl catalysis)


Trans stable

Cl Cl H

Cis reactive

1,3-Rearrangement (HCl catalysis)


Propagation Trans stable

Cl Cl H Cl Cl Cl

Cl

Cis reactive

Cl H Cl Cl

Cl
Propagation

Cl

H Cl Cl

1,3-Rearrangement (HCl catalysis)

Cl H Cl
Cis reactive

Cl

Trans stable

Cl

Cl

H Cl

etc.

1,3-Rearrangement (HCl catalysis)

Scheme 3.5

six-center transition state in the catalytic presence of HCl or metal chloride dike ZnCl2 (Scheme 3.6). Structural irregularities such as allylic or tertiary chlorine atoms eliminate much faster than do normal secondary chlorines in the chain.

Cl

H Cl H Cl Zn Cl H

Cl H H Cl H H

H Cl Cl H Cl Cl Cl H

Cl H H Cl Cl H H Cl H H

Cl

Zn Cl

Cl

Cl

Cl

Scheme 3.6

The second and third steps of the processes constitute the chain elimination, regardless of the initiation site. In the second step, an HCl molecule is eliminated from a cis -alkyl-allyl chlorine through a six-center transition state, generating a conjugated diene or polyene.

434

PVC Stabilizers

Next is an HCl-catalyzed, 1,3 chlorine rearrangement, generating a new cis -alkyl-allyl chlorine from the conjugated polyene. The second and third processes may continue as long as HCl is still present in the system. Elimination of HCl stops the 1,3-rearrangement of chlorine and consequently, the reaction chain. This explains the very important role of HCl in PVC degradation. All the metallic chlorides that are Lewis acids and can form complexes with chloroalkanes may have a similar role. (Scheme 3.7).

H Cl Cl

H Cl

Cl Cl

Cl Zn Cl Cl Cl

Cl Zn Cl

Scheme 3.7

The analogy with chloroalkanes and chloroalkenes provides important support for the above mechanism [25], as it was shown that activation parameters for the initiation of PVC degradation correlate with the activation parameters of gas phase elimination from secondary chloroalkanes and fall on the same straight line. This suggests that both processes may have the same mechanism: a 1,2-elimination of hydrogen chloride from a synperiplanar conformation involving backbone chlorine through a transition state of four centers. In the same way, it was shown that the chain reactions may have the same mechanism as the elimination from cis -alkyl-allylic chlorides (chloroalkenes): a 1,4-elimination of hydrogen chloride through a transition state of six atoms from a cis allylic structure [25]. Only the cis-allylic system with one double bond that has a relatively low activation enthalpy of dehydrochlorination is reactive in chain propagation in the degradation process. Trans conjugated polyenes are much more stable in the absence of HCl. To dehydrochlorinate, they require a 1,2-elimination at the one end that has a much higher activation enthalpy. This process is similar to a random initiation and is much more likely statistically to occur at a different place in the polymer molecule or on a different polymer molecule entirely. In this way, a chain reaction, once interrupted, does not continue, and the sequences of conjugated polyenes remain as such in the system. Trans conjugated polyenic systems are known to be very stable relative to their cis isomers because of their favorable conformation that allows polymer packing to occur [29].

Thermal Degradation and Stabilization of PVC

435

3.2.1.2 Thermal Oxidative Degradation of PVC During processing, in addition to thermal dehydrochlorination, the polymer is exposed to thermo-oxidative degradation resulting from oxygen; in addition, mechanical stress may cause chain scission. The main feature in thermo-oxidative degradation is dehydrochlorination as in thermal degradation. The presence of oxygen causes the dehydrochlorination process to accelerate, but the discoloration is not as severe as during thermal degradation [5, 30]. The polyene sequences are shorter as a result of the reaction between them and oxygen. The overall activation energy of dehydrochlorination is practically the same for thermal and thermo-oxidative processes [30]. The initial dehydrochlorination proceeds by the same mechanism. The most significant damage during the commercial processing of PVC occurs
Cl CH2 Cl CH2 Cl Cl CH2

Shear Sher
Cl CH2 Cl CH2 Cl CH2 Cl Cl

+
Cl CH2 CH CH2 Cl

Thermal dehydrochlorination

Cl CH2 CH

Cl CH2

Cl

O2

Cl CH

Cl

Cl CH2

. O

CH

Cl CH

Cl

Cl CH2 OH

Cl CH2 CH

Cl O. CH2

Cl

CH

.OH

Cl Cl

.
O Cl CH2 Cl CH2 CH O CH3 CH2 Cl

CH2

CH

Thermo dehydrochlorination

Scheme 3.8

436

PVC Stabilizers

as a result of mechano-chemical reactions in the presence of entrapped oxygen. The shear forces cause chain scission, generating radicals. Thermally-initiated HCl loss is followed by radical oxidation of polyenes to form peroxy radicals and hydroperoxides. Hydroperoxides decompose to generate alkoxy and hydroxy radicals that accelerate the oxidation process and form ketones and acid chlorides [31]. Ketoallylic chlorides initiate the thermal dehydrochlorination process, as described earlier. Although the radical process is much more complex than shown in Scheme 3.8, it is clear that thermo-oxidative degradation does not differ in any essential way from thermal degradation. Dehydrochlorination, the most important process, has the same mechanism in both types of degradations. 3.2.1.3 Secondary Processes in PVC Degradation In PVC degradation at low dehydrochlorination levels, polyene concentrations increase linearly and in parallel with HCl evolution. At higher dehydrochlorination levels, the increase in polyene concentration levels off. The plateau value is lower when degradation temperatures and oxygen pressures are higher. When the plateau is reached, dehydrochlorination level for all double bond sequences show that no consecutive reactions to longer polyenes take place [32]. In the absence of oxygen during the thermal degradation of solid samples, a measurable increase in molecular weight occurs and the molecular weight distribution becomes wider and shifts toward higher values. At some point during degradation, the melt viscosity increases considerably, as can be observed by increasing torque in a Brabender Plasticorder experiments [33]. The crosslinking process is catalyzed by HCl [34]. The most important crosslinking reaction is the Diels-Alder condensation of cisoid trans-trans dienes with other polyenes [35] (Scheme 3.9).
Diels Alder condesation Cl Cl Cl Cl

Cl Cl Benzene formation

Cl Cl

Cl Cl

Cl Cl
Scheme 3.9

Cl Cl

Thermal Degradation and Stabilization of PVC

437

Benzene is formed in very small amounts even at temperatures as low as 160 to 170 C by an intramolecular process [36, 37] (Scheme 3.9). At higher temperatures, substituted benzenes and condensed aromatic hydrocarbons are formed by radical scission of DielsAlder condensation products and radical cyclization of polyenes [38]. In the presence of oxygen, the same reactions take place, but they are more complex because of the processes described earlier. Oxidative scission of the chain predominates and, in general, the molecular weight of the polymer decreases [33].

3.2.2 Heat Stabilization of PVC


The degradation of PVC at elevated temperatures required in thermoplastic processing is an intrinsic characteristic of the polymer and consists of dehydrochlorination, auto-oxidation, mechano-chemical chain scission, crosslinking, and condensation reactions. This degradation must be controlled by the addition of stabilizers. The heat stabilizer must prevent the dehydrochlorination reaction that is the primary process in degradation. There are two ways the stabilizer can act: By reacting with allylic chlorides, the intermediates in the zipper degradation chain. This process should be faster than the chain propagation itself, requiring a very active nucleophile. However. the reactivity of the nucleophile should not be so high as to react with the secondary chlorine of the PVC chain, a process that rapidly exhausts the stabilizer. To be effective, the stabilizer must be associated by complex formation with polymer chlorine atoms, which means it should have a Lewis acid character. This association should take place in regions where the polymer molecules have maximum mobility; in other words, where the conformation of the polymer can favor the degradation processes. Once the degradation starts, it is very fast and can be stopped only if the stabilizer is already associated with the chlorine atom that becomes allylic. These regions are the surfaces of the primary particles of PVC, where the stabilizer molecules are associated with the chlorine atoms. The exceptional effectiveness of such stabilizers at very low concentrations is explained by their entropically favorable position for stopping degradation. In general, these stabilizers, because of their effectiveness, prevent the formation of polyenes longer than four to five double bonds and maintain very good early color in the polymer. These stabilizers are called primary stabilizers. Scavenging the hydrogen chloride generated by degradation is another way to stop the process as the HCl is a catalyst for the chain propagation reaction and the initiation step. However, the diffusion of HCl is quite slow because HCl is associated with the double bond where it was generated. When HCl diffuses away from the reaction center, the zipper degradation reaction stops. The stabilizer should scavenge HCl with high effectiveness to avoid its catalytic effect in chain initiation that starts another zipper dehydrochlorination chain. Because this type of stabilizer cannot prevent the dehydrochlorination in its early stages, polyenes longer than four to five double bonds are formed. PVC discolors and the initial color is not maintained. However, by scavenging HCl, this type of stabilizer avoids the autocatalytic degradation and consequently, overall degradation is much slower. These stabilizers provide very good long term stability and are usually referred to secondary stabilizers.

438

PVC Stabilizers

To have good stabilization of PVC with good early color and long term stability, the two types of stabilizer should be combined appropriately for each particular PVC formulation. Stabilization is complicated by the fact that primary stabilizers become strong Lewis acids by reacting with the HCl that catalyzes the initiation and propagation of PVC degradation. To avoid this, secondary stabilizers should react efficiently with HCl to protect the primary stabilizers. Another possibility is to include compounds called costabilizers in the system. Costabilizers form relatively stable complexes with the chloro derivatives of primary stabilizers (the Lewis acids) and suppress their degradative effect. The most important classes of stabilizers and how they act in PVC stabilization is briefly discussed below. 3.2.2.1 Alkyltin Stabilizers The most commercially important alkyltin derivatives are the mono and dimethyl-, butyland octyltin alkyl thioglycolates, mercaptopropionates and alkyl maleates. All these compounds react with HCl to form the corresponding tin chlorides (Scheme 3.10). However, their stabilization effect does not correlate with the amount of HCl reacted nor with the rate of this reaction [39]. It has been established that in the stabilization of PVC with alkyltin alkyl thioglycolates, alkyl thioglycolates are released by reaction with HCl [40]. These alkyltin compounds consequently function as secondary stabilizers, but this is not the main mechanism of their action.
H H3C Cl H3C S CH2 H3C Sn Cl O C O O H
Scheme 3.10

S CH2 H3C Sn S O C O O CH2 R R C O

R R

CH2

During PVC stabilization with alkyltin alkyl thioglycolates, thioglycolate groups are incorporated into the polymer chain as was determined by 113Sn and 14C tagging [41, 42]. Alkyltin thioglycolates exchange thioglycolate groups with chlorine atoms in reactions with model allylic chlorides and the reactivity in this process parallels the PVC stabilization effect [42]. The main stabilization mechanism of these compounds is consequently substitution of allylic chlorine and they are primary stabilizers (Scheme 3.11). In alkyltin thioglycolates, one thioglycolate group bonds to tin to form a complex and is not active in stabilization. Alkyltin mercaptopropionate groups do not form such complex structures; all mercaptopropionate groups are active in stabilization and their activity is higher compared to the corresponding thioglycolates on a molar basis. In general, monoalkyltins are more reactive than dialkyltin derivatives; however, as a result of the very fast exchange of

Thermal Degradation and Stabilization of PVC


H3C S CH2 H3C Sn Cl O C O R Cl

439

O R O CH2 C C O CH2 S H3C Sn O S H3C Cl Cl

O C O R

S CH2

Scheme 3.11

thioglycolate groups, compositions comprised of at least 40 to 50% mono content exhibit activity equal to that of the monoalkyltin derivatives themselves. Consequently, pure monalkyltin derivatives are not required to obtain maximum stabilization. [43, 44]. Based on their high compatibility with PVC and difficulty of extracting them from PVC blends, it has been postulated that tin stabilizers associate with chlorine atoms at the surface of PVC primary particles which explains their high efficiency in PVC stabilization [43, 44] (Scheme 3.12).
O R CH2 C S H3C Sn O CH2 C O R S H3C O Cl Cl Cl Cl

Cl

PVC primary particles


PVC i 3.12 Scheme ti l

Mercapto compounds generated by the reaction of alkyltin mercapto derivative with HCl add to double bond sequences and by this process, retard PVCs discoloration [45]. Dialkyltin di(alkyl maleates) are able to add in a Diels Alder reaction to polyene sequences and reduce the discoloration of degraded PVC [46]. In both cases, the average polyene sequence length is shortened, thereby shifting the absorption maximum toward the ultraviolet and away from the visible wavelengths. 3.2.2.2 Mixed Metal Stabilizers Metal carboxylates stabilize PVC by either mechanism, depending on the metal. Strongly basic carboxylates derived from metals such as K, Ca, or Ba, which have little or no Lewis acidity are mostly HCl scavengers. Metals such as Zn and Cd, which are stronger Lewis acids and form covalent carboxylates, not only scavenge HCl, but also substitute carboxylate for the allylic chlorine atoms [7] (Scheme 3.13). It has been shown that when the concentration of the metal carboxylates is decreased, the ester group introduced into the backbone by direct substitution can be eliminated by

440

PVC Stabilizers

O R

_ _ O O Ca+2 R C O Cl Cl O

O R + 2 HCl CaCl2 + 2

O O R R O Cl O C O + R C O

R Zn O C

O Zn Cl

R Cl Cl H O C O Cl Cl +
Scheme 3.13

O R C O H

reaction with HCl or by thermal degradation at higher temperatures (reversible blocking mechanism) [7, 48] (Scheme 3.13). IR spectroscopy has shown that Zn carboxylates associate with PVC molecules at the surface of primary particles [49] and are, consequently, very effective in the substitution of allylic chlorine. The synergism between Zn or Cd carboxylates and Ba or Ca carboxylates is attributed to fast exchange reactions between zinc or cadmium chlorides and barium or calcium carboxylates. These reactions regenerate the active zinc or cadmium carboxylates and also avoid the catalytic effect of zinc or cadmium chlorides in PVC degradation (Scheme 3.14). However, it has been shown that the synergistic effect is increased by preheating zinc and calcium stearates together [50]. In this way, a complex zinc stearate is formed that is more active in allylic chlorine substitution (Scheme 3.14).
R O C R _ O _ O O C R + ZnCl2 CaCl2 + O C O Zn O C R R O C R _ _ O +2 O Ca O C R O + C O Zn O C R
Scheme 3.14

+2 Ca

R O C R O

R O O C O Ca+2 _ O O

_ Zn O O C R

Thermal Degradation and Stabilization of PVC

441

The damaging effect of Zn or Cd chlorides in PVC degradation can be considerably reduced by using costabilizers that form metal complexes with them. The most common costabilizers used with solid Cd and Zn carboxylates are polyols [51]. 3.2.2.3 Alkyl Phosphites Stabilizers Dialkyl phosphites have no effect on PVC degradation. Trialkyl phosphites scavenge HCl by an Arbuzov reaction and form dialkyl phosphites. They react also with allylic chlorides, but this process plays a secondary role [27, 52, 114] (Scheme 3.15). When used alone, phosphites are secondary stabilizers, giving good long term stability but poor early color. However, in the presence of zinc di(dialkyl phosphites) (formed from zinc salts and trialkyl phosphites as stabilization proceeds), allylic substitution is considerably increased and becomes the dominant process in PVC stabilization. The early color is very considerably improved [27].
H P OR OR RO _ Cl

H RO

Cl P

H P O

OR OR

RCl OR OR

Cl P RO OR OR

Cl

_ RO

+ OR OR

RCl O
Scheme 3.15

OR OR

3.2.2.4 - Diketones Stabilizers -Diketones and similar compounds with active methylenes react in the presence of Zn carboxylates as catalysts with allylic chlorides generated by PVC degradation by a Calkylation process [54, 114]. The stabilization effect increases with the CH acidity of these compounds [55]. 3.2.2.5 Epoxidized Fatty Acid Esters Stabilizers Epoxides are HCl scavengers and are also reported to be effective in allylic chlorine replacement in the catalytic presence of Zn and Cd salts (Scheme 3.16) [56, 57].

442

PVC Stabilizers

HCl O

HO

Cl

Zn Cl2 O Cl

Cl OZnCl

Cl

O ZnCl2

Scheme 3.16

3.2.2.6 Hydrotalcites Stabilizers Hydrotalcite, a natural mineral, is the hydroxycarbonate of Mg and Al with the exact formula: Mg 6Al2 (OH)16CO3.4H2O. It is constituted from infinite sheets of octahedra of Mg2+ six-fold coordinated to OH -, sharing edges (brucite-like sheets), where Al3+ substitutes for some of the Mg2+ ions. A positive charge is generated in the hydroxyl sheet that is compensated for by CO 32- anions, which lie in the interlayer regions between two sheets. In the free space of these interlayers, there is water of crystallization, associated by hydrogen bonds with both OH- and CO32- anions. Hydrotalcite-like clays with anions of weak acids react with strong acids such as HCl and exchange the anions with Cl-. This reaction allows hydrotalcite-like clays to be used as HCl scavengers in PVC stabilization [58, 59].

3.3 Product Groups and Their Specific Chemical and Application Characteristics
Despite the great variety of thermostabilizers known, only a few have gained industrial importance. According to their chemical composition, they are usually divided into four groups: tin stabilizers, mixed metal carboxylates, lead stabilizers, and metal free stabilizers. Besides the thermostabilizers, there is the important group of costabilizers, which are products with no significant efficiency alone, but which are used together with stabilizers to provide strongly enhanced effects.

3.3.1

Tin Stabilizers

As early as 1936, Yngve recommended not only tetraalkyltin but also alkyltin carboxylates as PVC stabilizers [60]. In 1950, Firestone filed patent applications for organotin mercaptides, which became extremely important in further developments in PVC technology [61]. Organotin compounds with at least one tin-sulfur bond are generally called organotin mercaptides, sulfur-containing tin stabilizers, or thiotins. Organotin salts of carboxylic acids

Product Groups and Their Specific Chemical and Application Characteristics

443

mainly maleic acid or half esters of maleic acid are usually known as organotin carboxylates, and the corresponding stabilizers are sometimes called sulfur-free tin stabilizers. 3.3.1.1 Organotin Mercaptides and Organotin Sulfides Sulfur-containing organotin compounds are among the most efficient and most widely used heat stabilizers. They can be described by the following structures (Scheme 3.17):
R R
Sn- R1 CH3 -/ Methyltin n-C4H9 - Butyltin n-C8H17 - Octyltin
1

Sn

S R

2 2

S R R
1

Sn S S R

R
2

S R

Scheme 3.17 S- R2 -S-CH2 -CO-O-alkyl thioglycolates (alkyl is mostly ethylhexyl or iso-octyl) -S-CH2~ CH2 -CO-O-alkyl mercaptopropionates -S-CH2 -CH2-O- CO alkyl mercaptoethanol esters (so-called reverse esters) -S-alkyl alkylmercaptides -Ssulfides

Among these, the liquid thioglycolates are the predominant group on the market. Mono- and diorganotin mercaptides are often used in combination, because these mixtures improve initial color as well as the long-term heat stability of PVC (synergistically) [62 65]. This is shown in Fig. 3.2.

Fig 3.2 Synergism of mono- and dioctyltin isooctyl thioglycolates exemplified by yellowness Index (YI) as a function of milling time (t) in the dynamic heat stability test on a two-roll mill at 200 C a: dioctyltin-bis(isooctylthioglycolate), b: monooctyltin-tris(isooctylthioglycolate), c: 80% dioctyltin-bis(isooctylthioglycolate) and 20% monooctyltin-tris(isooctylthioglycolate)

444

PVC Stabilizers

The term dialkyltin is also used for mixtures of dialkyltin with smaller amounts of monoalkyltin compounds in order to exploit the synergistic effect of the combination of mono- and dialkyltin. Very efficient, solid stabilizers of a type not mentioned above are derived from -mercaptopropionic acid (Scheme 3.18):
R Sn R CH2 CH2 C O n O n R R Sn S CH2

CH2

O C O

Scheme 3.18

Sulfides of mono- and diorganotin are used in liquid mixtures with certain tin stabilizers, mainly together with thioglycolates and reverse esters. The heat stabilizing effect of these organotin stabilizers depends on the type of mercapto group they contain. These groups are directly involved in the stabilizing reaction; they can either replace the labile chlorine directly according to Scheme 3.11, or add onto polyene sequences after intermediate formation of the mercaptide HSR [45, 66]. Organotin mercaptides are able to react with HCl, to annihilate initiating sites by substitution and also help impede auto-oxidation. The combination of these functions gives the organotin mercaptides exceptional thermostabilizing properties not found in any other class of stabilizer. Details about the mechanism of organotin stabilization can be found in Section 3.2 and in references 3642. The organotin-sulfur stabilizers, especially as mixtures of mono- and dialkyl-tin i-octyl thioglycolates, can be used in all PVC applications where high thermostability is required. They can stabilize all homopolymers, emulsion, suspension, and bulk PVC (E-, S-, MPVC), as well as copolymers of vinyl chloride, graft polymers, polyblends, and postchlorinated PVC (CPVC). One of the most appreciated properties of the whole organotin stabilizer group is the absolute crystal clarity of finished articles, an advantage in the manufacture of rigid PVC packaging and transparent film, bottles, and containers. Organotin thioglycolates are used also in the manufacture of plasticized PVC hoses, profiles, sheet, and transparent top coats or layers. Sulfur-containing organotin stabilizers are not, in general, self-lubricating. Therefore, the high processing temperatures necessary for optimum transparency may cause the hot melt to adhere to the metal surfaces of processing equipment, unless suitable lubricants are added. Adding high-polymeric processing aids based on PMMA to organotin-stabilized PVC imparts better flow properties and improves the surface quality of finished articles, e.g., in calendering films, extruded profiles and sheet, blown bottles, and injection molded fittings.

Product Groups and Their Specific Chemical and Application Characteristics

445

Organotin mercaptides should not be used with cadmium- or lead-containing stabilizers or pigments because the resulting formation of cadmium or lead sulfide can discolor the PVC (sulfur staining). Organotin stabilizers migrate from rigid PVC only very slightly [6769]. This fact, together with favorable toxicological properties, is the basis for the worldwide approval of certain types of methyl- and octyltin isooctylthioglycolates for use in food packaging and potable water pipe [70, 71]. As described in the previous Section (3.2.2.), organotin stabilizers are transformed during processing into the corresponding organotin chlorides. Methyltin chlorides have considerably higher vapor pressure than the analogous butyl- or octyltin compounds. Because of their volatility during processing, the maximum allowed concentration (MAC) for tin (0.1 mg/m3) must be monitored and enforced strictly, especially in open systems such as calendering. Butyltin mercaptides are widely used as stabilizers in the production of films, sheet, injection moldings, floor tiles, and wall coverings. In the US, they are also used for pipe extrusion, and siding with high titanium dioxide content is manufactured almost exclusively with organotin mercaptides. Transparent and translucent articles for outdoor use can be stabilized with sulfur-containing organotin stabilizers only if suitable UV absorbers are also present. A special application for sulfur-containing organotin stabilizers is the production of foamed, rigid PVC profiles and sheet. Reverse esters are mercaptides of mono- or di-methyl or butyltin, based on mercaptoethanol esters of long chain fatty acids. They are especially effective in the manufacture of PVC pipe and siding. Approvals for water pipe exist for certain organotin reverse esters. The liquid estertin iso-octyl thioglycolates (alkyl-O-CO-CH2CH2)2Sn(SCH2COO-i-octyl)2 are also efficient non- toxic stabilizers, but they have not developed significant market share. 3.3.1.2 Organotin Carboxylates Only carboxylates carboxylate with the following structures are of practical interest (Scheme 3.19):
R R
1 1

Sn

O O

CO CO

R R

Scheme 3.19 R1 = Butyl or Octyl, R2 = Alkyl or -CH=CH-CO-O-Alkyl

446

PVC Stabilizers

As already mentioned, organotin derivatives of maleic acid may have an additional stabilizer function, i.e., the Diels-Alder reaction [46]. Their performance is good in all types of suspension, emulsion, and bulk PVC. Optimum results are obtained when they are combined with small amounts of phenolic antioxidants, particularly in plasticized PVC, impact-modified PVC, and PVC copolymers. Because stabilizers containing maleic acid occasionally lead to eye and mucous membrane irritations, there have been many attempts to replace them with other systems. For many years, organotin stabilizers free of maleic acid have been on the market. These consist of a combination of organotin carboxylates, e.g., laurates, and a small amount of an organotin mercaptide [72]. Just as with sulfur-free organotin stabilizers, when used in a suitable formulations, this combination gives rigid PVC high transparency and excellent weathering stability. In the melt, PVC stabilized with alkyltin maleates tends to stick to hot contact areas of the processing equipment. However, this problem can be prevented by suitable lubricants. Organotin carboxylates work especially well in the manufacture of rigid or plasticized PVC articles for outdoor use, as for example, transparent and translucent double-walled panels for greenhouses, siding, and window profiles [74, 75], particularly when pigmented [73]. 3.3.1.3 Recommended Formulations Organotin Mercaptides Glass-clear Rigid Film for Food Packaging S/M-PVC (K-value: 57 to 60) MBS Modifier Processing Aid Di n-octyltin Mercaptide Internal Lubricant External Lubricant Potable Water Pipes (Pressure Pipes) S-PVC (K-value: 68) Ca Carbonate Dimethyl or Dioctyltin Mercaptide Ca Stearate Paraffin Wax PE Wax Oxydized PE Wax Siding Top and Base S-PVC (K-value: 64 to 67) Impact Modifier Ca Stearate Ca Carbonate TiO2 Capstock 100 parts 4.0 6.0 parts 1.3 1.75 parts 0.0 5.0 parts 6.0 10.0 parts Substrate 100 parts 4.0 6.0 parts 1.0 1.4 parts 8.0 12.0 parts 1.0 2.0 parts

100 parts 40. 12 parts 0.5 2.0 parts 1.0 1.5 parts 0.2 1.0 parts 0.3 0.6 parts 100 parts 00. 2.0 parts 0.3 0.5 parts 0.4 1.0 parts 0.6 1.0 parts 0.0 0.6 parts 0.1 0.2 parts

Вам также может понравиться