Вы находитесь на странице: 1из 14

Combined Relaxation Study of Poly(vinyl chloride) Blends with Chlorinated Poly(ethylene), Hydroxyl-Terminated Poly(butadiene), and Ethylene-Propylene-Diene Terpolymer

S. STOEVA, 1 P. KARTALOV, 2 K. JANKOVA 1


1

University Prof. Assen Zlatarov, 8010 Bourgas, Bulgaria Plovdiv University, 4000 Plovdiv, Bulgaria

Received 19 April 1995; revised 19 December 1997; accepted 22 December 1997

ABSTRACT: A number of blends based on suspension poly(vinyl chloride) and stabilizers

with poly(ethylene) chlorinated in a uidized-bed reactor containing 21.8% chlorine, hydroxyl-terminated poly(butadiene), and ethylene-propylene-diene terpolymer have been studied using such methods as thermally stimulated current depolarization and dynamic mechanical analysis. Some dielectric and thermodynamic parameters ( tmax , to , Ea , DH*, DS* , DG*, meff ) have been determined. Blends containing randomly E chlorinated poly(ethylene) exhibited dipoledipole interactions between the macromolecules of poly(vinyl chloride) which decreased at the expense of the long sequences of nonchlorinated methylene groups. Simultaneously, an increased physical interaction between poly(vinyl chloride) and the additives was observed in blends containing chlorinated poly(ethylene) and/or hydroxyl-terminated poly(butadiene), and ethylene-propylene-diene terpolymer. On the basis of the data of dynamic mechanical analysis obtained a heterogeneous structure of the blends is suggested. The development of a boundary interfacial layer with a proper region of relaxation proves the formation of compatible structures between the components. 1998 John Wiley & Sons, Inc. J Polym
Sci B: Polym Phys 36: 15951608, 1998

Keywords: PVC blends; thermally stimulated current depolarization; dynamic mechanical analysis; physical quantities of relaxation transitions

INTRODUCTION
The excellent mechanical characteristics of poly(vinyl chloride) (PVC) and its accessibility provide wide utilization of PVC in industry. Unfortunately, solid PVC has some drawbacks (e.g., poor processability and low impact strength).1 Modication of PVC with chlorinated polyethylene (CPE) results in polymer materials suitable for the manufacture of sheets, coatings, shaped articles of improved elasticity, and impact strength.2 The usual chlorine content in these materials varCorrespondence to: S. Stoeva
Journal of Polymer Science: Part B: Polymer Physics, Vol. 36, 15951608 (1998) 1998 John Wiley & Sons, Inc. CCC 0887-6266/98/101595-14

ies within 2540% 3 and content of 36% is known to provide for the best impact strength.4 There is a considerable body of literature concerning the changes in properties and compatibility of PVC-CPE blends.4 11 Siegmann and Hiltner 4 reported the incompatibility of the components in PVC-CPE blends. Locke and Paul 5 and Zelinger 6 studied PVC-CPE blends of chlorine content 36, 42, and 48%, respectively, by using dynamic mechanical measurements. Shur and Ranbys 7 study of the gas permeation of similar PVC-CPE blends found that PVC-CPE (36% Cl) and PVC-CPE (42% Cl) blends formed largely incompatible structures, whereas PVC-CPE blends containing 48% Cl were compatible to some extent, only. Locke and Paul 5 explained the im1595

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1596

STOEVA ET AL.

provement in stress-strain properties of polymer blends containing CPE and PVC by the improved interfacial adhesion resulting from the interaction of CPE with the blend components, rather than by the increasing mutual compatibility of the components. On the other hand, del Val et al.8 conducting thermally stimulated creep measurements of PVC blends with CPE of various chlorine distributions observed that some degree of compatibility existed. It depended on the mode of chlorine atoms distribution along the macromolecular chain. The compatibility of PVC and CPE, however, was found to depend not only on the chlorine content and its distribution, but also on the amount of CPE in the blends.9 According to Varkalis et al.10 weak interactions between polymer components took place in PVC-CPE blends provided CPE content did not exceed 20%. Doube and Walsh 11 studied the thermodynamic compatibility of PVC blends with chlorinated in solution poly(ethylene) Hypalon 48. The authors found that the two polymers were compatible within a denite concentration range. PVC blends with CPE of industrial application containing 36% Cl have low impact strength at low temperatures. This is due to the comparatively high Tg value of CPE. The impact strength of PVC-CPE blends was increased considerably by adding modiers with elastomeric properties. The introduction of the following modiers has been reported: A copolymer of ethylene with vinyl acetate, 12 various butadiene and butadiene-acrylonitrile oligomers, 13 acrylic resin, 14 ethylene-propylene-diene terpolymer.15 The elastomeric additives produced better adhesion between the components at the boundary interfacial surface. The PVC-CPE blends mentioned above are mainly concerned with CPE prepared by chlorination in solution or suspension. There is little information on the application of poly(ethylene) (PE) chlorinated in powdery state.16,17 Frey et al.16 prepared a thermoplastic composition of PVC and modied PE. The latter was obtained by gasphase chlorination of commercial PE in a uidized-bed reactor and contained 36.5% Cl. Highmolecular PE was agitated with chlorine in a tubular reactor. The product of 3342% chlorine content was introduced into PVC blends, which improved their impact strength.17 Chlorinated poly(ethylene) of molecular weight ca. 100,000 prepared in a uidized-bed reactor was introduced into PVC. According to our previous studies, 18 the resulting blends had an improved impact strength,

provided the chlorine content in CPE was 21 22%. The high-impact strength of compositions at both the ambient and lower temperatures were prepared by treating the PVC-CPE blend of 21.8% Cl content by addition of 3% hydroxyl-terminated poly ( butadiene ) and / or ethylene-propylene-2ethylidene-5-norbornene terpolymer.19 The aim of the present work is to study the effect of poly(ethylene) chlorinated in a uidizedbed reactor and combined with elastomeric additives on the main kinetic, dynamic, and dielectric properties of PVC in its transition from glassy into highly elastic state.

EXPERIMENTAL
Materials A suspension PVC with Ficentcher constant of V 68.1, Mn 86,000, bulk mass of 540 kg/m 3 and moisture content of 0.03% was used in all experiments. The chlorinated polyethylene (CPE) was prepared by chlorination of high-density poly(ethylV ene) with Mh 96,500 and particle size of 0.125 0.250 mm in a uidized bed with a gaseous mixture of chlorine and nitrogen at temperatures of 293353 K. The combined chlorine content was determined by burning out a 1525 mg sample according to the Schoniger method. This was fol lowed by potentiometric titration with 0.1 N AgNO3 solution. The PVC-CPE blends with CPE containing 21.8% Cl were prepared. Melting point of CPE was 394.5 K, glass transition temperature 289.2 K, and degree of crystallinity 45.7%. The hydroxyl - terminated poly ( butadiene ) (HTPB) was prepared by radical polymerization in 2-propanol in the presence of 60% aqueous solution of H2O2 .20 The characteristics of HTPB were: V Mn 3150, viscosity at 298.2 K, 41.5 Pa.s, Tg 182.2 K, functionality 2.2. The ethylene - propylene - diene terpolymer (EPDM) was Vistalon 2504 (Exxon) and contained 56.6% ethylene, 39.0% propylene, and 4.4% 2-ethylidene-5-norbornene. The corresponding viscoelastic characteristics were: viscosity (Moony) 40, tensile strength 13 MPa, elongation at break 1455%. A series of PVC blends and polyblends was prepared. The following additives were employed and their amount was calculated with respect to PVC: CPE 21.8% Cl, 10 wt%, HTPB 3 wt%, EPDM 5 wt%. A combination of stabilizers such as calcium

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

COMBINED RELAXATION STUDY OF PVC BLENDS

1597

stearate (1.0 wt%), modied dibutyltin maleate BT-22 (4.0 wt%), pentaerythrito-tetrakis[3-(3,5ditert-butyl-4-hydroxyphenyl)] propionate-Irganox 1010 (0.25 wt%) was added to each polymer blend. The ingredients were rst dry-mixed and then homogenized on rolls at 445451 K for 8 10 min. Sheets were then pressed at 453 K under pressure of 3.0 MPa and an average cooling rate of 25 /min.

tively, are due to the cooperative character of dipole-segmental relaxation involving N kinetic units. Values of N can be determined from the dependence of the frequency factor no on temperature, which is expressed as: 22,23
01 no t o n * (Ea /RT ) f 01 .[1/( f 0 1)!] (4) 0 f 1 N

METHODS
Thermal Depolarization Analysis (TDA) 1. If an electric eld is applied to a polymer material, residual polarization is known to be induced. The magnitude of the thermodepolarization current related to a unit of area gives the current density i. By applying the experimental values for i, one can calculate the residual polarization P and the relaxation time t from eqs. (1) and (2), providing that t t 0 and i r 0 at t:
t

where n * denotes the frequency of heat vibrations 0 of kinetic units and amounts to no 10 12 s 01 , and f is a transition parameter accounting for the variation of N values. If ln to f (Ea ) function is linear, the so-called compensation law is considered valid: 24,25

to tc exp[0E( t )/krTc ]
Then eq. (3) can be expressed as:

(5)

t(T ) tcexp

E( t ) k

1 1 0 T Tc

(6)

P(t)
t0

i(t) dt

(1) (2)

t(T ) P(T )/i(T )

From the plot ln t f (10 3 /T ) one can dene the temperature intervals corresponding to a linear relationship. Then using the equation of Arrhenius-Frenkel (3) the activation energy Ea and pre-exponential factor to for each interval can be found:

where tc and Tc are known to represent compensation time and compensation temperature, respectively. According to the second approach, relaxation time is associated with the life-time of the activated complex.26 This approach makes use of the Eyring equation, 27 equation of Arrhenius-Frenkel, and the thermodynamic dependences of DH*, DG*, and DS* : E

DH* Ea 0 RT DG* DH* 0 T DS* DS* R[ln h 0 ln( torxrerkrT )] E

(7) (8) (9)

t to expEa ( t )/RT

(3)

According to Lushcheikin 21 to ranges from 10 013 to 10 010 s and from 10 025 to 10 018 s for dipolegroup and dipole-segmental relaxation processes, respectively. The corresponding values of Ea are within 50120 and 150800 kJ/mol. By introducing to values of the dipole-segmental relaxation processes and applying two different approaches one can estimate: (i) The number of kinetic units N that determine the length of the chain segment taking part in the relaxation process; (ii) the variation in enthalpy ( DH*), entropy ( DS* ), and free E Gibbs energy ( DG*) associated with the transitions of a given kinetic unit from polarized to depolarized state. The rst approach takes into account that the low to values or high values of no 1/ to , respec-

where h is Plancks constant, x is transmission coefcient, whose value for dipole-segmental processes is 0.1, and e the Neper is number. Applying the standard procedure described elsewhere 28 one can calculate DH*, DG*, and DS* values for the E corresponding temperature intervals DT. 2. The activation energy Ea can be calculated by means of several methods. (a) Chens method 29 provides the calculation of activation energies Ea ( v ), Ea ( t ), and I I H Ea ( d ), referring to the total temperature range and the low- and high-temperature sides of the interval, respectively. It involves preliminary estimation of the Chen

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1598

STOEVA ET AL.

criterion, the so-called form-factor XFF , according to eq. (10):

H H H XFF T2 0 Tmax /T2 0 T1

(10)

where T1 and T2 are the half-intensity temperatures on the low- and high-tempera ture side of the peak, respectively [ i(T1 ) 2 ) i(Tmax )/2 imax /2]. The value of i(T XFF obtained predetermines the kinetics of the corresponding relaxation process which is monomolecular for XFF 0.40 and bimolecular for XFF 0.55. (b) Using Turnhouts 3 0 or Grosweiners method,31 one can calculate the activation energy Ea (Th) and Ea (Gw) of the low-temperature part of the plot i f (T ) only. (c) Method according to the literature.29 From the dependence: Ea (imax ) RrT 2 /brtmax max (11)

atom groups and is determined by utilization of some literature data.34 Assuming, however, that Pmax is provided by a nondipole constituent alone, i.e., by current carriers of elementary electric charge qo 1.6 1 10 019 C, the surface density of current carriers can be evaluated from the expression n Pmax /qo . Now the volume density is nally obtained according to the expression n * n n . The n/n * ratio gives the number of atomic I complexes per current carrier (a.c./c.c.).

Dynamic Mechanical Analysis (DMA) DMA is an informative method concerning the mobility of macromolecular segments, as well as the supermolecular formations.35 Dynamic mechanical tests are rather sensitive to structural heterogeneity and to the morphology of multiphase systems, such as crystalline-amorphous polymers, polymer blends, and compositions.36 The main dynamic mechanical characteristics: shear modulus G , loss modulus G , and factor of mechanical losses tan d were measured by free damping oscillation of the reverse torsion pendulum. Then applying a specic form of the Arrhenius-Frenkel equation: Ea RrTmaxrln( tmax / to ) (14)

where b dT/dt is the rate of linear heating of the sample during depolarization, Ea for the whole temperature range of the plot i f (T ) can be calculated. 3. The values of the increment of specic dielectric permeability D1 are calculated by the equation:

D1 1stat 0 1 Pmax / 1orEp

(12)

where 1o 8.85 1 10 012 F/m is the absolute dielectric permeability in vacuum, Ep 5 1 10 5 V/m is the applied electric eld to induce polarization at 383 K. 4. The effective dipole moment meff of the corresponding atomic complexes is evaluated by means of a procedure described elsewhere 29 34 which is based on the Langevin-Debye equation: 33

Ea values for all maxima of the tan d versus temperature dependence can be calculated. It is assumed that to 5 1 10 012 s represents the period of vibration of free chain segments with respect to their momentary equilibrium position.37 The value of tmax was found to be 1.6 s, as calculated from the dependence vrtmax C, where C 10 for chain segments and large kinetic atomic groups 34 and v 2pn with n 1 Hz.

meff 3rkrTprPmax /Eprn

(13)

Analytical Procedures The thermodepolarization analysis of the blends studied was carried out according to a standard procedure.32 Specimen dimensions were: length 40 mm, width 30 mm, thickness 1.1 mm. Aluminium electrodes were applied to the samples which were subsequently placed into a heating chamber and subjected to polarization in a d.c. electric eld of Ep 5 1 10 5 V/m at 383 K for 2 h. Samples then, were cooled to room temperature in the same electric eld for 2 h and stored at room tem-

where k 1.38 1 10 023 J.K 01 is Boltzmanns constant; Tp is the temperature of polarization; Ep 5 1 10 5 V/m; Pmax is the experimentally determined maximum residual polarization assuming that it is provided by the dipole component of atomic groups in the corresponding complex; n is the number of atomic complexes per volume unit. Van der Vaals volume V of each atomic complex represents the summarized volumes of all participating

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

COMBINED RELAXATION STUDY OF PVC BLENDS

1599

Figure 1. Dependence of thermodepolarization current density on temperature for PVC (1) and PVC blends with 3% HTPB (2), 5% EPDM (3), 3% HTPB / 5% EPDM (4), 10% CPE / 3% HTPB (5), 10% CPE / 5% EPDM (6), 10% CPE / 3% HTPB / 5% EPDM (7).

perature with short-circuited electrodes for 18 20 h. Thermodepolarization was conducted in the temperature range 293403 K with an average heating rate of the polarized sample of 1.8 K/min. The basic dynamic mechanical characteristics G , G , and tan d were determined according to

the method of the free damping vibrations of the reverse torsion pendulum. The method described by Illers and Breuer 42 was adopted. Samples of 40 mm length, 7 mm width, and 1.1 mm thickness were attached rmly to the pendulum and then introduced into a thermocryogenic chamber. The measurements were conducted at the frequency of 1 Hz within the temperature interval 293383 K and at average heating rate 1.5 K/min. On the basis of the mathematical transformations described by Ferry 43 the values of G , G , and tan d were calculated by a specically designed computing program.44 The glass transition temperature ( Tg ) of PVC and its blends with HTPB, EPDM, and CPE containing 21.8% Cl was determined by using PerkinElmer DSC 2C instrument in nitrogen atmosphere at scanning rate 40 K/min.

RESULTS AND DISCUSSION


Thermally Stimulated Current Depolarization
Figure 2. Dependence of thermodepolarization current density on temperature for PVC blends with 3% EPDM (1), 5% EPDM (2) and 7% EPDM (3).

Thermally stimulated current depolarization (TSCD) and thermally stimulated creeping techniques have recently been widely utilized to study relaxation processes in PVC containing stabiliz-

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1600

STOEVA ET AL.

Table Ia. Physical Quantities of the Relaxation Dielectric Processes in PVC Blends with HTPB, EPDM, and CPE (21.8% Cl) Tmax (K) 375 imax 1 107 (A/m2) 4.25 Pmax (mC/m2) 172

No. 1

Composition 3% HTPB

D1 at 383 K
38.8

tmax (s)
153

DT (K)
330341 341367 367375 326341 341365 365374 329339 339364 364380 332346 346366 366376 330348 348366 366375 334343 343363 367378

Ea(t) (kJ/mol) 88.5 160.1 276.1 94.9 165.2 274.9 103.1 137.8 221.4 165.1 94.9 243.8 152.3 116.0 232.7 192.5 87.8 226.7

t0 (s)
1.48 1 1009 1.54 1 10020 4.62 1 10037 1.45 1 10010 2.44 1 10021 4.76 1 10037 8.94 1 10012 3.82 1 10017 3.74 1 10029 8.17 1 10022 3.32 1 10011 1.68 1 10032 9.12 1 10020 2.67 1 10014 5.64 1 10031 5.37 1 10026 5.26 1 10010 2.54 1 10030

7 27 8 27 4 27 9 21 6 2 19 13 19

5% EPDM

372

6.17

231

55.7

176

3% HTPB / 5% EPDM 10% CPE / 3% HTPB 10% CPE / 5% EPDM 10% CPE / 3% HTPB / 5% EPDM

377

6.75

260

58.8

153

375

4.00

188

42.4

155

374

6.37

293

66.1

175

376

3.03

127

28.6

162

ers, plasticizers, lubricants, 45 48 etc. The application of these techniques for the study of relaxations in PVC-CPE blends, however, is limited.8,49 The advantage of TSCD is the analysis of the distributions of relaxation parameters such as relaxation time, activation energy, etc.45 Figure 1a,b gives the experimental dependence of thermally stimulated current density imax on temperature for all blends studied. The sharp maxima in the temperature range 371377 K is considered to correspond to a dipole-segmental depolarization associated with devitrication in PVC. Their shift to higher temperature probably results from specic interactions between the components. Simultaneously, an increase in imax was observed, as compared to imax of PVC containing stabilizers only 48 which amounted to 1.13 1 10 07 A/m 2 . The increase in imax is provided by the contributions of all components: (1) By the depolarization of dipole groups or their associates in CPE; (2) by depolarization of polar hydroxyl groups in HTPB with m 5.27 1 10 030 C/m, as well as by the contribution of the additional current term of HTPB and EPDM; (3) by enhanced mobility of the polarized segments of the PVC, preferably, in the presence of some additives. The larger increase in imax was observed with the addition of EPDM alone or, also, its combination with HTPB and CPE (Fig. 1a,b, 2, and Table Ia). Figure 2

shows that the increased concentration of EPDM in the blends of 37 mass % resulted in the corresponding increase of imax from 2.5 1 10 07 to 9.8 1 10 07 A/m 2 . At the same time, the value of Pmax changed within 92.7452.3 mC/m 2 . It is very likely that EPDM as a nonpolar elastomer, situated among structural formations of PVC, would facilitate both the group and segment mobility and, hence, the preliminary orientation of dipoles. This suggestion is conrmed by the high values of Pmax for samples 2, 3, and 5 (Table Ia), as compared to Pmax value of PVC containing stabilizers only 48 which is 62 mC/m 2 . Moreover, the Pmax value of the PVC / 10% CPE (21.8% Cl) 49 was 223 mC/m 2 , whereas the Pmax value of the polyblend PVC / 10% CPE (21.8% Cl) / 5% EPDM was 293 mC/m 2 . The lower values of imax and Pmax for blends containing HTPB resulted from the possible dipoledipole interaction between their polar groups which is likely to retard the orientation during polarization. The lowest imax value was observed with the polyblend containing all components studied (Fig. 1, curve 7). The pattern of change in the increment of relative dielectric permeability D1 is similar to that of Pmax , i.e., as EPDM is introduced into one- and two-component blends, D1 values are higher (Table Ia). To calculate the relaxation times tmax eq. (2) was used. The tmax values of all studied blends

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

COMBINED RELAXATION STUDY OF PVC BLENDS

1601

Table Ib. Kinetic Parameters of the Relaxation Dielectric Processes in PVC Blends with HTPB, EPDM, and CPE (21.8% Cl)

No. 1

Composition 3% HTPB

DH* (kJ/ mol)


85.7 157.0 273.0 92.1 162.2 271.8 100.3 134.9 218.3 162.3 92.0 240.7 149.4 113.1 229.6 189.7 84.8 226.6

DS* E (kJ/mol 1 K)
065.3 143.7 459.3 046.5 159.1 459.2 023.5 78.8 308.0 168.4 034.9 372.0 129.3 024.3 342.9 248.4 057.8 330.3

DG* (kJ/ mol)


107.7 106.3 102.6 107.6 106.0 102.0 108.1 107.2 102.8 105.2 104.3 102.8 105.6 107.0 102.6 105.7 105.2 103.5

Ea(imax) (kJ/ mol) 241

Ea(Th) (kJ/ mol) 273

Ea(Gw) (kJ/ mol) 283

XFF

Ea(v) I (kJ/ mol) 285

Ea(t) I (kJ/ mol) 280

H Ea(d) (kJ/ mol) 285

0.40

5% EPDM

206

353

364

0.48

317

361

262

3% HTPB / 5% EPDM 10% CPE / 3% HTPB 10% CPE / 5% EPDM 10% CPE / 3% HTPB / 5% EPDM

242

313

324

0.44

311

321

281

246

374

361

0.45

329

358

285

216

286

296

0.38

306

293

245

237

319

329

0.42

317

326

295

were lower than the tmax of PVC containing stabilizers only which was 250 s with b 1.8 K/min.48 The lower tmax values of the blends proved the increased mobility of the structural units. The Arrhenius ln t f (10 3 /T ) plots are presented in Figure 3. The linear sections of the plots were used to calculate the Ea values according to eq. (3). The calculated values are listed in Table Ia. An increase in Ea ( t ) with the temperature rise was observed for blends containing HTPB and/or EPDM. For blends containing CPE, the lowest Ea values were found in the middle of the temperature range (343366 K). Similar changes were observed in the values of the pre-exponential factor to estimated according to eq. (3). The increase of Ea ( t ) and the corresponding decrease of to are known to prove the enhancement of the degree of cooperation for dipole-segmental relaxation.21 Accordingly, the number of kinetic units participating in the corresponding relaxation process was determined, based on the values of to 10 013 s which is characteristic for dipole-segmental processes. For blends containing HTPB and/or EPDM, the cooperative character of dipole-segmental process appeared at temperatures exceeding 339 K (Table Ia). The cooperative character was more strongly pronounced with PVC-

based blends containing one elastomeric additive (N 27 at T 365 K). Somewhat ambiguous change in N was observed with blends containing CPE (Table Ia). Here the cooperative character of dipole-segmental relaxation was clearly pronounced in the rst temperature range and, even better, in the third interval. The function ln to f (Ea ) for blends 16 (Table Ia) is presented in Figure 4. As seen from the gure, the compensation law here applied to the data obtained since a linear relationship was observed. On the basis of the compensation law in accordance with eq. (6), the parameters Tc and tc were determined. Values are: Tc 349.5 K and tc 21498 s. This fact suggests a molecular origin similar of the dielectric relaxation for all blends. To estimate the changes in the enthalpy, entropy, and Gibbs free energy, eqs. (7) (9) were used. The results are presented in Table Ib. The DH* values were close to the Ea ( t ) values and exhibited the same trend in their change. A similar correlation was observed between the DS* and E to values. The transition from polarized into depolarized state gave some values of DS 0. The negative values of entropy reect the inuence of electrical eld on the formation of a local sequence of the cinetic units.50 Such a sequence can exist

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1602

STOEVA ET AL.

Dipole Moments and Current Carrier Density of Characteristic Atomic Complexes in PVC Blends with HTPB, EPDM, and CPE (21.8% Cl)

k 1 1005 a.c./c.c.

5.66 3.16 2.80

4.75

I n* 1 10022 (m03)

3.52 6.06 6.57

4.02

2.26 7.82 1.83 1.77 10.37 22.92 0.003

I n 1 10015 (m02)

1.07 1.54 1.63

1.17

n 1 10028 (m03)

1.99 1.91 1.84

1.84

mtheor 1 1030 (C 1 m)

9.84 9.98 10.14

10.22

meff 1 1030 (C 1 m)

16.55 20.23 21.20

18.01

0.003 0.003

0.003

CH

15.36

10.53

1.70

0.79

2.23

7.64

Atomic group : single CHCl group ratio

CH2OH

Figure 3. Arrhenius plots of the time of relaxation for PVC blends with 3% HTPB (), 5% EPDM (l), 3% HTPB / 5% EPDM (), 10% CPE / 3% HTPB (), 10% CPE / 5% EPDM (), 10% CPE / 3% HTPB / 5% EPDM ( ).

0.001 0.001

0.001

CH|CH

0.001

at temperatures up to 340 K in the absence of CPE. The introduction of CPE results in the formation of the local sequence at temperatures within 343366 K. The DG* values were approximately the same for all blends studied. These values varied within 102.0108.1 kJ/mol. From the Table Ib data, the following conclusions can be made: (1) The activation energy Ea (w) determined by Chens method has values within 285 329 kJ/mol, whereas the corresponding values of Ea (imax ) are lower (206242 kJ/mol). The lowest values of Ea ( tmax ) were found for blends containing EPDM. We believe this proves the fact that the additive facilitates the dipole orientation polarization of PVC; (2) Ea values for the lowtemperature section of the dependence i f (T ) determined by different methods are very close; evidence of their universality; (3) Ea values for the high-temperature section of the dependence i

0.03 0.03

0.03

CH(CH3)

0.03 0.03

0.03 1.42 5

1.07 1.16 1.22

1.32

Table Ic.

No.

1 2 3

/ 8q5d$$4014

04-17-98 12:51:17

3% HTPB 5% EPDM 3% HTPB / 5% EPDM 10% CPE / 3% HTPB 10% CPE / 5% EPDM 10% CPE / 3% HTPB / 5% EPDM

Composition

1.49

CH2

0.03

0.03

polpa

W: Poly Physics 9504014

COMBINED RELAXATION STUDY OF PVC BLENDS

1603

Dynamic Mechanical Analysis The dependences of G , G , and tan d on temperature for the blends studied are shown in Figures 59. The G versus T plot (Fig. 5) showed an abrupt change of course in the temperature range 326348 K due to the PVC transition from glassy to a highly elastic state (basic a-relaxation). It is worth considering some peculiarities of the curves depending on the type and combination of polymers. The following phenomena were observed: (1) The slope of the G plot in the transition region increased for blends containing HTPB, EPDM, and CPE (Fig. 5b); (2) the values of G decreased considerably in the range 293340 K for blends with CPE added (Fig. 5b, curves 58). The drop in G value in this case proved the existence of weaker dipole-dipole interactions between the macromolecules of PVC, which were strongly affected by the sequences of nonchlorinated methylene groups in CPE. This was conrmed by the very low values of DCp for blends containing CPE (Table II). The (PVC / 3% HTPB) blend was the only one whose G value was higher (above 323

Figure 4. Plot of ln to vs. Ea ( t ) for PVC blends with 3% HTPB ( ), 5% EPDM ( l), 3% HTPB / 5% EPDM (), 10% CPE / 3% HTPB (), 10% CPE / 5% EPDM (), 10% CPE / 3% HTPB / 5% EPDM ( ).

f (T ) are lower, as compared to the corresponding values for the low-temperature section, since the depolarization processes are facilitated with the temperature increase; (4) the relaxation processes in the blends studied show approximately monomolecular kinetics, since the XFF values are within 0.380.48. Based on the chemical composition of the components in the blends studied and their content, the type of the atomic groups and their fraction with respect to a single chloromethylene (CHCl) group was evaluated. To calculate the effective dipole moment meff for each combination of atomic groups, the Langeven-Debye equation was used (Table Ic). The theoretical dipole moment mtheor can also be calculated from the literature data.51,52 Data in Table Ic revealed that meff mtheor for all the blends. Hence, the suggestion adopted previously that Pmax was determined by the dipole constituent in the atomic groups alone proved to be incorrect. Obviously, apart from the dipole constituent, the nondipole one also contributed to Pmax . Provided Pmax comprises a nondipole constituent alone, i.e., free charge carriers, one can calculate both the surface (n ) and bulk (n *) densities of the current carriers, as well as the coefcient k which accounts for the number of atomic complexes (a.c.) per current carrier (c.c.). Higher values of n and n * and lower values of k were found for blends containing EPDM. The blend with a composition PVC / 10% CPE / 3% HTPB / 5% EPDM represented an exception.

Figure 5. Temperature dependence of shear modulus (G ) of PVC (1) and PVC blends with 3% HTPB (2), 5% EPDM (3), 3% HTPB / 5% EPDM (4), 10% CPE (5), 10% CPE / 3% HTPB (6), 10% CPE / 5% EPDM (7), 10% CPE / 3% HTPB / 5% EPDM (8).

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1604

STOEVA ET AL.

Table II. Calorimetric Parameters of PVC Blends with HTPB, EPDM, and CPE (21.8% Cl) Composition 3% HTPB / 5% EPDM 363.8 5.0 10.4 248.6 10% CPE / 3% HTPB 361.7 2.9 10.0 164.5 10% CPE / 5% EPDM 360.4 1.6 10.4 149.9 10% CPE / 3% HTPB / 5% EPDM 362.6 3.8 11.1 147.8

Parameters Tg (K) Tgi Tgo (K) DTg (K) DCp (J/kg K)

PVC / Stabilizers 358.8 0 13.8 284.0

3% HTPB 361.7 2.9 12.9 276.2

5% EPDM 361.2 2.4 10.4 237.3

10% CPE 362.5 3.7 11.6 114.0

K) than the corresponding value of the initial PVC. An explanation of this fact is believed to be the enhanced physical interaction between PVC and HTPB polar groups. Probably, as a result of the increased number of crosslinks including hydrogen bonds, the resistance of the material toward mechanical treatment increased. Figures 6 and 7 demonstrate the dependences of G on temperature for the blends studied. A well-pronounced maximum in the region of a-relaxation of PVC was observed. For PVC containing stabilizers only, the same maximum was observed at 337 K (Fig. 6, curve 1). The maximum for the remaining blends appeared above 337 K

with the exception of the blend PVC / 10% CPE (335 K). The increase of Tmax (G ) probably was due to steric hindrances involving the solvated aggregates of PVC segments while changing their position. In addition, weak maxima were observed in the range 310320 K. According to Perepechko 53 they can be attributed to the enhanced mobility of PVC macromolecules associated with their irregular regions. The enhanced mobility probably is associated with the decomposition of dipolar bonds acting as local physical nodes in the macromolecular network. According to Bartenev and Sinitsina 54 this represents the so-called pCl

Figure 6. Temperature dependence of loss modulus (G ) for (1) PVC and (2) PVC blends with 3% HTPB, (3) 5% EPDM, (4) 3% HTPB / 5% EPDM.

Figure 7. Temperature dependence of loss modulus (G ) for PVC blends with (1) 10% CPE (21.8% Cl), (2) 10% CPE / 3% HTPB, (3) 10% CPE / 5% EPDM, (4) 10% CPE / 3% HTPB / 5% EPDM.

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

COMBINED RELAXATION STUDY OF PVC BLENDS

1605

Figure 8. Plot of tan d vs. temperature for (1) PVC and (2) PVC blends with 3% HTPB, (3) 5% EPDM, (4) 3% HTPB / 5% EPDM.

process of relaxation. In the case of the blends containing more than one polymer, the dependence G f (T ) was multiplet (Fig. 7). For (PVC / 10% CPE) blend two maxima were recorded the rst one was in the range 306318 K and was associated with the complex a-relaxation in CPE 55 57 as well as the decomposition of pCl -dipolar bonds. The other maximum was found at 335 K and can be assigned to the transition of PVC from the glassy into the highly elastic state. The remaining three- and four-component blends exhibited three regions of relaxation. The rst was in the range 301307 K and again was associated with the a-relaxation of CPE. The third region was referred to the main relaxation process in PVC. The well pronounced second (intermediate) region in the range 323327 K corresponded to the increased segment mobility of PVC macromolecules occupying the interglobular area and participating in the formation of a boundary interfacial layer. This assumption is based on the structuralmorphological model of impact-resistant PVC reported by Monakhova et al.58 According to the proposed model, in the case of impact-resistant PVC blends, interfacial layer is developed including modier molecules, globular fragments, and intermediate PVC macrochanges located in the interglobular area. The blends studied were thermodynamically incompatible. The blend (PVC / 10% CPE) (21.8% Cl) studied recently 49 has

shown similar behavior. Reductions in the temperature interval of glass transition ( DTg ) by 1 4 K for PVC blends support this suggestion (Table II). Apart from this, the formation of an interfacial layer with specic relaxation region proved the formation of joint structures. These structures are characterized by a variety of physical nodes caused by dipolar interactions, local Van der Waals forces involving C|C double bonds in adjacent macromolecules of HTPB, etc. As a result, a shift of Tg by 1.55 K toward higher temperatures was observed for PVC in the blends (Table II). Figures 8 and 9 illustrate the dependence of tan d on temperature. Obviously, microregions of different composition and size were formed in the polymer blends studied. These microregions were considered to be responsible for the appearance of two (Fig. 8) or three and more (Fig. 9) regions of dissipation. We know 59 that the appearance of additional maxima of (tan d vs. T ) function proves the existence of segment solubility of the components on the boundary phase surface. This effect was well pronounced with blends containing CPE and elastomeric additives (Fig. 9). In this case a wide temperature range (345380 K) was observed where the vitrication of PVC occurred

Figure 9. Plot of tan d vs. temperature for PVC blends with (1) 10% CPE (21.8% Cl), (2) 10% CPE / 3% HTPB, (3) 10% CPE / 5% EPDM, (4) 10% CPE / 3% HTPB / 5% EPDM.

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1606

STOEVA ET AL.

Table III. Activation Energy of Characteristic Relaxation Processes According to Dynamic Mechanical Spectrum (tan d) in PVC Blends with HTPB, EPDM, and CPE (21.8% Cl) Tmax1 (K) 349 352 348 347 347 346 346 345 Ea1 (kJ/mol) 76.9 77.5 76.7 76.4 76.4 76.2 76.2 76.0 Tmax2 (K) 362 363 357 353 Ea2 (kJ/mol) 79.7 80.0 78.6 77.8 Tmax3 (K) 374 376 375 377 373 378 378 372 379 Ea3 (kJ/mol) 82.4 82.8 82.6 83.1 82.2 83.3 83.3 81.9 83.5

No. 1 2 3 4 5 6 7 8

Composition PVC / stabilizers 3% HTPB 5% EPDM 3% HTPB / 5% EPDM 10% CPE 10% CPE / 3% HTPB 10% CPE / 5% EPDM 10% CPE / 3% HTPB / 5% EPDM

(Table III). The poly(ethylene) chlorinated in a uidized state can be regarded as a block copolymer consisting of sequences of methylene (CH2 ) and chloromethylene (CHCl) groups. Such a structure facilitates the adhesion between polar (PVC) and nonpolar (EPDM) or weakly polar (HTPB) ingredients and contributes to the formation of interfacial layers. The observed variety of relaxation processes in the blends conrms the validity of the above considerations. The Ea values at the maxima of the tan d versus temperature plots were calculated using eq. (14). Data in Table III showed a negligible increase in Ea as the temperature rose. Ea values varied in the range 76.0 83.5 kJ/mol.

CONCLUSIONS
A number of blends consisting of suspension poly(vinyl chloride) (PVC) with stabilizers, poly(ethylene) chlorinated in uidized state (CPE), hydroxyl-terminated poly(butadiene) (HTPB), and ethylene-propylene-diene terpolymer (EPDM) have been studied. Toward this purpose, the methods of thermally stimulated depolarization and dynamic mechanical analysis have been employed. As seen from the tables the temperatures Tmax3 (DMA) were found to have values higher (by 24 K) than Tmax (TSCD), i.e., some minor shift between the mechanical and dielectrical values of Tmax for the blends existed, concerning the

a-relaxation process. Glass transition temperatures (Tg ) determined by DSC had on average lower (by 1318 K) values than Tmax . Taking into account the difference between the frequencies of DMA and TSCD, a larger difference between the corresponding temperature positions should be expected. This difference appears to be compensated by the various mode of excitation of the molecular segments. Usually, mechanical stress is known to inuence the whole section of the segment, thus exciting the latter more easily than the electrical forces which, in turn, affect the polar part of the segment only.60 This assumption was conrmed by the lower values of Ea determined by DMA, as compared to Ea (TSCD). On the other hand, the lower values of Ea (DMA) probably are determined by the fact that chain segments of a smaller number of monomer units are involved in the mechanical a-relaxation process, whereas the corresponding number of monomer units taking part in the dielectrical relaxation is larger. The comparison of the TSCD and DMA (tan d ) spectra resulted in the detection of one a-relaxation peak only for the corresponding dielectrical spectra. Therefore, the blends studied should be considered as homogeneous. DMA data, however, pointed to the existence of heterogeneous blends, since a number of peaks in the a-relaxation region was observed. Consequently, the existence of only one region of glass transition in accordance with the TSCD and DSC data, does not necessarily mean homogeneity of the system.61 This ambiguity is also supported by the varying sensitivity of

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

COMBINED RELAXATION STUDY OF PVC BLENDS

1607

the different methods employed with respect to the dimensions of the domains subjected to the relaxation process. Since the dynamic mechanical method was more sensitive with respect to various kinds of molecular motions taking place in domains of small dimensions, 61 the dependence tan d f (T ) reected the multiplicity of relaxation transitions. The additives studied containing both the polar and nonpolar functional groups were found to affect PVC by various patterns. On one hand, these additives caused a decrease of the dipolar interactions in PVC at the expense of the nonpolar sequences of methylene groups in CPE and EPDM. This phenomenon was proved by the diminished relaxation times ( tmax ) of the blends, as compared to PVC alone and, also, the lower values of DCp and the higher values of Pmax . On the other hand, the so called physical nodes were formed as a result of the corresponding physical interactions between the polar and polarizing groups in PVC v v and the additives, such as CHCl, CHOH, u u {CH|CH{, etc. These interactions led to a weak increase of Tmax (TSCD), Tmax (G ), Tg (DSC), as well as an increase of height of the G peak. As a matter of fact, the effect of the additives on PVC was associated with two mutually contradicting phenomena. The rst represented the plasticizing effect which resulted in the increase of segmental mobility of PVC. The second effect was solvating associated with the formation of physical nodes in the macromolecular network of the polymer. This effect was more pronounced with the such polar additives as CPE and HTPB. The whole structure of the system became more microheterogeneous by the formation of aggregates which possessed both the various size and inner microstructure. Therefore, the kinetics of the mechanical and dielectrical a-relaxation of PVC blends is similar and is determined by the conformation changes in the chain segments of different length. The suprasegmental structure of PVC in all of the blends can be considered as one of the same kind since the systems studied conformed to the compensation law. The quantitative characteristics only are subject to change, depending on the kind of additives and the physical interactions between the latter and PVC.
The authors acknowledge the nancial support of the

Bulgarian Ministry of Science and Education, through the National Research Fund.

REFERENCES AND NOTES


1. A. Takahashi, Polym. Eng. Sci., 22, 48 (1982). 2. A. A. Dontsov, G. Ya. Lozovik, and S. P. Novitskaya, in Chlorination of Polymers, Khimiya, Moscow, 1979, p. 108 (in Russian). 3. C. F. Ryan and R. L. Jalbert, in Encyclopedia of PVC, L. I. Nass, ed., Marcel Dekker, New York, 1977. 4. A. Siegmann and A. Hiltner, Polym. Eng. Sci., 24, 869 (1984). 5. C. E. Locke and D. R. Paul, Polym. Eng. Sci., 13, 308 (1973). 6. J. Zelinger, J. Polym. Sci., C(16), 4259 (1969). 7. Y. J. Shur and B. Ranby, J. Appl. Polym. Sci., 20, 3105 (1976). 8. J. J. del Val, C. Lacabanne, and A. Hiltner, J. Appl. Phys., 63, 5312 (1988). 9. R. R. Blanchard and C. N. Burnell, SPE J., 24, 74 (1968). 10. A. Varkalis, J. Zican, M. Kalnin, and M. Ioelovich, Zinat. Akad. Vestis. Kim. Ser., 5, 571 (1987). 11. C. P. Doube and D. J. Walsh, Eur. Polym. J., 17, 63 (1981). 12. G. F. Ignatova, V. L. Trizno, A. F. Nikolaev, and N. V. Daniel, Plast. Massy, 7, 39 (1974). 13. D. Ferenc, M. Akutin, L. Kiss, and N. Tichonov, Plaste und Kautsch., 31, 136 (1984). 14. Xu Xi, Zh. Liwu, and Li Huilin, Polym. Eng. Sci., 27, 398 (1987). 15. Yu-Der Lee and Chi-Ming Chen, J. Appl. Polym. Sci., 33, 1231 (1987). 16. H. Frey, H. Klug, and A. Mukerjee, Ger. Patent, No. 2319044 (1974). 17. S. Bonotto and E. Wagner, US Patent, No. 3396211 (1968). 18. S. Stoeva, unpublished data. 19. S. Stoeva, M. Michailov, S. Malamusis, and K. Jankova, Bulg. Patent, No. 44287 (1988). 20. Iv. Mladenov, K. Jankova, Tsv. Karadzhova, Bulg. Patent, No. 41562 (1985). 21. G. A. Lushcheikin, in Electric Properties of Polymers. Investigation Methods, Khimiya, Moscow, 1988, p. 146 (in Russian). 22. Yu. V. Zelenev and L. A. Deltuva, Vysokomol. Soedin., A24, 2383 (1982). 23. R. Barrer, J. Phys. Chem., 61, 178 (1957). 24. J. D. Hoffman, G. Williams, and E. Passaglia, J. Polym. Sci., C14, 176 (1966). 25. C. Lacabanne, D. Chatain, J. C. Monpagens, A. Hiltner, and E. Baer, Solid State Commun., 27, 1055 (1978). 26. J. Hoffman and H. Pfeiffer, J. Chem. Phys., 22, 132 (1954).

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

1608

STOEVA ET AL.

27. S. Glasstone, K. Laidler, and H. Eyring, in The Theory of Rate Processes, McGraw-Hill, New York, 1941 (Russian transl.). 28. P. Kartalov, Ph.D. thesis, Plovdiv University, Plovdiv, 1974, pp. 6367 (in Bulgarian). 29. R. Chen, J. Appl. Phys., 40, 570 (1969). 30. J. van Turnhout, Polym. J., 2, 173 (1971). 31. L. J. Grossweiner, J. Appl. Phys., 24, 1306 (1953). 32. P. Kartalov, in Physical Methods for Investigation of Polymers, Plovdiv University, Plovdiv, 1987, pp. 4154, 274302 (in Bulgarian). 33. C. Bucci, R. Fieschi, and G. Guidi, Phys. Rev., 148, 816 (1966). 34. D. W. van Krevelen, in Properties of Polymers, Correlation with Chemical Structure, Elsevier, Amsterdam, 1972 (Russian transl.). 35. B. P. Shtarkman, in Plastication of Poly(vinyl chloride), Khimiya, Moscow, 1975, p. 234 (in Russian). 36. L. E. Nielsen, in Mechanical Properties of Polymers and Composites, Marcel Dekker, New York, 1974 (Russian transl.). 37. G. M. Bartenev, G. M. Sinitsyna, and I. L. Gorelova, Kompozitsionnye Materialy, 21, 8 (1984). 38. P. Kartalov, D. Donchev, and V. Lyudskanov, Nauchni trudove VPI-Plovdiv, Fizika, 6, 67 (1968). 39. A. K. Doolittle, J. Appl. Phys., 22, 1471 (1951); ibid, 23, 236 (1952). 40. P. Kobenko, E. Kuvshinskii, and N. Shishkin, Zh. Teoret. Fiz., 14, 10 (1944). 41. P. Kobenko, in Amorphous Materials, AN SSSR, Moscow, 1952, p. 278 (in Russian). 42. K. H. Illers and H. Breuer, Kolloid-Z., 176, 110 (1961). 43. J. D. Ferry, in Viscoelastic Properties of Polymers, John Wiley and Sons, New York, 1970 (Russian transl.). 44. Kh. Ponevski and M. Mateev, Fizika, 21, 23 (1983).

45. A. L. Kovarskii, S. A. Mansimov, and A. L. Buchachenko, Polymer, 27, 1014 (1986). 46. M. Kalman, Polym. Bull., 22, 213 (1989). 47. T. S. Gancheva and P. D. Dinev, Eur. Polym. J., 18, 159 (1982). 48. P. Kartalov, S. Stoeva, Kh. Polizov, and M. Mateev, Fizika, 27, 95 (1989). 49. P. Kartalov, S. Stoeva, S. Kartalov, and M. Mateev, Fizika, 28, 142 (1990). 50. M. D. Migahed, M. Ishra, and T. Fahmy, J. Phys. D., 27, 2216 (1994). 51. V. A. Afanasiev, in Physical Methods in Studying the Structure of Organic Compounds, ILIM, Frunze, 1968, p. 24 (in Russian). 52. B. M. Tareev, in Physics of Dielectric Materials, Energoizdat, Moscow, 1982, p. 119 (in Russian). 53. I. I. Perepechko, in Akustic Methods in Polymers, Khimiya, Moscow, 1973, p. 109 (in Russian). 54. G. M. Bartenev and G. M. Sinitsina, Vysokomol. Soedin., A38, 799 (1996). 55. R. F. Boyer, in Transitions and Relaxations in Polymers, John Wiley and Sons, New York, 1966 (Russian transl.). 56. G. M. Bartenev, R. M. Aliguliev, and D. M. Khiteeva, Vysokomol. Soedin., A23, 2003 (1981). 57. R. Popli, M. Glotin, and L. Mandelkern, J. Polym. Sci. Polym. Phys. Ed., 22, 407 (1984). 58. T. G. Monakhova, L. I. Batueva, T. B. Zavarova, and A. P. Savelev, Vysokomol. Soedin., A33, 1634 (1991). 59. Encyclopaedia of Polymers, Vol. 3, Soviet Encyclopaedia, Moscow, 1977, p. 437 (in Russian). 60. G. M. Bartenev and A. G. Barteneva, in Relaxation Properties of Polymers, Khimiya, Moscow, 1992, p. 245 (in Russian). 61. Polymer Blends, Vol. 1, D. R. Paul and S. Newman, eds., Academic Press, New York, 1981 (Russian transl.).

/ 8q5d$$4014

04-17-98 12:51:17

polpa

W: Poly Physics 9504014

Вам также может понравиться