You are on page 1of 9

FUEL PROCESSING TECHNOLOGY

ELSEVIER
Fuel ProcessingTechnology49 ( 1996)65-73

Thermal and catalytic degradation of commingled plastics


Manjula M. Ibrahim, Eric Hopkins, Mohindar S. Seehra
Physics Department, West Virginia Uniuersity, Morgantown, WV 26506-6315, USA

Received 2 1 September 1995;accepted28 February 1996

Abstract Tbermogravimetry, in situ electron spin resonance (ESR) spectroscopy and in situ and ex situ X-ray diffraction (XRD) are employed to investigate the thermal and catalytic degradation of a sample of commingled plastics (CP). XRD studies show CP to contain about 90% polyethylene (PE) and 10% polypropylene (PPE) and a smaller amount of TiO,. Analysis of the weight loss data in argon to 550C yields an activation energy E = 38 kcal mol- I for the thermal decomposition of CP. In XRD studies, a melting point of 135C is inferred whereas the onset of irreversible degradation begins only around 360C. The in situ ESR experiments of CP and CP loaded with 10% Al,O, (nanoscale) and 10% sulfur, both under 500 psig of H,, show that for CP alone an ESR signal indicative of degradation is first seen near 380C whereas for the loaded CP this degradation temperature is reduced to 280C. This enhanced catalytic degradation detected by ESR is believed to be due to elemental sulfur. Keywords: Activation energy; Electron spin resonance; Plastics degradation; X-ray diffraction

1. Introduction The properties and thermal degradation of model polymers such as polymethylene (PM), polyethylene (PE) and polypropylene (PPE) are reasonably well known [1,2]. The

degradation/depolymerization process is considered to be a free radical chain reaction involving initiation, propagation, transfer and termination with activation energy E = 45 to 60 kcal mol- and significant degradation occurring only above 380C [ 1,2]. In this

* Corresponding author. 0378-3820/%/$15.00 Copyright 6 1996Elsevier Science B.V. All rights reserved. PII SO378-3820(96)01023-5

66

M.M. Ibrahim et al./Fuel

Processing Technology 49 (1996) 65-73

paper, we report the results of our recent investigations of the composition and thermal and catalytic degradation of a sample of commingled plastics (CP) employing the techniques of X-ray diffraction (XRD), thermogravimetry and in situ hightemperature/high-pressure electron spin resonance (ESR) spectroscopy. This project is a part of a larger program of coprocessing waste plastics and coals into useful liquefaction products.

2. Experimental procedures The sample of CP used in these investigations was supplied by Dr. Larry Anderson of the University of Utah who obtained it from the American Plastics Council. The elemental analysis of the CP provided by Dr. Anderson is as follows: C, 83.3%; H, 13.4%; N, 0.08%; S, 0.024%; ash, 0.98%; oxygen (by difference), 2.21%. The XRD investigations were carried qut with a Rigaku/DMax diffractometer under both in situ and ex situ heating conditions. Thermogravimetric investigations were done with a Mettler 3000 system. For in situ ESR investigations at 9 GHz, we employed a gold-plated TE ,a2 microwave cavity, internal modulation at 100 KHz, sapphire sample tubes for high-pressure capabilities and an indirect sample heating system details of which have been given elsewhere [3,4].

3. Results and discussion The room temperature X-ray diffractogram of the CP sample is shown in Fig. 1, where we have identified the peaks due to PE and PPE. A crude estimate of the relative concentrations of PE and PPE in the sample can be made from the relative areas under the peaks, this estimate being 90% PE and 10% PPE in the CP sample. A more accurate determination should be possible by developing calibration curves using calibrated mixtures of PE and PPE. This work is now in progress. However, for the sake of these investigations, it is clear that the CP sample contains primarily high-density PE. The effects of heating the sample in air ex situ at different temperatures for 30 min, followed by furnace cooling to room temperature and X-ray diffractometry, are also shown in Fig. 1. For heating at 100, 200 and 300C followed by cooling to room temperature, PE in the sample appears to recrystallize since the prominent PE peaks are still visible. For heating at 360C the intensity of the PE peaks diminishes considerably and peaks due to TiO, become distinct. For heating at 42OC, the PE peaks disappear completely suggesting irreversible degradation of PE. In Fig. 2, we show the results of a second experiment in which the CP sample was heated in situ during XRD measurements in nitrogen atmosphere using a special high-temperature furnace. The weaker PPE peaks are no longer visible partly because of the attenuation caused by the X-ray window of the furnace, and the broad peak near 28 = 10 is due to the Duco cement used for gluing the sample. The significant result here is that the PE peaks disappear between 130 and 135C which coincides with the

M.M. Ibrahim et al./ Fuel Processing Technology 49 (1996) 65-73

6-l

??

Ti02 lines

unheated

100-c

360C

i420C

10

15

20

25

30

35

2-theta

(degrees)

Fig. 1. Room temperature X-ray diffmctograms of the commingled plastics heated ex situ in air to the temperatures indicated. PE and PPE respectively represent polyethylene and polypropylene and the numbers in parentheses show the Miller indices of the lines.

melting

temperature of PE [l]. Above this temperature, PE loses its long-range crystalline order resulting in disappearance of the Bragg peaks. The broad peak near 28 5: 18 at higher temperatures is due to amorphous PE [5]. Combining this result with the results of Fig. 1, we conclude that irreversible degradation of CP begins only around 360C whereas 135C is its melting temperature. Our conclusion that thermal degradation of CP begins near 360C is quite consistent with earlier experiments on pure PE [2]. Using the data of Fig. 2 and the procedure outlined by Aggarwal and Tilley [5] involving deconvolution and measuring the areas under the peaks after correcting the intensity for atomic scattering and diffraction angle, we show in Fig. 3 the relative intensity of the amorphous and crystalline components of CP as a function of temperature. At room temperature, the crystallinity of CP is about 50%. The crystalline component disappears near 135C, resulting in a net increase of the amorphous component at this transition. With further increase in temperature, there is a systematic loss of the amorphous component, so that by 450C even the amorphous component disappears, most likely due to loss of material. Further confirmation of the above conclusions comes from our thermogravimetric experiments on CP carried out at heating rates of 5, 10 and 20C min- in argon flow

M.M. Ibrahim et al./Fuel

Processing Technology 49 (1996) 65-73

PE (ua 1

PE (200)

24'C

25O'C

10

15

20

25

30

35

a-theta

(degrees)

Fig. 2. In situ X-my diffractograms of the heated commingled plastics to the temperatures shown. The broad component near 20 = 18is due to the amorphous component of polyethylene.

(Fig. 4). The corresponding rate of mass loss curves against temperature are shown in Fig. 5. Since the mass loss is indicative of the scission of the chemical bonds leading to volatilization, it is clear from Fig. 4 and Fig. 5 that even at the lowest heating rate, this process becomes noticeable only above 300C. This is consistent with the results from the XRD studies described above. These results can be made more quantitative by the calculation of the associated activation energy E using the temperature T, = 480, 505 and 5 15C where the rate of mass loss ( - dm/dt) is maximum (Fig. 5) for the heating ,) against l/T, should rate p = 5, 10 and 20C min- respectively. A plot of ln(P/T yield a straight line with the slope = -E/R [6], where R is the gas constant. This plot, shown in Fig. 6, yields E = 37 kcal mol - for the thermal decomposition of the CP

M.M. Ibrahim et al. / Fuel Processing Technology 49 (1996) 65-73


140001 I

69

Comingled Plastics

t
12000 : 0 '>b', -aQ Q 0 0 n Amorphous Total Crystalline

o'-10000

I1 \

100 1

!!

90

8000 -

6000 50 40 4000 30 2000 20


10

8 b a s 'g 5 oc

0
0

1
50

I,

I,

~a*_r?_~______n______n--____a______*-______n

I, I, I,
200

I1

..I
I,,

I,,

I,,

I,,

100

150

250

300

350

400

450

500

Temoerature

( Cl

Fig. 3. Temperature dependence of the crystalline, amorphous and total (amorphous plus crystalline) components of CP as determined from diffractograms of Fig. 2.

o ,, 300

, , , , , , , 350

,x>
400 450 500 5 i0

T ( C)
Fig. 4. Remaining weight versus temperature for the commingled plastics in argon flow for the three heating rates of 5, 10 and 20C min- . The horizontal lines represent constant decomposition fraction cx= 0.5 and Q = 0.7.

70

M.M. Ibrahim et al. / Fuel Processing Technology 49 (I 996) 65-73

T ( C)
Fig. 5. The rate of weight change (dm/dr) versus temperature data for the commingled plastics in argon for three heating rates. T,, is the temperature of maximum weight loss = -dm/dt.

sample. The data in Fig. 4 can also be used to determine E by the constant (Y(fraction of decomposition) method of Flynn and Wall [7] where 1 din/3 E =--(1) R b d(l/T) where b is a constant = 1. By retaining higher-order terms in the calculations, terms

-14.0
E _ 37 kcal/mol
c

-16.0 L 1.25

1.30 l/Tm(10-3/K)

1.35

Fig. 6. Plot of In@/Ti) versus l/T, for the heating rates p = 5, 10 and 20C min- . The solid line is the least squares fit, the slope yielding E = 37 kcal mol- .

M.M. Ibrahim et al. / Fuel Processing Technology 49 (I 996) 65-73

71

-1.0

E-

42 kcal/mol

Q 2 s F -2.0 -

-2.5 -

J 1.25 1.30 1/T(10-3/ K)


Fig. 7. Plot of In p versus l/T for a = 0.5 and 0.7, the solid lines being the least-squares tit through the points. See text for activation energies.

1.35

1.40

which

are weakly

dependent on temperature, we have derived the following expression

for b
-b=l+(~,R)

2T

T2

(2)

The plots of In B versus l/T for (Y= 0.5 and (Y= 0.7 are shown in Fig. 7. For E = 37 kcal mol - and temperatures used in Fig. 7, we find b = 1.08. Using this value of b, we = 0.5 yielding an find E = 42 kcal mol - for (Y= 0.7 and E = 36 kcal mol- for OL average value of 39 kcal mol- determined above by a different method, although using the same data. For comparison, for pure PE, values of E from 45 to 70 kcal mol- have been reported by different investigators [1,2]. We also measured thermogravimetric weight loss data in air. This yielded uneven curves most likely due to interaction with oxygen. Consequently, we were not able to determine the relevant activation energies using the above procedures. We next consider the results of the ESR experiments. First, it is noted that the CP sample does not give a detectable ESR signal at room temperature. In Fig. 8, we show the results of careful ESR experiments carried out under 500 psig of H, pressure for two cases: (i) CP alone and (ii> CP plus 10% nanoscale Al,O, plus 10% elemental sulfur. The nanoscale Al,O, was obtained from Nanophase Technologies Corporation (453 Commerce St. Burr Ridge, IL 60521, USA) and it has a particle size of 20 nm with 45 m2 g-r surface area. Elemental sulfur was added to simulate the recent successful ESR experiments with Fe-based catalysts [8], and because with coal liquefaction, sulfur

72
Commingled

MM. Ibrahim et al./Fuel Plastics (CP) 244OC

Processing Technology 49 (19%) 65-73 CP +lO% A&O, +lO%S

360C 351C

. -

361c ----++-

Fig. 8. The recordings of the electron spin resonance spectra for CP and for CP plus AI,O, and sulfur, close to the degradation temperatures resulting in an ESR signal.

appears to inhibit sintering of the particles and enhance liquefaction [9]. In any case, the ESR results of Fig. 8 show the following dramatic results: with CP alone, the first hint of an ESR signal due to its degradation is seen near 36OC,with a clear signal at 381C whose intensity then increases with increase in temperature. On the other hand, with CP plus A&O, and sulfur, we see a clear ESR signal at 28OC,nearly 100C lower than the thermal case. At higher temperatures, the signal again increases. By carrying out experiments with Al,O, and with sulfur loading alone, we have determined that Al,O, plays no role in the degradation of CP and that the degradation at the lower temperature is only due to elemental sulfur. In addition, H, pressure is not necessary for this degradation since it is observed even in a vacuum-sealed sample as long as sulfur is present. Additional experiments for determining the mechanism of this effect due to sulfur are now in progress. In recent experiments by other researchers [lo- 131, a number of acid catalysts have been found to be effective in the depolymerization and liquefaction of polymers and CP. We will test these catalysts in the near future by our ESR methodology described above. In summary, the results presented here show that PE in the CP samples begins to degrade thermally near 380C and this degradation temperature is lowered by about 100C due to the action of the elemental sulfur. Additional experiments with this and other catalysts are in progress.

MM. Ibrahim et al./ Fuel Processing Technology 49 (1996) 65-73

73

Acknowledgements
This work was supported in part by the U.S. Department of Energy through the Consortium for Fossil Fuel Liquefaction Science, Grant No. DE-FC22-93PC93053.

References
[1] Encyclopedia of Polymer Science and Technology, Vol. 4, Interscience, New York, 1966. [2] L.L. Wall, S.L. Madorsky, D.W. Brown, S. Straus and R. Simha, J. Am. Chem. Sot., 76 (1954) 3430-3437. [3] M.M. Ibrahim and M.S. Seehra, Energy Fuels, 5 (1991) 74-78. [4] M.M. Ibrahim and M.S. Seehra, Am. Chem. Sot., Div. Fuel Chem., Prepr., 37 (1992) 1131-1140. [5] S.L. Aggarwal and G.P. Tilley, J. Polym. Sci., 18 (19.55) 17-26. [6] M.K.I. Ismail and S.L. Rogers, Carbon, 30 (1992) 229-239. [7] J.H. Flynn and L.A. Wall, Polym. Lett, 4 (1%6) 323-328. [8] M.M. Ibrahim and M.S. Seehra, Energy Fuels, 8 (1994) 48-52. [9] F. Derbyshire and T. Hager, Fuel, 73 (1994) 1087-1092. [lo] P. Sivakumar, Heon Jung, J.W. Tiemey and I. Wender, Fuel Process. Technol., 49 (1996) 219-232. [I 11 Z. Feng, J. Zhao, J. Rockwell, D. Bailey, and G. Huffman, Fuel Process. Technol., 49 (1996) 17-30. 1121W.B. Ding, W. Tuntawiroon, J. Liang and L.L. Anderson, Fuel Process. Technol., 49 (1996) 49-63. [13] K. Liu, and H.L.C. Meuzelaar, Fuel Process. Technol., 49 (1996) I-15.

You might also like