Sie sind auf Seite 1von 8

Kinetic

2. Rapid entrained

studies
pyrolysis reactor

of rapid oil shale pyrolysis:


of oil shales in a laminar-flow

Ming-Shing Shen, Alain and Gary J. Morris?

*, Lawrence P. LUI

J. Shadle,

Guo-Qing

Zhangt

Morgantown Energy Technology Center, Morgantown, tDepartment of Mechanical Engineering, West Virginia 26506, USA (Received 26 March 1991; revised 31 May 1991)

WV 26507, USA University, Morgantown,

WV

Rapid pyrolysis of Kentucky New Albany shale was conducted in a laminar-flow entrained reactor (LFER) to obtain a fundamental understanding of thermal reactions, which occur during high-heating-rate retorting processes. The reactor configuration was designed to simplify operation and allow accurate modelhng. Temperature characterization and flow visualization in the LFER were conducted to provide the data necessary to proceed with a kinetic study on the rapid pyrolysis of oil shale. The reactor with gas preheaters was constructed to achieve high particle heating rates and to feed oil shale fine particles generated from beneficiation. The rapid pyrolysis of raw and beneficiated oil shales was carried out in nitrogen at temperatures between 700C and 85OC, with gas preheat temperatures up to 98OC. For each temperature, the sampling probe was set at different positions along the length of the reactor tube to obtain different residence times. A pyrolysis kinetic model of LFER has been developed to calculate the particle heat-up rate and residence time under each set of conditions. Comparisons are presented to evaluate the effects of beneficiation, temperature and heating rate as well as residence time. The effects of particle size and gas environment on heat transfer rates and conversion yields were also studied, and the results were used to validate the heat transfer model and to evaluate the imnact ofdevolatilization behaviour on shale combustion in a circulating fluidized-bed reactor.
(Keywords: oil shale; pyrolysis; kinetic)

Oil shale research at the Morgantown Energy Technology Center has been in high-heating-rate oil shale pyrolysis using flash lamp and entrained flow reactors to simulate the rapid heat-up attained in fluidized-bed combustion and retorting processes. Flash lamp pyrolysis studiesle3 on small micrometre-sized shale particles provided mechanistic insight into the nature of thermal reactions that occur during rapid retorting processes. The overall conversion of shale kerogen to liquids plus gases is higher for flash pyrolysis conditions than for slow-heating processes such as in Fischer Assay and thermogravimetry research. The rate of the initial phase of kerogen decomposition in flash pyrolysis is extremely rapid. Complete conversion of the kerogen in oil shale can be achieved in a millisecond time frame; however, a large portion of the product is gas. Compared to slow-heating processes, flash lamp pyrolysis products: (1) contain more unsaturated gas species (benzene, toluene and xylenes) and heteroatomic liquid species; (2) are higher in heavy molecular weight liquids; and (3) have greater gas to liquid ratios. We found that the controlling reaction pathways that govern liquid quality are determined by the heating rate (incident flux), while the overall conversion and the extent of cracking, polymerization or condensation reactions
Presented at Eastern Oil Shale Symposium, 6-S November 1990. Lexington, KY, USA * Present address: EC&G Washington Analytical Services Center, Inc.. Morgantown Operations, Morgantown Energy Technology Center, 3610 Collins Ferry Road, Morgantown, WV 26507, USA

are determined by the peak temperature (net flu~)~. We found that the liquids from retorting at elevated temperatures were mainly high molecular weight aliphatic components. The process conditions that affect the nature of these products are also expected to affect the nature of the yields and product distribution. We found that shale pyrolysis mechanisms are heating rate dependent: the number of kerogen decomposition pathways are more diverse at high-heating rates than at low-heating rates. The optimum conditions for producing liquids from rapid retorting of oil shale is at a slower heating rate than could be attained in a flash lamp, thus, leading to the development of an entrained reactor. Many studies on the rapid devolatilization of coal have been conducted using entrained reactors4P9; however, the resulting kinetic rates from these and other types of rapid heating experimentsloP13 vary by several orders of magnitude. Several researchers4.5.7.9.12 report that the kinetic rates for coals of different rank were within a factor of 2 to 5 when determined in the same reactor, while the rates obtained for the same coal rank, but from different research groups, varied by several orders of magnitude. Thus, coal type is not the reason for the variations reported. The inconsistency in these rapid heating experiments has been cited as caused by variations in estimations of particle temperature history-. Temperature-dependent variations in the thermal properties and the physical structure of coal are often cited as one potential cause for the uncertainty in the temperature histories of coal in rapid heating

0016-2361/91/11127748 c 1991 Butterworth-Heinemann

Ltd.

FUEL, 1991, Vol 70, November

1277

Kinetic studies of rapid pyrolysis: Table 1 Composition


of samples Amount Colorado Composition (dry) Organic matter Mineral matter. LTA C or8

M-S.

Shen

et al. The beneficiated New Albany shale was prepared by the Mineral Resources Institute, Alabama, using fine grinding followed by froth flotation. The elemental analyses suggested that the beneficiation resulted in a three-fold enrichment over the raw shale. The particle densities were measured using mercury porosimetry, and the pressure required to fill the interstitial particle volume was determined using non-compressive, non-porous glass particles ground to similar sizes according to the technique recommended by Gan and co-workers 5.
Entrainedflow Figure 1 reactor

(wt%) New Albany Beneficiated

Li
H N S (total) Ash Moisture (X as recovered) Particle density (g cm-3) Fischer Assay (I t - ) Mean particle diameter (pm)

21.00 79.00 17.02 5.37 1.98 0.58 0.79 59.79 0.21 1.89 135.60 70.00

13.82 86.18 11.14 0.16 1.26 0.32 5.01 79.43 1.18 2.04 52.50 79.00

35.97 64.03 27.76 0.14 3.09 0.79 4.61 56.69 1.69 1.66 149.20 4.20

experiments 8-9,11 . While the physical structure and thus the physical properties of coal particles are dominated by the formation of a melt, often referred to as metaplast, which flows, swells and becomes glassy at various stages during heatingi4, the physical structure in oil shale is dominated by the mineral matter. This important distinction makes the estimation of thermal and physical properties, such as thermal conductivity, heat capacity, density and diameter, much more reliable, at least when the shale particles contain a normally representative distribution of minerals and kerogen. To study the causes and effects of oil yield enhancement caused by rapid heating, a 5 cm inside diameter, laminar-flow entrained reactor (LFER) was constructed to achieve heating rates approaching those anticipated in a fluidized bed. Fast heat-up and subsequent rapid quench in this system enable the study of the initial stages of retorting. The primary objective is to determine the kinetics for shale devolatilization at short residence times and high-heating rates. Measuring kinetic parameters at rapid heating conditions is necessary since there is some variability in the low-heating rate data, and since extrapolation for rapid heating rate applications can be risky. Rapid heating of oil shale involves complex mass and heat transfer effects and chemical reactions that are dependent on a wide range of parameters, particularly heating rate, final temperature, particle size, shale type and grade. The kerogen in oil shale is such a complicated heterogeneous mixture of organic compounds that the reported pyrolysis kinetics are undoubtedly an average for many different reactions that give the oil product. We present a critical assessment of the factors involved in determining shale pyrolysis rates using the LFER. Enough similarities exist to other analyses for the evaluations made here to be applicable to other rapid devolatilization experiments. EXPERIMENTAL
Samples

Experiments were conducted with Colorado shale, New Albany shale and beneficiated New Albany shale in the LFER. The raw shales were crushed once through a hammer mill, sieved and then aerodynamically cleaned and separated with the turbo classifier at the Fluidization Research Center of West Virginia University. Proximate and ultimate analyses are presented in Table 1.

shows a schematic diagram of the reactor set-up. The geometry of the reactor system was simplified in design to minimize the effect of mixing zones on residence time, and to facilitate modelling the Bow. Oil shale particles were fed from a syringe pump feeder, entrained in a nitrogen carrier gas (primary gas), and introduced to the reaction zone using an injection probe. The syringe pump feeder was modified to accommodate oil shale fine particles generated from beneficiation. One of the reasons that most experimentation in entrained reactors is conducted with particle sizes >40 pm is that static forces that result in cluster formation tend to be much lower in larger particles5v798. It is likely that researchers who failed to observe the expected particle size dependence4 caused by differences in heat transfer rates were feeding large clusters rather than discrete particles. To disrupt these clusters, a porous gas distributor was inserted in the bottom of the solids container and an auxiliary gas line was introduced below the distributor to permit fluidization of the solids. An auxiliary gas vent outlet was installed at the top of the solids reservoir chamber to vent the nitrogen gas required for fluidization in excess of that needed for primary gas injection to the entrained reactor. During operation, the solids feed bed was slumped several times prior to fluidization by pressurizing the solids reservoir and releasing the pressure using a valve in the vent line. This aided the breakdown of particle clumps and allowed even fluidization. The result was a feeder that provided a continuous flow of well-dispersed particles rather than clusters of particles. The injection probe outlet was located at #the inlet of a Lindberg three-zone tube furnace that was maintained at reaction temperature. Preheated gas (secondary gas) was introduced to the reactor zone through a flow straightener; the gas contacted the entrained oil shale at the probe exit to heat the shale and primary gas to the desired temperature. A collection probe was used that traversed the reaction zone to quickly quench the entire product stream. Both the injection probe and collection probe were water-cooled, providing a well-defined residence time within the reactor4. The position of the cooled collection probe was adjusted to control residence times. Immediately downstream from the collection probe was a 0.5 pm ceramic thimble filter used to collect the solid residue. While this configuration permits complete collection of all particles, it has the disadvantage that condensible liquids may deposit back on the spent shale. In fact, there is some evidence that this happens. Thus, the observed weight loss is only the light oil that is not condensed and the heavy oil that deposits on the walls of the

1278

FUEL, 1991,

Vol 70, November

Kinetic

studies

of rapid

pyrolysis.

M.-S.

Shen

et al.

SOLIDS CARRIER GAS

4 A
20 PSI HEATING GAS

-4

-_.

First Preheater L

In

. Vibrator
Second Preheater

ST

Adjusting

Figure 1

Schematic

diagram

of laminar-flow

entrained

reactor

water-cooled collection probe. The amount of oil that condensed on the spent shale and the filter was determined by simply washing with an organic solvent, toluene. The extract was distinguished from what we typically consider as bitumen, because it could be washed out in a Soxhlet apparatus in 2 h, while bitumen would continue to be extracted from shale after 24 h. In an entrained reactor, the environment temperature and the entraining gas flow are critical parameters in determining the shale particle temperature and residence times. Shale particle temperature and shale flow rate must be determined to measure the kinetics of oil shale pyrolysis. In the LFER, the temperatures of the furnace walls and the entraining gas were independently controlled to match reactor gas and wall temperatures. Flow visualization tests ensured that proper inlet design and operating conditions were sufficient to prevent dispersion of shale particles and to ensure uniform treatment conditions for all particles.

available in the secondary gas stream, and so ample heat was available to ensure that there were no heat limitations or requirements to make up significant portions of the heat in the reaction zone. The furnace temperature was set to be isothermal using a cluster thermocouple prior to each test run as presented previously16. The oil shale particles were pyrolysed at 976, 1026, 1077 and 1125 K in a nitrogen atmosphere. For each temperature, separate tests at different residence times were conducted by setting the sampling probe at different positions along the length of the reactor tube. For each set of conditions, the organic conversion rate was determined by using an ash tracer techique. The amount of solid spent shale was collected and weighed and its ash content was determined. The equation for the ash tracer method for calculating weight loss is

(1)
where AW=weight loss per cent (daf basis), A,= proximate ash in dry oil shale and A = proximate ash in dry spent shale. The results were modified from a daf basis to a dmmf basis by correcting for the mineral matter as determined from the low-temperature ash (LTA). However, volatiles that evolve from the minerals during pyrolysis, such as CO, from the decomposition of carbonates or H,O from hydrated minerals, will report as weight loss in this type of analysis.

Shale was fed into the reactor at a rate of -4 g h- . Nitrogen was used as the primary and secondary gas at standard temperature and with 0.4 and 13.0 1 min- pressure, respectively. The gas profiles were characterized using the techniques of Flaxman and Hallett and the operation was established to be steady laminar as described elsewhere16. The sum of the primary gas and the shale represented a thermal load of < 10% of that

FUEL, 1991,

Vol 70, November

1279

Kinetic

studies

of rapid

pyrolysis:

M. -S. Shen

et a I.

The ash recovery efficiency for all of these experiments was >98%. The reproducibility of these measurements was tested by comparing the results of replicate tests at the same conditions. The average standard deviation from four sets of duplicate measurements of the weight loss was k 1.4%, which corresponds to a precision of k 3.9% at the 95% confidence limit.
Particle trajectory

Equation (3) was solved numerically to find the particle temperature, Tp, as a function of time, t. The distance that particles travel, 2, was also a function of time. Particle acceleration was modelled as the sum of the drag force and gravity: d2Z d,=CD;;d,(VgP

I,)+G

(4)

model

The reactor was designed to provide short reaction times without the complications caused by mixing, solid-solid interactions or solid-wall interactions. The geometry of the reactor system was made simple to minimize the effect of mixing zones on the residence time and also to facilitate the modelling of the flow velocity and temperature profiles. Since there was a dilute flow, it was reasonable to consider a single particle as the control volume:

where C, = drag coefficient and G =acceleration of gravity. Equation (4) was solved numerically for distance Z as a function of time. RESULTS AND DISCUSSION
Particle trajectory

particle volume < gas volume

1 10000 >

The particle was approximated as a homogeneous, spherical particle. Fluid and particle properties8-21, including gas density, thermal conductivity, viscosity, and gas and shale heat capacity, were determined as functions of particle temperature. The arithmetic mean of particle temperature and the bulk gas temperature were taken as the temperatures of the fluid boundary layer:

T=Tp+Tp f
-I

(2)

Thermal mixing of the primary and secondary gas streams was assumed to be instantaneous. Particle heat-up was modelled as the sum of the convection plus radiative heat input minus the heat of reaction; as a first approximation, the conversion rate of oil shale pyrolysis was assumed to follow the Arrhenius first-order rate expression discussed previously16. No attempt was made to model the cooling rate of the sample, which was expected to be approximately the same as the heating rate, based on similar heat transfer mechanisms. The gas must first cool by contacting the cold walls but the smaller diameter in the collection probe (relative to that of the reactor) provides better gas contact and relatively high gas convection rates. It is accepted that some reaction could have taken place in the collection probe during the particle quench but this was considered a relatively small error. Convective and radiative heat transfer relationships and the heat of reaction for the control volume were used to calculate the particle temperature during the period immediately following injection into the LFER:

Typical particle velocity profiles in the reaction zone, calculated from Equation (4), are shown in Figure 2 for Colorado shale. The profiles for New Albany shale and beneficiated New Albany shale were similar. The primary jet was introduced at a higher linear velocity than the secondary stream so that initial particle velocity was higher than gas velocity. Because instantaneous equilibration of the primary and secondary gas velocities was assumed, drag forces decelerated the particle until it approached the gas velocity. A minimum velocity is predicted at -0.03 s. The assumption of instantaneous velocity equilibration must be investigated further, since flow visualization tests indicated that the central jet was kept intact and thus may have penetrated into the flow profile of the secondary gas. The particles then accelerated with the gas along the reactor centreline as the laminar velocity profile developed. The particle velocity reached a stady-state value at -0.3 s when the gas flow reached a fully developed state. After this point, the particles were fully entrained and flowing at near the velocity of the gas at -90 cm s-l, while the terminal velocity was much less at only 10-20 cm s-l. New Albany shale equilibrated at slightly higher velocity than Colorado shale.This was because the New Albany shale had a slightly larger particle size and higher density. For beneficated New Albany shale,

where pp = particle density (g cm - 3), C, = particle heat VP =particle volume (cm3), capacity (J cm -3 K-l), h, = convection heat transfer coefficient (J cm- K s- ), d, = particle diameter (cm), cp = particle emissivity (0.9 for oil shale), F,_, =shape factor between the particle and wall and pK = organic density (g cmm3).

50 _I

I 0.0

I 0.1

0.2

I 0.3 Time

I 0.4 (set)

I 0.5

I 0.6

0 17

Figure 2 Predicted particle Colorado shale. Temperature

velocity profile in the reaction zone for (K): a, 976; b, 1026; c, 1077; d, 1125

1280

FUEL,

1991,

Vol 70, November

Kinetic

studies

of rapid

pyrolysis:

M.-S.

Shen et al.

llOOG e 5 5 & r-" f m 4 r L? 500400300 I 0.0 I 0.1 I 0.2 I 0.3 I 0.4 I 0.5 I 0.6 I 0.7 lOOO900aoo700600-

Time (set)

Figure 3 Predicted particle temperature profile in the reaction zone for New Albany shale. Temperature (K): a, 976; b, 1026; c, 1077; d, 1125

the initial particle velocity was higher but equilibrated at lower velocity because of its fine particle size and lower density, reducing the contribution of its terminal velocity. Typical temperature history profiles for shale particles at the radial centre of the reaction zone are displayed in Figure 3. The initial increase in particle temperature of the Colorado shale was greater than that of New Albany shale. This was also because the New Albany shale had a slightly larger particle size and higher density. For the fine, beneficiated shale particles, the initial temperature rise was very fast indeed. The beneficiated shale particles reached the final temperature in - 1 ms because of the small particle sizez2. Analysis of pyrolysis rate data

beneficiated shale particles heated quickly to reach the final temperature, but the depletion of reactive starting material was notably slower than that for the raw shale, especially at the lower temperatures22. The temperature dependence was greater for beneficiated shales, being more in line with extrapolations from the literature at lower reaction temperatures. When the weight loss data is analysed based on total mass of organic material lost, the beneficiated and the raw New Albany shales exhibited comparable devolatilization rates ~ at least at the lowest temperatures. This suggests that vaporization of the products may be controlling the initial devolatilization rate. Such a vaporization process would be more important in the richer particles of the beneficiated shales because of the larger heat requirements and mass fluxes. The conversion rate of Colorado shale was also normalized to its asymptotic yield at 14.7%. The maximum volatile yields corresponded to -70% of the organic matter (Figure 5). Preliminary data on toluene extraction of the spent shale residues indicated that heavy oils condensed on the solid particles either in the gas stream or in the collection thimble. The extraction yields were greater in the Colorado shale than in the New Albany shale. The fit between experimental data and

5 2 E 's 8 ;

0.6 0.7 0.6 0.5

When analysing the weight loss data in terms of a decomposition reaction with first-order dependence on the concentration of shale organic matter, the weight loss was normalized to the maximum yield. In Figures 4 and 5. the maximum yield was assumed to be the asymptotic yield, w* at the highest test temperature. In a single first-order reaction model, the maximum yield is independent of temperature, particle size and heating rate. Although this is not completely consistent with existing evidence on oil shales, this type of analysis was deemed appropriate for comparative purposes. In any case, the asymptotic yield was 13.2% for the New Albany shale. In the absence of other effects, this corresponded to a volatile yield of nearly 95% of the organic matter in a period of < 1 s. At the present time, the distribution of ~olatiles into gases and liquids is not known quantitatlvely. The presence of clays that are known cracking catalysts would favour the formation of gas over liquids. The rate data for the New Albany shale exhibited a general consistency with literature rate data. The experimental data compared with the model curve fits are shown in Figure 4. The experimental data agreed with the model fits within the experimental confidence limits, having a relative standard error of 3.9%; however, the temperature dependence appeared too low and systematic variations existed in the fits. The maximum yield for the beneficiated New Albany shale**, 23.7%, was much higher than that for the raw New Albany shale. The small

0.1 0.0 0 5 10 15 20 25 30 35 40 45 50

-1
55 Distance (cm)

Figure 4 Comparison of experimental data and curves calculated using the kinetic parameters for New Albany shale. Temperature (K): A, 976; n , 1026: +, 1077; 0, 1125 -

3 2 k .g Ii : s E .g 3

0.6 0.7 0.6 0.5 0.4 0.3 0.2 0.1 -1

lb

15

20

25 Distance

30 (cm)

35

40

45

50

55

Figure 5 Comparison of experimental using the kinetic parameters for Colorado 976: n , 1026; +, 1077; 0, 1125

data and curves calculated shale. Temperature (K): A,

FUEL, 1991,

Vol 70, November

1281

Kinetic

studies

of rapid

pyrolysis:

M. -S. Shen

et a I.

calculated curves was not quite as good as for the New Albany shale, possibly because of the complications from mineral carbonate decomposition and heavy oil material condensation. The temperature dependence of the Colorado shale was intermediate between that of the raw and beneficiated New Albany shales. This ordering is consistent with the suggestion that the early stages of rapid shale devolatilization are controlled by the rate of vaporization where there is a dependence on oil shale grade. In addition, the fact that the eastern shale responded favourably to the rapid heating experiments indicated that the pyrolysis mechanism in an aromatic shale follows a set of parallel reaction pathways such that aromatics not immediately evolved undergo retrogressive reactions to form char. On the other hand, the conversion efficiencies in the western shale were not significantly higher than those obtained at slow-heating rates, probably because of the lower coking tendency of these highly aliphatic kerogens. This would suggest that a set of parallel reaction pathways may not be necessary to explain rapid pyrolysis rate data for the western oil shales.
Particle size considerations

The optimum particle size for these raw shales was found to be 170 by 200 mesh 23. At this size fraction, there was little segregation of components and the particles retained a representative mixture of kerogen and minerals. This minimizes the possibility that individual particles of kerogen or minerals will skew the data. In addition, the once-through grinding technique minimizes the potential that kerogen will become fluid because of heat and compression during grinding, and thence deposit as a coating on the external surfaces of particles. This situation has been observed when repeated fine grinding to a small particle size (< 100 pm) was desired. Obviously, the kerogen conversion in such externally coated particles would experience unnaturally large heat and mass transfer rates. Model calculations indicated that particle size was a most important parameter in determining particle temperature history in the LFER. Recent attempts to measure coal particle temperatures in this type of reactor have indicated that the heat-up times are faster than models estimate. This has been assumed to be because the heat transfer coefficient was underestimated; however, recent evidence obtained using temperature and particle size measurements on single particles suspended in an electrodynamic balance is inconsistent with this explanation . One implication from this work is that the heat transfer calculations in all studies that assume spherical particles will grossly overestimate the volume of irregularly shaped, non-spherical particles. Since the spherical particle assumption minimizes the surface area to volume ratio, the calculations result in higher thermal inertia and thus lower heat-up rates. For instance, a cube having a characteristic length (largest dimension) equal to that of a sphere has a nearly equal surface area but only two-thirds the volume, and so it has about a 40% higher area to volume ratio. The spherical-volume equivalent size of the shale particles are reported in Table I as determined using the Coulter counter technique and assuming an ellipsoid geometry. Scanning electron microscopy (SEM) photomicrographs of the shales and subsequent image analysis

indicated that the sphericity (ratio of the sphericalequivalent diameter to the spherical diameter) was 0.96-0.97; however, the SEM is a two-dimensional technique and the third dimension was assumed to be the same as the smallest measured length. In fact, the thickness of these particles could have been significantly smaller than assumed, especially for the raw shale in which the particle size was equal to or larger than the thickness of the bedding layer23. If the net heat flux is comparable to the rate of devolatilization, decomposition takes place during heat-up. Transient heat transfer equations have been solved24 using the Biot number, a dimensionless parameter, to determine the relative resistance for heat transfer externally (to a particle) and internally (within a particle). For example, if a Colorado oil shale (135.6 1 t- ) is heated in a hot nitrogen environment (lOOOC), the error in assuming that the particle temperature is uniform is < 5% when the particle size is < - 125 pm. A characteristic heating time for the intraparticle heating of shale particles of 70 pm is - 5 ms, while that for external shale heating is -80 ms. These calculations indicate that for small non-reacting shale particles, the rates of heat transfer to the particles are much higher than the rates of heat transfer within the particle. On this basis, an intraparticle temperature gradient is insignificant. Experiments on Colorado oil shale with different particle sizes were conducted to evaluate mass and heat transfer effects. The results indicate that as the mean particle diameter reduced from 70 to 63 to 60 pm, the conversion (weight loss fraction based on shale) increased from 12.0 to 17.2 to 18.7% at a reactor temperature of 800C. At 85OC, the conversion increased from 14.7 to 18.1 to 19.7%, respectively. This magnitude is not inconsistent with the effects expected because of external heat transport to the particle. Smaller particles had longer residence times because of slower particle velocities. Smaller particles heat up more quickly than bigger particles at the same residence time. Differences arise between experimental data and model calculations for the small particles only. These could be a result of several things. First of all, the model assumes a single particle as a control volume, but there might be a cluster formation. The model also assumes the primary gas heated up instantaneously and does not take into account the heat transfer of primary gas around the particle. Furthermore, these particle sizes may be smaller than the thickness of the bedding layer, causing some segregation of kerogen and mineral components, which would, in turn, affect the heat and mass transfer rates, and, thence, the kerogen conversion.
Additional heat tramfer considerations

Experiments on Colorado oil shale with helium gas were also conducted to evaluate mass and heat transfer effects. The results indicated that using helium as the carrier gas resulted in an increased weight loss of 6% at 800C and an increase of 14.2% at 850C. Helium has a smaller density, lower viscosity and higher heat conductivity than nitrogen. The smaller density and viscosity will generate a shorter residence time, while the higher heat conductivity will produce a higher heating rate. The higher heating rate and shorter residence time would enhance the pyrolysis conversion rate. The trend

1282

FUEL,

1991,

Vol 70, November

Kinetic

studies

of rapid

pyrolysis:

M.-S.

Shen

et al.

of experimental results was in agreement with the model calculation results. However, the weight loss rate predicted by the model was greater than that of experimental results, since the nitrogen gas was used as a secondary gas in these experimental tests. The maximum potential conversion obtained in helium or with the smaller particles was larger than that generated for 70 pm particles in nitrogen. The LFER computer model was used to conduct a sensitivity analysis on particle size and gas atmosphere. The model can account for the differences in the time-temperature profile and, therefore, the changes in the conversion rate. The model can accurately simulate the data for time dependence of conversion at a single temperature. However,the maximum conversion had to be adjusted to match the experimental values for the different particle sizes or gas atmospheres. The ultimate organic conversion (or the maximum recoverable organic fraction in the shale studied) had to be varied for the different particle sizes and gas atmospheres. This suggests that the changes in heating conditions altered the extent of retrogressive reactions such that higher heating rates and mass transfer rates favoured higher conversion levels even at these short residence times (0.54.6 s).

bonds to aliphatic CH groups did not change significantly with temperature. CONCLUSIONS The techniques applied to investigate rapid oil shale pyrolysis have provided an insight into the reaction mechanisms. The conversion of oil shales has been demonstrated to be an extremely rapid process, with maximum yields being achieved in < 1 s at temperatures near 800C. The conversion of the beneficiated shale exhibited a greater temperature dependence than the feed shale. The weight loss data suggested that vaporization may play an important role in the initial stages of rapid devolatilzation. If vaporization is a rate-controlling process, the implications will be quite important in the design of processes in which shales of various grades are rapidly devolatilized. The rate data for the raw shales indicated that eastern and western shales exhibited different reaction mechanisms: the conversion efficiency in the eastern shale depended strongly on heating rate while the western shale did not. Thus, different reaction models may be necessary to describe the rapid devolatilization of these two shales. The LFER experiments require further improvement, especially with regard to the treatment of the particle size and shape determination, the heat transfer and velocity calculations in the injection region of the reactor, particle cluster formation in the reactor, determination of the product yield, and separation of solids and vapours. Further validation of the particle trajectory model is desired and measurements of particle residence times, temperatures, velocities and size are required. In addition, further work is required on gas and oil generation/separation to determine carbon distribution in char, gases and liquids. The effect of particle size and heat transfer rates (using gases of different thermal conductivity) on rates and yields should be studied in more detail, and the results used to validate the heat transfer model. Kinetic analysis of the data is being expanded to include an analysis of the shale conversion as a number of parallel first-order reactions in which the measured products are lumped together, and the overall activation energy can be estimated as a function of conversion. In this method, the relationship between the log of the time required to reach a fixed level of conversion and the reciprocal temperature is used to determine the activation energy at that level of conversion. This analysis provides a straightforward determination of the error limits on activation energy and bounds the activation energy whereas other approaches do not. It also provides a method for determining the fraction of each component in the reacting system in cases where the activation energy varies with conversion. REFERENCES
1 Shadle, L. J.. Gaston, M. H. and Rosencrans, R. D. Proc. of the 1985 Eastern Oil Shale Symposium, Kentucky Center for Energy Research Laboratory, Lexington, 1985 Shadle, L. J., Rosencrans, R. D., Shen, M. S. cr al. Proc. of the International Conference on Oil Shale and Shale Oil, Chemical Industry Press, Beijing, 1988 Shadle, L. J., Hobbs, G. R., Shen, M. S. et al. Flash Pyrolysis of Green River Shale, Technical Note. DOE/METC-88/4080, NTIS/DE88001077, 1987, p. 34

Liquid

product

composition

Part of the pyrolysis products from LFER was retained on the spent shale, and this part was Soxhlet extracted with toluene. The toluene extract was analysed using FT-i.r.. In the toluene extract of the New Albany spent shale, the relative aromatic concentration appeared to be high. There were also many C=O and C-O structures. These observations demonstrated that the LFER pyrolysis liquid products had similar chemical character to asphaltenes (hexane insolubles). The fraction of asphaltenes in toluene solubles increased with temperature. This was attributed to the formation of condensed and non-condensed polycyclic aromatic compounds from less complex aromatic compounds as the pyrolysis temperature increased. For a given temperature, the relative amount of oils in toluene solubles increased with decreasing residence time. For a given residence time and temperature, the fraction of oils in toluene solubles was in the order: Colorado shale > beneficiated New Albany > New Albany shale The oils from LFER samples were analysed by FT-i.r.. It was observed that ester-type structures were rather prominent in LFER oils from New Albany shale, while Colorado shale oils contained a weak ester peak only at 850C. This may be attributed to the higher (almost twice) organic oxygen content of New Albany shale than Colorado shale. For the New Albany shale, the relative concentration of the C=O bonds to aliphatic CHs increased with temperature. This may have resulted from reduced decarboxylation of the ester groups under higher heating rates that prevailed in the reactor. In a slow-heating-rate process, such as a Union B Retort, the ester structures would be readily thermolysed. For the beneficiated New Albany shale, the C=O bond stretching mode appeared at higher frequency indicating that the R group in the ester RCOOR was more alkyl than aryl in nature. The relative concentration of C=O

FUEL,

1991,

Vol 70, November

1283

Kinetic studies of rapid pyrolysis:


4 5 6 7

M-S.

Shen et al.
17 18 19 20 Perry, R. H. and Chilton, C. H. Chemical Engineers Handbook, 5th Edn, McGraw-Hill Book Co., New York, 1973 Camp, D. W. Proc. of the 1987 Eastern Oil Shale Symposium, Kentucky Energy Cabinet Laboratory, Lexington, 1987 Bauahman. G. L. Svnthetic Fuels Data Handbook, 2nd Edn, Cameron Engineers,*Inc., Denver, 1978 Johnson, D. R., Smith, J. W. and Young, N. B. Stratigraphic Variation of Oil Shale Enthalpy of Retorting Through the Green River Formation on the Colorado C-a Tract, LETC/RI-79/9, 1979 Shaw, R.J. Specific Heat of Colorado Oil Shale, USBM Report no. 4151, 1947 Shen, M. S., Shadle, L. J. and Zhang, G. Q. Proc. of the 1990 Oil Shale and Tar Sand Contractors Review Meeting, Morgantown Energy Technology Center, Morgantown, 1990 Grimm, U. and Shadle, L. J. Chemical Separations. 11. Applications, Litarvan Literature, Denver, 1986 Holman, J. P. Heat Transfer, 5th Edn, McGraw-Hill, New York, 1981 Szladow, A. J. and Given, P. H. Ind. Eng. Chem. Des. Deo. 1981, 20, 27

8 9 10 11 12

13 14 15 16

Badzioch, S. and Hawksley, P. G. W. Ind. Eng. Chem. Des. Dev. 9, 521 Kobayashi, H. PhD Thesis Massachusetts Institute of Technology, 1976 Flaxman, R. J. and Hallett, W. L. H. Fuel 1987, 66, 607 Maloney, D.J. and Jenkins, R. G. Proc. of the Twentieth Symposium on Combustion, The Combustion Institute, Pittsburgh, Ann Arbor, 1984, p. 1435 Fletcher, T. H. Cornbust. Flame 1989, 78, 223 Solomon, P. R., Serio, M. A., Carangelo, R. M. and Markham, J. R. Fuel 1986,65, 182 Freihaut, J. D. and Proscia, W. M. Energy &Fuels 1989,3,625 Maloney, D. J., Monazam, E. R., Woodruff, S. D. el al. Combusl. Flame in press Anthony, D. B., Howard, J. B., Hotell, H. C. er al. Proc. of the Fifteenth Symposium (Int.) on Combustion, The Combustion Institute, Pittsburgh, 1975, p. 1303 Hertzberg, M. and Zlochower, I. A. Combusr. Flame in press Van Krevelen, D. W. Coal, Elsevier, Amsterdam, 1961 Gan, H., Nandi, S. P. and Walker Jr., P. L. Fuel 1987, 51, 51 Shen, M. S., Zhang, G.-Q. and Shadle, L. J. Thermochim. Acta 1989, 154, 355

21 22

23 24 25

1284

FUEL,

1991,

Vol 70, November

Das könnte Ihnen auch gefallen