Вы находитесь на странице: 1из 9

1250

Energy & Fuels 2006, 20, 1250-1258

Kinetic Modeling for the Combined Pyrolysis and Steam Gasification of Petroleum Coke and Experimental Determination of the Rate Constants by Dynamic Thermogravimetry in the 500-1520 K Range
D. Trommer and A. Steinfeld*,,
Department of Mechanical and Process Engineering, ETH-Swiss Federal Institute of Technology Zurich, 8092 Zurich, Switzerland, and Solar Technology Laboratory, Paul Scherrer Institute, 5232 Villigen, Switzerland ReceiVed September 7, 2005. ReVised Manuscript ReceiVed February 9, 2006

The detailed kinetic mechanisms for the combined endothermic pyrolysis and steam gasification of petroleum coke are formulated using elementary chemisorption and reaction steps. The pyrolysis model is based on a linear combination of first-order decomposition rates of pseudo components. The gasification model is based on the oxygen-exchange mechanism describing reversible O-transfer surface reactions followed by a unidirectional gasification step, and on an extended mechanism describing, in addition, reversible steam sorption as OH/H groups and irreversible surface chemistry. The kinetic rate laws of Langmuir-Hinshelwood type are derived by assuming sorption equilibrium and by considering inhibition due to recombination of adsorbed oxygen and competition for active sites. The Arrhenius parameters of the rate constants for the pyrolysis and gasification of Flexicoke and Petrozuata Delayed coke are experimentally determined by dynamic thermogravimetric analyses with Ar, H2O, CO2, and H2O-CO2 mixtures in the range 500-1520 K.

Introduction Hybrid solar/fossil endothermic processes, in which fossil fuels are used exclusively as the chemical source for H2 production and concentrated solar power is used exclusively as the energy source of process heat, offer a viable route for fossil fuel decarbonization and create a transition path toward solar hydrogen.1 An important example of such hybridization is the endothermic steam gasification of petroleum coke (petcoke) to synthesis gas (syngas). The advantages of supplying solar energy for process heat are 4-fold: (1) the calorific value of the feedstock is upgraded, (2) the gaseous products are not contaminated by the byproducts of combustion, (3) the discharge of pollutants to the environment is reduced, and (4) the need for energy-intensive processing of pure oxygen is eliminated. A 2nd-Law analysis for generating electricity using the solar gasification products indicates the potential of doubling the specific electrical output and, consequently, halving the specific ` CO2 emissions, vis-a-vis conventional petcoke-fired power plants.2 The overall chemical process can be represented by the simplified net reaction:

CHxOy + (1 - y)H2O )

x (2 + 1 - y)H + CO
2

(1)

where x and y are the elemental molar ratios of H/C and O/C in petcoke, respectively. Sulfur compounds and other impurities
* Corresponding author. E-mail: aldo.steinfeld@eth.ch. ETH-Swiss Federal Institute of Technology Zurich. Paul Scherrer Institute. (1) v. Zedtwitz, P.; Steinfeld, A. Energy (Amsterdam, Neth.) 2003, 28, 441-456. (2) Trommer, D.; Noembrini, F.; Fasciana, M.; Rodriguez, D.; Morales, A.; Romero, M.; Steinfeld, A. Int. J. Hydrogen Energy 2005, 30, 605618.

contained in the raw materials are omitted from eq 1; these are of course important, but their exclusion does not affect the main conclusions. The moisture content is also not accounted for in eq 1, but the stoichiometric addition of water can be adjusted accordingly. The approximate main elemental chemical composition, the low heating value (LHV), and elemental molar ratios of H/C and O/C are listed in the Appendix for two different types of petcoke: Flexicoke and Petrozuata Delayed (PD) coke. The chemical thermodynamics indicate that the net process is endothermic by about 50% of the feedstocks LHV and proceeds at above 1300 K to produce, in equilibrium, an equimolar mixture of H2 and CO and less than 5% CO2. The solar chemical reactor technology, featuring a continuous vortex flow of steam laden with petcoke particles and directly exposed to concentrated solar radiation, was demonstrated at a power level of 5 kW in a high-flux solar furnace.3 The steam gasification of petcoke is characterized by the simultaneous occurrence of pyrolysis (i.e., thermal decomposition in the absence of a gasifying agent) and a combined H2OCO2 gasification. Previous kinetic models describing pyrolysis include chemical and physical processes among carbonaceous structures,4 irreversible and first-order decomposition reactions with continuously distributed activation energies,5-7 and a linear combination of parallel, independent decompositions of pseudo components.8,9 Previous kinetic models describing gasification
(3) ZGraggen, A.; Haueter, P.; Trommer, D.; Romero, M.; de Jesus, J. C.; Steinfeld, A. Int. J. Hydrogen Energy, in press. (4) Migliavacca, G.; Parodi, E.; Bonfanti, L.; Faravelli, T.; Pierucci, S.; Ranzi, E. Energy (Amsterdam, Neth.) 2005, 30, 1453-1468. (5) Ulloa, C.; Gordon, A. L.; Garca, X. J. Anal. Appl. Pyrolysis 2004, ` 71, 465-483. (6) Pitt, G. J. Fuel 1962, 41, 267-274. (7) Anthony, D. B.; Howard, J. B. AIChE J. 1967, 22, 625-656. (8) Meszaros, E.; Varhegyi, G.; Jakab, E.; Marosvolgyiet, B. Energy ` ` Fuels 2003, 18, 497-507.

10.1021/ef050290a CCC: $33.50 2006 American Chemical Society Published on Web 03/28/2006

Combined Pyrolysis and Steam Gasification of Petcoke

Energy & Fuels, Vol. 20, No. 3, 2006 1251

include those based on the oxygen-exchange mechanism for gasification with H2O, CO2, and H2O-CO2 mixtures at below or near atmospheric pressure,10-16 and those that, in addition, consider reversible sorption mechanisms.2,17-21 This article presents the formulation of kinetic models for the pyrolysis and steam gasification of petcoke in the absence of O2, the detailed derivation of the rate laws, and the experimental determination of the rate constants by thermogravimetry. Kinetic Models Pyrolysis. The rate of pyrolysis is calculated as a linear combination of first-order decomposition rates of pseudo components:

from the solid to form CO(g):

CO 9 CO + nC* 8

k3

(7)

Inhibition occurs by the recombination of adsorbed oxygen with either CO or H2. Equation 6 typically controls the movement toward equilibrium, while eqs 5 and 6 provide for the watergas shift equilibrium. Assuming elementary reactions, the rate laws for the formation/consumption of the gas species H2O, CO2, CO, and H2 are obtained as functions of their partial pressures in the bulk phase and the fraction of surface covered with adsorbed oxygen, O:

rH2O ) k-2pH2O - k2pH2O(1 - O) rCO2 ) k-1pCOO - k1pCO2(1 - O) rH2 ) k2pH2O(1 - O) - k-2pH2O rCO ) k1pCO2(1 - O) - (k-1pCO - k3)O

(8) (9) (10) (11)

dXP dt
where

ciri ) ci dt i i

dRi (2)

dRi ri ) ) ki(1 - Ri) dT

(3)

where ri ) 1 dni mpetcoke dt (12)

and the temperature dependence of the rate constant is given by the Arrhenius law:

ki ) k0,i exp

- EA,i RT

(4)

ri, with i ) H2O, CO2, CO, and H2, is expressed in units of moles of i transformed per time per mass of solid, mol/(g s). Further assuming sorption equilibrium yields:

Gasification. Two mechanisms for the gasification of carbon with H2O and CO2 are considered: the oxygen-exchange and the extended mechanisms. The rate constants of the oxygen exchange mechanism are denoted ki, and the rate constants of the extended mechanism are denoted ki. The oxygen-exchange mechanism postulates reversible oxygen-exchange surface reactions

O )

k1pCO2 + k2pH2O k1pCO2 + k-1pCO + k2pH2O + k-2pH2 + k3

(13)

Substituting O in eqs 8-11 gives the rate laws in terms of the partial pressures and temperature:

\ CO2 + C* {k } CO + CO
-1

k1

(5) (6)

ri )

Ni 1+ 1 (k p + k-1pCO + k2pH2O + k-2pH2) k3 1 CO2

(14)

\ H2O + C* {k } H2 + CO
-2

k2

with the species-dependent numerators Ni being

followed by the unidirectional gasification step at higher temperatures through combination of oxygen and carbon atoms
(9) Porada, S. Fuel 2004, 83, 1191-1196. (10) Laurendeau, N. M. Prog. Energy Combust. Sci. 1978, 4, 221-270. (11) Fredersdorff, C. G. v.; Elliott, M. A. Coal Gasification. In Chemistry of coal utilization; Lowry, H. H., Ed.; Wiley: New York, 1963; pp 8921022. (12) Gadsby, J.; Long, F. J.; Sleightholm, P.; Sykes, K. W. Proc. R. Soc. London 1948, A193, 357-376. (13) Reif, A. E. J. Phys. Chem. 1952, 56, 785-788. (14) Ergun, S. Kinetics of the reactions of carbon dioxide and steam with coke; Bulletin 598; U.S. Bureau of Mines: Washington, DC, 1962; pp 1-38. (15) Mentser, M.; Ergun, S. A study of the carbon dioxide-carbon reaction by oxygen exchange; Bulletin 664; U.S. Bureau of Mines: Washington, DC, 1973; pp 1-42. (16) Gadsby, J.; Hinshelwood, C. N.; Sykes, K. W. Proc. R. Soc. London 1946, A187, 129-151. (17) Muller, R.; Zedtwitz, P. v.; Wokaun, A.; Steinfeld, A. Chem. Eng. Sci. 2003, 58, 5111-5119. (18) Long, F. J.; Sykes, K. W. Proc. R. Soc. London 1948, A193, 377399. (19) Blackwood, J. D.; McGrory, F. Aust. J. Chem. 1957, 11, 16-33. (20) Muhlen, H. J. Zum Einfluss der Produktgase auf die Kinetik der WasserdampfVergasung in Abhangigkeit Von Druck und Temperatur; Universitat Essen (D), 1983. (21) Muhlen, H. J.; Heek, K. H. v.; Juntgen, H. Fuel 1985, 64, 944 949.

NH2O ) -k2pH2O -

k2k-1 k1k-2 p p + p p k3 H2O CO k3 H2 CO2 NH2 ) -NH2O

(15) (16) (17)

NCO2 ) -k1pCO2 +

k2k-1 k1k-2 pH2OpCO p p k3 k3 H2 CO2

NCO ) 2k1pCO2 + k2pH2O -

k2k-1 k1k-2 pH2OpCO + p p k3 k3 H2 CO2 (18)

The rate for the carbon consumption is obtained by applying carbon mass conservation:

rC ) -rCO - rCO2 ) -k3O ) k1pCO2 + k2pH2O (19) 1 1 + (k1pCO2 + k-1pCO + k2pH2O + k-2pH2) k3
Since the gaseous products CO and H2 are continuously withdrawn from the reaction zone, the reverse reactions of eqs

1252 Energy & Fuels, Vol. 20, No. 3, 2006

Trommer and Steinfeld

5 and 6 are neglected (k-1 ) k-2 ) 0), leading to

k1pCO2 + k2pH2O rC ) 1 1 + (k1pCO2 + k2pH2O) k3

(20)

The extended mechanism assumes that H2O undergoes dissociation at the carbon surface into hydroxyl radical and hydrogen atom groups, both being rapidly chemisorbed at adjacent active sites, followed by a hydrogen transfer to the adsorbed hydrogen,18-21 and that CO2 reacts by oxygen exchange with the solid surface according to the previous mechanism. Inhibition occurs by recombination of CO with adsorbed oxygen and by H2/CO adsorption on active sites. Considering reversible sorption of gaseous species onto the carbon surface and irreversible reactions among adsorbed species and with molecules in the gas phase, we get

In the mathematical treatment that follows, eqs 28 and 29 are omitted from consideration since the reduced system represents data reasonably well in the 0-1 bar pressure range used in this study. CO inhibition can occur both by recombination with an adsorbed oxygen (eq 21 reverse) and by competition for active sites (eq 22). Further assuming sorption equilibrium for CO, O, H2, and the intermediate component OH-H representing the adsorbed water,

OH,H: k4pH2Ov ) (k-4 + k5)OH,H CO: k2pCOv ) k-2CO H2: k6pH2v + k5OH,H ) k-6H2 O: k5OH,H + k1pCO2v ) k-1pCOO + k3O
with

(30) (31) (32) (33)

\ CO2 + C* {k } CO + CO
-1

k1

(21) (22) (23) (24) (25) (26)

v ) 1 - (OH,H + CO + H2 + O)

(34)

CO + C* {k } CCO \
-2

k2

Assuming elementary reactions, the rate laws for the formation and consumption of the gas species H2O, H2, CO, and CO2 are obtained as functions of their partial pressures and i:

8 CO 9CO + nC* H2O + C* {k } COH + CH \


-4

k3

k4

rH2O ) k-4OH,H - k4pH2Ov rH2 ) k-6H2 - k6pH2v

(35) (36)

8 COH + CH9 CO + CH2 \ H2 + C* {k } CH2


-6

k5

rCO ) k1pCO2v - k-1pCOO + k-2CO - k2pCOv + k3O (37) rCO2 ) k-1pCOO - k1pCO2v (38)

k6

Equations 21-23 are pertinent to CO2 gasification, whereas eqs 23-26 are pertinent to H2O gasification, with eq 23 describing desorption of the surface oxygen complex common to both H2O and CO2 gasification. Gadsby et al.12 proposed the system of eqs 21 (forward), 22, and 23, in which CO inhibition occurs due to adsorption of CO on the carbon surface, competing with CO2 for active sites. Ergun and co-workers14,15 proposed the system of eqs 21 and 23, in which CO inhibition occurs when surface oxygen is detached by CO from the gas phase (eq 21 reverse). Both mechanisms lead to the same form of the rate law for carbon consumption by gasification using only CO2 as gasifying agent:

The system of eqs 30-34 is solved analytically for the unknown i, and the solution inserted in eqs 35-38 to yield the rate expressions of Langmuir-Hinshelwood type is

ri ) {Ni} 1 +

/{

k1 k4(k5k-6 + k3k-6 + k3k5) pCO2 + pH2O + k3 k3k-6(k-4 + k5)

k6 k -1 k-1k2 2 k2 pH 2 + + pCO + p + k-6 k-2 k3 k-2k3 CO k-1k6 k-1k4 k5 + k-6 p p + p p k3k-6 k-4 + k5 H2O CO k3k-6 H2 CO
with the following numerator terms:

(39)

rC ) 1 + a2pCO2 + a3pCO

a1pCO2

(27)

k4k5 k-1 k4k5 N H2 O ) pH2O p p k-4 + k5 k3 k-4 + k5 H2O CO NH2 ) -NH2O NCO ) 2k1pCO2 +

(40) (41)

with a1, a2, and a3 being lumped rate constants. Reif13 analyzed in detail the importance of the two CO inhibition paths described above and found experimentally that the recombination of CO and CO is considerably faster than CO adsorption. For gasification at elevated pressures, Blackwood and Ingeme22 and Muller et al.17 extended Erguns mechanism by adding

k4k5 k-1 k4k5 pH 2 O p p k-4 + k5 k3 k-4 + k5 H2O CO (42) k-1 k4k5 p p k3 k-4 + k5 H2O CO (43)

8 CO2 + CCO92CO + CO 8 CO + CCO9CO2 + 2C*


k8

k7

(28) (29)

NCO2 ) -k1pCO2 +

(22) Blackwood, J. D.; Ingeme, A. J. Aust. J. Chem. 1960, 13, 194209.

Equation 39 represents the complete solution of the reactions described by the system of eqs 21-26. Neglecting the mixed and squared terms in the denominator since they have negligible effect in the low-pressure range and replacing the elementary

Combined Pyrolysis and Steam Gasification of Petcoke

Energy & Fuels, Vol. 20, No. 3, 2006 1253

rate constants by lumped complex constants yields

r H2 O )

-K2pH2O - K2K3pH2OpCO 1 + K4pCO2 + K5pH2O + K6pH2 + K7pCO rH2 ) -rH2O

(44) (45) (46)

rCO ) rCO2 )

2K1pCO2 + K2pH2O - K2K3pH2OpCO 1 + K4pCO2 + K5pH2O + K6pH2 + K7pCO -K1pCO2 + K2K3pH2OpCO 1 + K4pCO2 + K5pH2O + K6pH2 + K7pCO

(47)
Figure 1. Schematic of the thermogravimeters steam furnace for experiments up to 1523 K with reactive gas containing up to 100% steam.

where the Ki values are defined as

K 1 ) k1 K2 ) k4k5 k-4 + k5 k-1 k3 k1 k3 (48)

K3 )

K4 ) K5 )

k4(k5k-6 + k3k-6 + k3k5) k3k-6(k-4 + k5) k6 K6 ) k-6 K7 ) k2 k-1 + k-2 k3

The temperature dependence of the rate constants derived by the oxygen-transfer mechanism (eq 20) and by the extended mechanism (eq 50) is given by the Arrhenius law (eq 4). Active Surfaces. Being a heterogeneous process, the rate of the reaction is strongly affected by the availability of active surface, which in turn depends on, among other variables, the particle size, porosity, temperature, reaction extent, and impurities contained in the feedstock. The following proportionality is used:

rC a

()
1 dP

(52)

where a is the specific active surface per unit mass of petcoke, and n is determined experimentally for each type of petcoke. Limiting cases are n ) 1 for spherical particles, with the reaction taking place only on the outer surface, and n ) 0 for particles, with the reaction taking place in the entire volume. Thermogravimetric Analysis Experimental Setup. Experimentation was carried out on a thermogravimeter system (TG, Netzsch STA 409 CD) equipped with two furnaces: a conventional high-temperature electric furnace with a maximum working temperature of 1823 K and suitable for reactive atmospheres having a dew point below room temperature, and a special electric furnace with a maximum working temperature of 1523 K and suitable for reactive atmospheres containing up to 100% steam at 1 bar total pressure. The latter is shown schematically in Figure 1, as used for the gasification runs with H2O-CO2-Ar mixtures. The reactive gas entered the furnace chamber from the top and flowed downward past a thin layer of petcoke sample mounted on a 17-mm diameter flat sample holder. This arrangement avoided stagnant gas around the sample and minimized mass/heat transfer resistances between the coke sample and the bulk gas.24,25 The steam furnace was coupled with a steam generator unit (Bronkhorst Hitec CEM) via a transfer line heated to 473 K, for avoiding steam condensation. The mass flow rates and the steam concentration in the reactive gas were adjusted by mechanical flow controllers for Ar and CO2 (Vogtlin Q-FLOW) and electronic flow controllers for water (Bronkhorst LIQUIFLOW). Original petcoke was milled and sieved for
(24) Ollero, P.; Serrera, A.; Arjona, R.; Alcantarilla, S. Fuel 2002, 81, 1899-2017. (25) Gomez-Barea, A.; Ollero, P.; Arjona, R. Fuel 2005, 84, 1695 1704.

The rate of carbon consumption is calculated in analogy to the oxygen-exchange mechanism by applying carbon mass conservation:

rC ) -rCO - rCO2 ) -k3O ) K1pCO2 + K2pH2O 1 + K4pCO2 + K5pH2O + K6pH2 + K7pCO (49)

Further neglecting the inhibition effect of the product gases in eqs 44-47 and 49, the final expression for the rate of carbon conversion becomes

K1pCO2 + K2pH2O rC ) 1 + K4pCO2 + K5pH2O

(50)

Note that eq 20 obtained by applying the oxygen-transfer mechanism is a special case of eq 50, with the additional boundary conditions

K4 )

K1 K2 and K5 ) a4 a4

(51)

reducing the number of independent parameters from four in eq 50 to only three in eq 20. Note also that eq 39 does not consider the dual-site mechanism or the two-center adsorption.23
(23) Walas, S. M. Reaction Kinetics for Chemical Engineers; Butterworths: Boston, 1989.

1254 Energy & Fuels, Vol. 20, No. 3, 2006

Trommer and Steinfeld

represents only 0.33% weight for PD coke and 0.40% weight for Flexicoke. Thus, the measured rate of gasification is based on the weight loss data recorded by the TG:

-rC )

1 dXG 1 - XG dT

(55)

Figure 2. TG and DTG data for the pyrolysis of Flexicoke and PD coke. dp ) 250-355 m, ) 10 K min-1.

obtaining samples of different particle diameters using different grinding techniques. A cutting mill was used to obtain particles in the 80-1000-m range, which were sieved through a tower with 80-, 160-, 200-, 250-, 355-, and 500-m sieves. A ball and jet mills were used to get powder with mean particle diameters of 3.8 and 1.3 m, respectively. Samples of 40 mg of petcoke were placed as thin layers on the flat sample holder. The TG furnace was first evacuated and purged with Ar three times to obtain an air-free atmosphere. Thereafter, the sample was heated to the desired temperature at a rate of 10 or 20 K/min, under a flow of the reactive gas H2O, CO2, or H2OCO2 mixture, at a total pressure of 1 bar. Product gas composition at the furnace exit was analyzed by gas chromatography (two-channel Varian Micro GC, equipped with a Molsieve-5 and a Poraplot-U columns for the detection of H2, CO, CO2, CH4, C2H2, C2H4, C2H6, H2S, and COS). Experimental Results. On the basis of proximate analysis, we subdivided the coke in two fractions: volatiles and fixed carbon. Both are assumed to participate in the gasification reaction, but only the volatiles are subjected to pyrolysis. The total weight loss of the coke sample is defined as the sum of the pyrolysis and gasification weight loss:

X ) XG(1 - XP) + XP
where

(53)

X(T) ) 1 -

m(T) m0

(54)

XP and XG are the conversions due to pyrolysis and H2O-CO2 gasification, respectively (0 e XG e 1, 0 e XP e ci). Adjustment due to the ash content is not necessary because it

Since pyrolysis and H2O-CO2 gasification occur simultaneously, a set of experiments using only inert gas was performed, and the resulting weight loss curve was subtracted from that obtained using reactive gases H2O/CO2 for calculating the rate of H2O/CO2 gasification. The TG and DTG curves obtained for the pyrolysis of Flexicoke and PD coke are shown in Figure 2. The experimental runs were performed with 250-355-m particles, subjected to an Ar flow of 210 mL/min and a linear heating rate of 10 K min-1 in the 500-1520 K temperature interval. Both curves indicate pyrolysis in a two-step process: (1) at 956 and 860 K, producing a weight loss of 1.7 and 8.4% for Flexicoke and PD coke, respectively, and (2) at above 1090 and 1210 K, producing a weight loss of 6.7 and 4% for Flexicoke and PD coke, respectively. Total weight loss during pyrolysis up to 1520 K is 8.4% for Flexicoke and 12.4% for PD coke. Gases evolved during the first step of the pyrolysis of Flexicoke, shown in Figure 3a, include CO2 (with a peak about twice as much as the one for PD coke) and small amounts of H2. In contrast, gases evolved during the first step of the pyrolysis of PD coke, shown in Figure 3b, include a C2H4 peak at 773 K, C2H6 at 796 K, CH4 at 846 K, CO2 at 839 K, and H2 at 1007 K. At temperatures above 1250 K, both types of petcoke released CO, CO2, and H2. The results for PD coke are similar to those typically observed during pyrolysis of coal and other types of coke.26,27 The weight loss and product gas composition as a function of temperature for the steam gasification of Flexicoke and PD coke are shown in Figure 4. The experimental runs were performed with 250-355-m particles subjected to a 60% H2O-Ar mixture and a linear heating rate of 10 K min-1 in the 500-1520 K interval. The curves represent the superposition of pyrolysis, taking place above 650 K, and steam gasification, starting at about 1000 K. As expected for Flexicoke, pyrolysis is negligible compared to steam gasification. In contrast, for PD coke, it affects both the weight loss and gas composition. Further, the gasification of PD coke appears to proceed following two regimes, a faster one up to 1310 K, and a slower one afterward due to a thermally induced deactivation. In addition to the pyrolysis gas products already considered in Figure 3, the main gasification products are H2, CO, and CO2.

Figure 3. Weight loss and product gas composition as a function of temperature for the pyrolysis of (a) Flexicoke and (b) PD coke. dp ) 250-355 m, ) 10 K min-1.

Combined Pyrolysis and Steam Gasification of Petcoke

Energy & Fuels, Vol. 20, No. 3, 2006 1255

Figure 4. Weight loss and product gas composition as a function of temperature for the steam-gasification of (a) Flexicoke and (b) PD coke. pH2O ) 0.6 bar, dp ) 250-355 m, ) 10 K min-1.

Figure 5. Weight loss as a function of temperature for the gasification of (a) Flexicoke, and (b) PD coke in binary H2O-Ar and CO2-Ar mixtures. dp ) 250-355 m, ) 20 K min-1.

Figure 6. Weight loss as a function of temperature for the steam-gasification of (a) Flexicoke and (b) PD coke. The parameter is the mean particle size. pH2O ) 0.6, ) 10 K min-1.

At above 1210 K, steam transforms the sulfur content of petcoke into H2S. CH4 concentration is less than 0.013% for PD coke and less than 0.02% for Flexicoke. The effect of steam and CO2 concentrations is shown in Figure 5a,b for Flexicoke and PD coke, respectively, obtained from a series of TG runs with 250-355-m particles subjected to binary H2O-Ar and CO2-Ar mixtures in the 0-100% H2O/ CO2 concentration range, and a linear heating rate of 20 K min-1 in the 500-1520 K temperature range. The simultaneous occurrence of pyrolysis clearly affects the H2O/CO2 gasification of PD coke. It affects as well the gasification of Flexicoke, but
(26) Karcz, A.; Porada, S. Fuel 1995, 76, 806-809. (27) Kocaefe, D.; Charette, A.; Castonguay, L. Fuel 1995, 74, 791799.

in a manner less than that of PD coke. For PD coke, CO2 gasification starts at higher temperatures than H2O gasification. At higher concentration of the reactive gas, the gasification exhibits a saturation behavior typical for heterogonous gassolid reactions. A thermally induced deactivation at above 1310 K is observed for H2O gasification of PD coke. The effect of particle size is shown in Figure 6a,b for Flexicoke and PD coke, respectively, obtained with petcoke particles subjected to 60% H2O-Ar mixtures and a linear heating rate of 10 K min-1 in the 500-1520 K range. For Flexicoke, the reaction rate is not considerably affected by varying the mean particle size. In contrast, for PD coke, the reaction rate increases with decreasing mean particle size. Pyrolysis starts at somewhat lower temperatures (about 600 K)

1256 Energy & Fuels, Vol. 20, No. 3, 2006

Trommer and Steinfeld

Figure 7. Experimentally measured (data points) and model simulated (curves) conversion rates obtained from pyrolysis of (a) Flexicoke and (b) PD coke.
Table 1. Kinetic Parameters for the Pyrolysis of Petcoke Flexicoke k0 r1 r2 4.152 1.881 100 10-2 EA 5.157 1.210 105 104 ci 0.0136 0.0207 k0 1.905 7.528 10-3 10-1 PD coke EA 5.087 4.626 104 104 ci 0.0794 0.0393

and has a higher reaction extent for powder size fractions obtained by jet and ball mill. For particles bigger than 200250 m, the reaction does not go to completion before reaching 1310 K, and its rate undergoes attenuation by thermal deactivation. A residuum of 3.6 and 1.1% weight was obtained for Flexicoke and PD coke, respectively. Determination of the Rate Constants Pyrolysis. Figure 7 shows the measured (data points) and modeled (curves) conversion rate dXP/dt obtained for the pyrolysis of (a) Flexicoke and (b) PD coke. Smoothing algorithm based on least-squares quadratic polynomial fitting was applied to the measured TG curve. The kinetic parameters k0, EA, and the maximum conversion ci for two pseudo-components, defined by eqs 2-4, were determined by the least-squares method. The root-mean-square error (RMS) between the experimentally measured and the theoretically modeled reaction rate dXP/dT was minimized with an algorithm of MATLAB.28 MATLABs command fminsearch finds the minimum of a scalar function of several variables with the Nelder-Mead nonlinear minimization algorithm, generally referred to as unconstrained nonlinear optimization.28 The kinetic parameters are listed in Table 1 for Flexicoke and PD coke. The weight losses of Flexicoke for the first and the second reactions were 1.36 and 2.07%, respectively, giving a total of 3.43%. The weight losses of PD coke for the first and the second reactions were 7.94 and 3.93%, respectively, giving a total mass loss of 11.87%. The RMS of the absolute error between the measured and the modeled petcoke pyrolysis was 8.40 10-7 for Flexicoke and 2.24 10-6 for PD coke. Gasification. The Arrhenius parameters of the rate constants defined by eq 20 (derived by applying the oxygen-transfer mechanism) and eq 50 (derived by applying extended mechanism) were determined numerically using the same algorithm by MATLAB28 for minimizing the RMS error between measured and modeled data. The numerical procedure was carried out simultaneously for all temperatures and all combinations of partial pressures (H2O-CO2-Ar and CO2Ar mixtures) applied in the experiments. Table 2 lists the best(28) MATLAB R14; The MathWorks, Inc.: Natick, MA, 2005.

Table 2. Arrhenius Parameters of the Rate Constants Derived by Applying the Oxygen-Exchange Mechanism (Eq 20) for the H2O-CO2 Gasification of PD Cokea k1 (s-1 Pa-1) k0 (ki) EA (J/mol)
a

k2 (s-1 Pa-1) 1.602 1.618 105 10-1

k3 8.021 107 2.416 105

2.229 2.651 105 102

RMS ) 2.66 10-5.

fit kinetic parameters of the rate constants derived for the oxygen-exchange mechanism of the steam gasification of PD coke. The results are in agreement with data presented in gasification studies with pure H2O or CO2.10 Further, the gasification with H2O-CO2 mixtures can be described using the rate constants obtained from experiments with pure gases, as claimed previously.14 As far as the steam gasification of Flexicoke is concerned, the oxygen-exchange mechanism is not capable of describing with reasonable accuracy the reaction rate. As will be shown in the following figures, the addition of CO2 to a reacting gas containing more than 60% H2O results in a decrease of the reaction rate, which is not predicted by eq 20. Presumably, mass transfer is limited by slow diffusion of the reactive gas into the interior of the particles through their pores.14 Table 3 lists the best-fit kinetic parameters derived for the extended mechanism of the H2O-CO2 gasification of Flexicoke and PD coke. Figure 8 shows the comparison of the experimentally measured (data points) and the theoretically modeled (curves) rate of the reaction rC as a function of temperature for various H2O-CO2 mixtures. The parameter is the composition of the gaseous reactant, expressed in %H2O-%CO2 at 1 bar total pressure. Note that, for the measured data, the weight loss by pyrolysis in Ar has been subtracted from the weight loss by gasification in H2O-CO2 mixtures. It is shown that the extended model is closely obeyed over the 900-1310 K temperature range considered and over the complete range of H2O/CO2 concentration. The RMS of the absolute error between experimental and modeled rate of reaction is 3.02 10-5 for Flexicoke and 2.75 10-5 for PD coke. Figure 8 further shows that the rate of H2O gasification is comparable for both types of cokes, whereas the rate of CO2 gasification is considerably faster for Flexicoke than that for PD coke.

Combined Pyrolysis and Steam Gasification of Petcoke

Energy & Fuels, Vol. 20, No. 3, 2006 1257

Table 3. Arrhenius Parameters of the Rate Constants Derived by Applying the Extended Mechanism (Eq 50) for the H2O-CO2 Gasification of Flexicoke and PD Coke K1 (Pa-1 s-1) Flexicoke RMS ) 3.02 10-5 PD coke RMS ) 2.75 10-5 k0 (Ki) EA (J/mol) k0 (Ki) EA (J/mol) 1.445 2.214 105 2.251 101 2.408 105 101 K2 (Pa-1 s-1) 6.472 2.508 105 9.771 10-1 1.812 105 102 K4 (Pa-1) 1.155 6.137 103 4.683 10-8 7.148 102 10-5 K5 (Pa-1) 5.592 10-4 5.801 104 6.640 10-6 7.239 103

Figure 9 is a cross-plot of the previous figure at 1273 K and shows a contour of reaction rate rC, in s-1, as a function of the H2O and CO2 partial pressures, for (a) Flexicoke and (b) PD coke. Despite the differences in the values of the kinetic parameters K9 and K10 associated with the H2O terms of eq 50, the reaction rate with 100% H2O is similar for both types of coke, about 2.7 10-3 s-1. In contrast, the reaction rate with 100% CO2 is 7 10-4 s-1 for Flexicoke, more than twice the

value observed for PD coke. Interestingly, adding CO2 to a H2O-CO2 reactive gas at higher steam partial pressures results in a decrease of the reaction rate for Flexicoke and an increase of the reaction rate for PD coke. Figure 10 shows the reaction rate (log scale) as a function of the mean particle size for (a) Flexicoke and (b) PD coke and for various reaction temperatures. The data points correspond to the experimentally measured values. The curves correspond

Figure 8. Experimentally measured (data points) and the theoretically modeled (curves) reaction rate rC as a function of temperature for the gasification of (a) Flexicoke and (b) PD coke. The parameter is the different composition of the reacting H2O-CO2 mixture, expressed in % H2O - % CO2 at 1 bar.

Figure 9. Reaction rate rC, in s-1, as a function of the H2O and CO2 partial pressures, at 1273 K for (a) Flexicoke and (b) PD coke. Total pressure is 1 bar.

Figure 10. Experimentally measured (data points) and the theoretically modeled (curves) reaction rate rC as a function of the mean particle size for the gasification of (a) Flexicoke and (b) PD coke. The parameter is the reactor temperature.

1258 Energy & Fuels, Vol. 20, No. 3, 2006

Trommer and Steinfeld


Table A1. Approximate Main Elemental Chemical Composition in wt % (Ultimate Analysis), Low Heating Value (LHV), and Elemental Molar Ratios of H/C and O/C for the Two Types of Petcoke Flexicoke PD coke 88.21 4.14 1.46 2.28 4.16 35876 0.5581 0.0124

to the least-squares fit to the power proportionality of eq 52, with n ) 0.234 (RMS ) 0.0445) for Flexicoke, and with n ) 0.504 (RMS ) 0.0227) for PD coke. Summary and Conclusions We have derived kinetic rate laws for the combined pyrolysis gasification of petcoke and determined experimentally the Arrhenius parameters by dynamic thermogravimetric analyses of the Ar pyrolysis and H2O/CO2 gasification of two types of Venezuelan petcokes (Flexicoke and PD coke). The pyrolysis is well described by applying a two-pseudo-component decomposition mechanism. The gasification kinetics is modeled by applying two mechanisms: an oxygen-transfer mechanism based on reversible oxygen-transfer surface reactions and irreversible carbon gasification, and an extended mechanism that considers, in addition, OH/H sorption and hydrogen transfer. The former, simpler, mechanism fails to describe the Flexicoke gasification, presumably because of intraparticular mass transfer limitations. The latter mechanism is able to simulate, with good experimental agreement, the H2O/CO2 gasification of both types of petcoke. For 250-355-m particles, the rate of gasification with pure H2O, about 2.7 10-3 s-1, is comparable for both petcokes, while the rate of gasification with pure CO2, about 7 10-4 s-1, is twice as fast for Flexicoke than that for PD coke. Main evolved gases during pyrolysis include C2H6, C2H4, CH4, CO2, CO, and H2 for PD coke, and CO, CO2, and H2 for Flexicoke. In addition to the pyrolytic gas products, the main gasification products are H2, CO, CO2, and H2S. The dependence of the reaction rate on the active site availability was empirically found to be well described by a power proportionality to the mean particle size:

C H O N S LHV (kJ/kg) H/C (mol/mol) O/C (mol/mol)

92.70 0.67 0.92 0.90 1.98 32983 0.0859 0.0074

Nomenclature
Ri ) conversion of pseudocomponent (0 e Ri e 1) a ) active surface, m2/g ) linear heating rate (ramp) of the thermobalance, K/s ci ) maximum conversion of a pseudo component with respect to the overall sample mass C* ) carbon site on the coke surface EA ) activation energy, J/mol ki ) elementary rate constants of the oxygen exchange mechanism, units according to rate law ki ) elementary rate constants of the extended mechanism, units according to rate law Ki, ai ) lumped rate constants, units according rate law k0 ) pre-exponential factor, same units as rate constant m ) mass of coke sample, g ni ) mole number, mol n ) constant Ni ) species-dependent numerator of a rate expression pi ) partial pressure, Pa R ) ideal gas constant, J/(mol K) ri ) rate of reaction, s-1 i ) fractional surface coverage (0 e i e 1) T ) absolute temperature, K t ) time, s X ) overall coke conversion (0 e X e 1) XP ) coke conversion due to pyrolysis (0 e XP e ci) XG ) coke conversion due to H2O-CO2 gasification (0 e XG e 1) x ) H/C elemental molar ratio y ) O/C elemental molar ratio Subscripts G ) gasification i ) chemical species or number of reaction P ) pyrolysis 0 ) initial AbbreViations BET ) Brunauer-Emmet-Teller DTG ) differential thermogravimetry GC ) gas chromatograph LHV ) low heating value PD ) Petrozuata Delayed RMS ) root-mean-square (error) TG ) thermogravimeter
EF050290A

rC

()
1 dP

where n is determined experimentally for each type of petcoke.


Acknowledgment. This study was conducted within the framework of a joint PDVSA-CIEMAT-ETH project aimed at developing the chemical reactor technology for the solar gasification of petcoke. We thank A. Morales, D. Rodriguez, E. Lima, and J. C. de Jesus from PDVSA, M. Romero and A. Vidal from CIEMAT, and A. ZGraggen and P. Haueter from ETH for helpful discussions.

Appendix Table A1 shows the approximate main elemental chemical composition in wt % (ultimate analysis), LHV, and elemental molar ratios of H/C and O/C for the two types of petcoke.

Вам также может понравиться