Anda di halaman 1dari 7

702

Energy & Fuels 1997, 11, 702-708

Effects of Inertinite Content in Coal on Char Structure and Combustion


Angeles G. Borrego, Diego Alvarez,* and Rosa Menendez
Instituto Nacional del Carbon, CSIC, Apartado 73, 33080 Oviedo, Spain Received August 29, 1996X

This paper reports data on the combustion characteristics of seven high-volatile bituminous coals (VRr 0.65%) with increasing inertinite contents (10-68%). Structural parameters of pyrolysis chars are compared with the maceral compositions and distributions of maceral reflectances in the parent coals and with their combustion efficiencies. The microscopic study of pyrolysis chars obtained at 1000 C reveals that inertinite-rich coal chars are much denser than those of inertinite-lean coals. The reactivity of these chars at low temperatures (500 C) decreases with the increase of inertinite content of the parent coals, this being attributed to the highly cross-linked structure of the inertinites. However, at higher temperatures (1100 C) the burnout levels of the inertinite-richest coals were clearly higher than those of medium-to-low inertinite coals. This cannot be explained from the assumption of a diffusional control of the combustion process. Thus, a chemical control of the process is suggested where the mechanism governing the overall reaction rate at 1100 C would be the active site concentration in the molecular structure of pyrolysis chars, which, in turn, would depend on the aromaticity and degree of molecular ordering in the parent maceral structure.

Introduction The maceral composition of coals is one of the features that, together with rank, account for the wide range of variability in the physicochemical characteristics of coals. There is general agreement about the origin and evolution of the different maceral groups and also about the variations in their chemical compositions brought about by their distinct genesis. Nevertheless, despite the clear differences between these materials, the behavior of each maceral during any of the coal conversion processes is still a matter of discussion. Specifically, there is no clear understanding about the transformations undergone by the different macerals when submitted to the environmental conditions typical of pulverized fuel (p.f.) combustion, for instance, which of them are the most and the least likely to burn. Liptinite is the only maceral group that has a clear role during combustion. As a result of its high volatile matter and hydrogen contents and its low aromaticity, the presence of moderate amounts of liptinite is regarded as a positive feature that favors the onset of a stable flame.1 On the other hand, there is no clear evidence about how the amount and type of vitrinite and inertinite present in a coal affect its combustion performance. The fact that vitrinites show very homogeneous properties at any given rank and, moreover, that these properties vary with rank in a quite regular and predictable way2-5 points to the inertinites as being responsible for any anomalous behavior of coals during combustion. This
* To whom correspondence should be addressed. X Abstract published in Advance ACS Abstracts, March 1, 1997. (1) Smith, K. L.; Smoot, L. D.; Fletcher, T. H. In Fundamentals of coal combustion for clean and efficient use; Smoot, L. D., Ed.; Elsevier: Amsterdam, 1993; pp 131-298. (2) Field, M. A. Combust. Flame 1970, 14, 237-248. (3) Thomas, C. G.; Holcombe, M.; Shibaoka, B. C.; Young, L. F.; Brunckhorst, L. F.; Gawronski, E. Proc. Int. Conf. Coal Sci. 1989, 257260. (4) Bend S. L.; Edwards I. A. S.; Mash H. Fuel 1992, 72, 493-501.

maceral group combines very heterogeneous materials at any coal rank, and these do not show regular variations in their properties with coalification.6 Different limits, most of them inherited from coke-making, have been used over recent years to assess the amount of reactive and inert inertinite. Most criteria incorporate inertinite reflectance and coal rank,7-14 whereas others are based on maceral composition.15-17 The term inert, initially coined to group those macerals that remained unfused during carbonization, could be misleading when applied to combustion, since the interest here is placed on the ability of a coal to burn efficiently rather than to develop a highly porous structure. Moreover, combustion proceeds under substantially different operating conditions from those found in carbonization. Thus, the reaction time is much shorter (2-3 s), much higher temperatures are achieved (approximately 1500 C), and the atmosphere is oxygenrich. The common belief that the diffusion of oxygen
(5) Oka, N.; Murayama, T.; Matsuoka, H.; Yamada, S.; Yamada, T.; Shinozaki, S.; Shibaoka, M.; Thomas C. G. Fuel Process. Technol. 1987, 15, 213-224. (6) Stach, E. Brennst. Chem. 1952, 33, 368. (7) Falcon, R. M. S.; Snyman, C. P. An Introductin to coal petrography: Atlas of petrographic constituents in the bituminous coals of Southern Africa; The Geological Society of South Africa: South Africa, 1986; Review 2. (8) Skorupska, N. M.; Sanyal, A.; Hesselman, G. J.; Crelling, J. C.; Edwards, I. A. S.; Marsh, H. Proc. Int. Conf. Coal Sci. 1987, 827-831. (9) Jones, R.; McCourt, C.; Morley, C.; King, K. Fuel 1985, 64, 14601467. (10) Bailey, J.; Tate, A.; Diessel, C. F. K.; Wall, T. Fuel 1990, 4, 225-239. (11) Diessel, C. F. K. Fuel 1983, 62, 883-892. (12) Thomas, C. G.; Shibaoka, M.; Gawronski, E.; Gosnell, M. E.; Phong-anant, D. Fuel 1993, 72, 913-919. (13) Taylor, G. H.; Mackowsky, M. Th.; Alpern, B. Fuel 1967, 46, 431. (14) Shapiro, N.; Gray, R. Proc. Ill. Min. Inst. 1960, 68, 83. (15) Gray, R.; Goscinsky, J.; Schoenberger, R. Proc. J. Conf. Iron Steel Soc., Soc. Min. Eng. AIME 1978, 3. (16) Nandi, B. N.; Brown, T. D.; Lee, G. K. Fuel 1977, 56, 125-130. (17) Furimsky, E.; Palmer, A. D.; Kalkreuth, W. D.; Cameron, A. R.; Kovacik, G. Fuel Process. Technol. 1990, 25, 135-151.

S0887-0624(96)00130-2 CCC: $14.00

1997 American Chemical Society

Effects of Inertinite Content

Energy & Fuels, Vol. 11, No. 3, 1997 703


Table 1. Maceral (vol %) and Random Vitrinite Reflectance (Rr) Analyses of Coals GG inertinite fusinite semifusinite macrinite micrinite sclerotinite inertodetrinite vitrinite telinite collotelinite collodetrinite corpogelinite liptinite sporinite resinite cutinite Rr (%) 10.4 4.9 3.5 0.2 0.6 0.4 0.8 84.2 2.9 53.7 27.4 0.2 5.4 5.0 0.4 0.0 0.68 GE 11.1 3.7 2.6 1.4 1.8 0.0 1.6 74.3 1.4 42.7 29.8 0.4 14.6 11.8 0.0 2.8 0.58 CE 17.8 6.0 8.2 0.8 1.0 0.6 1.2 78.3 0.8 17.4 59.7 0.4 3.9 2.5 0.8 0.6 0.62 LE 29.1 18.7 8.8 0.2 0.0 0.6 0.8 65.9 1.2 41.6 22.9 0.2 5.0 4.2 0.2 0.6 0.71 DR 40.7 17.3 14.6 2.7 1.0 1.5 3.6 50.1 0.4 13.2 36.1 0.4 9.2 8.4 0.4 0.4 0.59 VD 67.6 4.8 31.7 17.8 1.5 0.0 11.8 27.4 0.0 10.0 17.0 0.4 5.0 4.4 0.6 0.0 0.73 KR 68.4 6.8 34.4 16.9 0.6 0.0 9.7 25.4 0.2 9.7 15.3 0.2 6.2 5.0 0.6 0.6 0.71

into the porous structure of coal chars is the mechanism that controls the overall combustion process led many research groups to identify the inert materials in the coking process with the least likely to burn fraction of coal, although there is no total agreement among the authors about the importance of coal plasticity during p.f. combustion. Indeed, higher combustion reactivities of semifusinite compared to vitrinite18 and higher reactivities in unfused inertinite chars than in fused chars coming from either vitrinite or inertinite19 have been reported. Thomas et al.20 critically reviewed the problems found when studying the influence of maceral composition on p.f. combustion and concluded that the final conversion achieved after coal combustion is related to the presence of a certain least-likely-to-burn fraction, not identifiable with any specific maceral, which generates dense chars in the pyrolysis stage, and these would be inefficiently gasified because of diffusional hindrances. In this paper, the influence of coal inertinite content on the structure of pyrolysis chars will be studied. The variation of combustion efficiency at 1100 C with both the inertinite content of the parent coals and the porosity of the corresponding pyrolysis chars will also be investigated and discussed using a series of highvolatile bituminous coals with similar vitrinite reflectances and inertinite contents ranging from 10 to 68%. Experimental Section
Coal Preparation. Seven coals with similar rank (about 0.65% random vitrinite reflectance) and different inertinite content were used in this study. These coals were selected from different parts of the world, nominally, Gedling (GE) from England, El Cerrejon (CE) from Colombia and five more from the former Gondwana continent: Drayton (DR) and Lemington (LE) from Australia, and van Dykes (VD), Kromdraai (KR), and Groote Geluk (GG) from South Africa. Representative coal samples were ground and sieved to a size range 36-75 m in diameter. Dry instead of wet sieving was performed in order to prevent coals from being oxidized. Particle size distributions were determined by Coulter counter using a Multisizer Accucomp II, which revealed identical distributions for the size fractions from all the studied coals, with sharp peaks around 53 m and spread within the limits of 36-75 m. Coal Characterization. Ultimate, proximate, and petrographic analyses were carried out on these size fractions. Additionally, a reflected light Zeiss photomicroscope was used for random reflectance (Rr) measurements on all the macerals selected by a point-counting procedure (500 points), the experimental conditions being the same as those used for standard vitrinite reflectance analyses.21 Char Preparation. Five grams of each coal were pyrolyzed at 1000 C under nitrogen in an entrained flow reactor (EFR), with an estimated heating rate of 104-105 C min-1. The gas flow rate was maintained at 37.5 L min-1, which ensured the existence of laminar flow conditions and the required residence time of 1 s. Coals were fed at a rate of 3 g h-1. Char Characterization. Morphological characterization of the individual char structures was performed by reflected
(18) Tsai, C.-Y.; Scaroni, A. W. Meeting of Society of Mining Engineers, AIME, Denver, CO, 1984. (19) Thomas, C. G.; Gosnell, M. E.; Gawronski, E.; Nicholls, P. M. CSIRO Internal Report CET/IR 487; CSIRO: New South Wales, Australia, 1996. (20) Thomas, C. G.; Gosnell, M. E.; Gawronski, E.; Phong-anant, D.; Shibaoka, M. Org. Geochem. 1993, 20, 779-788. (21) International Committee for Coal Petrology. International handbook of coal petrology; Centre National de la Recherche Scientifique: Paris, 1971, 1975.

light microscopy using a Zeiss microscope attached to an image analysis system (Vidas 2.1, Kontron). The selection of particles was performed by automatic point counting on polished surfaces, individual particles being chosen when the crosswire fell on their carbonaceous material. Porosities, surface-tovolume ratios, equivalent (circle) diameters, and shape factors were determined on digitized binary images of these sections.22 The optical texture of the carbonaceous matter was also examined under the microscope using crossed polars and a 1 retarding plate that generates different interference colors depending on the local orientation of the polyaromatic molecules that form the carbonaceous matter. Reactivity and Burnout Experiments. Char combustion profiles were obtained in a thermobalance SETARAM TAG24. A quantity of 10 mg of each char sample was heated to 700 C at 50 C min-1 under nitrogen. Once the weight of the sample was stabilized, the temperature was reduced to 500 C and the gas was changed from nitrogen to air. The char samples were maintained under these conditions for 2 h, sufficient to achieve complete combustion. Mass losses were monitored as a function of residence time. Burnout experiments were performed in a drop-tube furnace operating at 1100 C under air, the oxygen being 50% in excess of the stoichiometric amount required and the residence time around 0.3 s.

Results Chemical and Petrographic Characterization of Coals. The random vitrinite reflectance values of the coals used in this study are in Table 1. All these values lie within a narrow interval (Rr ) 0.58-0.73%); consequently, no rank effects on either char structure or combustion reactivity are expected, and any variations should be due to differences in the maceral composition of the coals. The main petrographic features of this set of coals can be summarized as follows. Liptinite contents are generally low, the highest values corresponding to DR (9.2%) and GE (14.6%). Inertinite contents increase from 10.4% (GG) to 68% (VD and KR) with a parallel decrease of vitrinite contents from 84% to 25% (Table 1). Semifusinite and fusinite account for most of the inertinite macerals except in VD and KR, where semifusinite and macrinite make up the bulk of the inertinite macerals. Fusinite and semifusinite contents are similar in GG, GE, and DR, whereas a higher proportion of fusinite compared to semifusinite was found in LE. CE exhibits a detrital aspect reflected in
(22) Alvarez, D.; Borrego, A. G.; Menendez, R. Fuel, in press.

704 Energy & Fuels, Vol. 11, No. 3, 1997


Table 2. Proximate and Elemental Analyses of Coalsa ash (% db) GG GE CE LE DR VD KR
a

Borrego et al.

V.M (% daf) 39.8 40.5 39.4 37.4 39.4 29.1 30.0

F.R. 1.51 1.47 1.54 1.67 1.54 2.44 2.33

C.V. (MJ kg-1 daf) 32.93 32.69 33.04 33.54 33.30 32.86 32.55

C (% daf) 81.3 81.3 80.4 83.2 82.7 77.0 75.8

H (% daf) 5.3 5.3 5.4 5.3 5.3 3.9 3.9

N (% daf) 1.8 1.7 1.6 1.8 1.7 1.7 1.7

S (% daf) 0.7 1.0 0.5 0.4 0.6 0.3 0.3

Odif (% daf) 10.9 10.7 12.1 9.3 9.7 17.1 18.3

O/Cb 0.10 0.10 0.11 0.08 0.09 0.17 0.18

H/Cb 0.78 0.78 0.81 0.76 0.77 0.61 0.62

11.0 3.0 6.8 11.5 13.4 14.1 12.8

db ) dry basis; daf ) dry ash free basis. b Atomic ratio.

Figure 1. Cumulative coal reflectograms.

its high proportion of collodetrinite.23 Part of the collotelinite in GG (13%) shows early oxidation cracks (pseudovitrinite). Only slight variations in chemical composition with increasing inertinite content of coals were observed (Table 2). In fact, a significantly higher fuel ratio (F.R.) and lower volatile matter, C and H contents, and H/C and O/C atomic ratios were observed only in KR and VD, the highest inertinite coals. This indicates the low sensitivity of chemical parameters to detect variations in the maceral composition of coals. Combined maceral and reflectance analyses were also carried out on these coals (500 points) with the aim of getting a more quantitative description of the different macerals, particularly the inertinites, whose reflectances are a good estimate of their degree of alteration. Thus, the higher the reflectance of a given inertinite, the more intense the alteration it has undergone. These data are shown as cumulative histograms (Figure 1) in order to allow for comparisons between coals to be made. All the curves show three distinguishable sections with different slopes that correspond, in increasing order of reflectance, to liptinite, vitrinite, and inertinite with low degrees of overlapping among the reflectances of the different maceral groups. GG, GE, and CE, which are the lowest inertinite coals (<18%), show similar profiles with steep slopes corresponding to vitrinites, the inertinites showing a homogeneous distribution of reflectances over the range 1.0-3.0%. LE (29% inertinite) also shows a uniform distribution of inertinite reflectances and a high fusinite content. The profile of DR (40% inertinite) displays some remarkable features due to the high proportion of low-reflecting inertinite (Rr < 1.5%). The curves from LE and DR become nearly confluent at a reflectance value of about 1.5%, both coals having very similar fusinite contents, though the latter has higher amounts of semifusinite (Table 1). Finally, KR and VD show very similar curves and maceral compositions. The distribution of inertinite reflectances
(23) ICCP Vitrinite Classification 1994. 46th ICCP Meeting, Oviedo, Spain, 1994.

in these coals shows modes at about 1.7% with more than 50% of the total components of both coals having reflectances higher than 1.5%. Conversely, in the other five coals (GG, GE, CE, LE, DR), more than 80% of the total macerals display reflectances below 1.5%. In general, the inertinite reflectances mirror what could be expected from the corresponding maceral compositions. Thus, for those coals in which semifusinite and fusinite make up the bulk of their inertinite contents, the distributions of inertinite reflectances do not show any modal values, whereas those in which semifusinite and macrinite account for the majority of the inertinites (VD and KR) show a Gaussian-like distribution of inertinite reflectances. The fact that only small differences have been observed in the chemical data of these coals (Table 2) can thus be better understood from their distributions of inertinite reflectances. Over the past 30 years, many different criteria have been suggested to determine the proportions of nonreactive inertinite, based on experimental data. Most of them rely on the experience accumulated in the field of coke-making. Some of these methods are based on inertinite reflectance through the choice of a reflectance value above which inertinites would not react. Others establish threshold values dependent on the vitrinite reflectance of the coal and propose margins of tolerance over that particular value. Some proposed margins are 0.15-0.20% (Taylor et al.13), 0.3% (Falcon and Snyman7), and 0.5% (Skorupska et al.8). Several authors stated that inertinite fusibility was rank dependent and agreed on a threshold value of about 1.4-1.5% for highvolatile bituminous coals.9-12,24 Bailey et al.10 obtained best-fit formulas, in which vitrinite reflectance was also the independent variable, to evaluate the percentages of semireactive and nonreactive inertinite. Other methods use maceral composition to discriminate between reactive and nonreactive inertinite.15-17 Considering that semifusinite was generated by partial alteration of vegetal tissues and on the basis of their cumulated experience in coke-making, first Gray et al.15 and then Furimsky et al.17 stated that only one-third of the semifusinites react while the rest of the inertinites remain unchanged. On a similar basis, Nandi et al.16 considered that macrinite, originating from humic gels and having undergone different degrees of alteration, can be transformed during pyrolysis whereas fusinite, semifusinite, and micrinite cannot. Finally, Kruszewska25 proposed the mean value of 500 reflectance measurements on all the macerals as a threshold value; the reactivity of inertinite was made here dependent on the maceral composition of the coal. Table 3 shows the proportions of inert inertinite calculated for the coals
(24) Diessel, C. F. K.; Wolff-Fischer, E. M. Int. J. Coal Geol. 1987, 9, 87-108. (25) Kruszewska, K. Fuel 1989, 68, 753-757.

Effects of Inertinite Content


Table 3. Percentages of Inert Inertinite in Coals Calculated Using the Different Criteria Found in the Literature GG Falcon and Skorupska et al.8 Jones et al.9 Bailey et al.10 Thomas et al.12 Gray et al.15 Nandi et al.16 Snyman7 7.8 7.4 5.0 4.0 5.4 9.2 9.4 GE CE LE DR VD KR 9.8 11.6 20.0 34.2 65.4 67.2 9.0 8.4 17.6 29.6 61.6 62.0 7.4 5.6 14.0 23.2 49.6 52.4 6.1 4.4 13.2 19.2 44.0 47.0 6.7 5.4 14.6 21.4 55.2 55.0 10.2 15.1 26.2 35.8 57.0 56.0 8.1 15.8 28.1 34.4 38.0 41.8

Energy & Fuels, Vol. 11, No. 3, 1997 705

Table 4. Morphological Parameters of Pyrolysis Chars size (m2) GG GE CE LE DR VD KR 1768 1663 1579 1662 1505 1851 1519 Dmean (m) 45.2 43.5 42.5 41.6 43.4 46.4 42.1 Dmin/Dmax 0.73 0.70 0.72 0.68 0.69 0.63 0.63 porosity 0.66 0.66 0.64 0.58 0.60 0.40 0.39 Sv (m-1) 1.46 1.63 1.51 1.32 1.41 1.09 1.11

in this study following several criteria in the literature. Two outstanding features come from these results: first, the wide dispersion of results, with discrepancies in the quantification of inert inertinites of up to 50% of the total inertinite content and, second, the high proportion of inertinite considered as inert even by those methods aimed at the prediction under p.f. combustion conditions.3,9-10 Char Characterization. The mean values of the main structural parameters measured in the char particles from the different samples are shown in Table 4. The variations in particle size among different samples were studied by measurements of the equivalent diameters of the particle cross sections. Estimates of 3-D particle sizes could have been done through simple mathematical treatments of these values, but the reduced number of particles collected from each sample does not offer a good statistical basis for the onset of accurate results. Thus, the 2-D data from the image analyzer were used straightforwardly, since they have sufficient validity for a comparative study. The values of mean diameter given in Table 4 randomly spread in the range 41.6-46.4, not showing any definite relationship with the inertinite content of the parent coals, the differences observed being attributed to experimental error. The aspect ratios (Dmin/Dmax) of the chars slightly decrease with the increase of inertinite content, although this effect is only apparent in the chars from KR and VD, the highest inertinite coals (Table 4). The greatest differences in the morphology of the studied chars were found in their internal structure, i.e., porosity and surface-to-volume ratio (Sv), the parameters that give a better estimate of the accessibility of the carbonaceous matter to the reacting oxygen. The mean values of char porosity listed in Table 4 clearly show how the inertinite content of coals reduces the porosity development of chars (from 0.66 down to 0.39), especially in the highest inertinite coals KR and VD. Also, the Sv decreases with increasing inertinite content of the parent coals (from 1.46 to 1.11), and again, this effect is especially pronounced in KR and VD (Table 4). The observed variations in the structure of chars seem to be only appreciable for the highest inertinite coal chars. However, the char from LE (30% inertinite) starts to manifest a tendency toward a reduction in sphericity,

porosity, and SV compared with the lowest inertinite coals, although these tendencies are somewhat neglected by the higher values found for DR (Table 4). This is in accordance with the fact that most of the inertinite in DR shows low reflectances and thus is supposed to display relatively good plastic properties. In general, the structure of chars is expected to change in a quite regular fashion toward more elongated particles with lower porosities and accessibilities (SV ) with increasing inertinite content of coals if the degree of alteration of these inertinites is at least moderate to high. Another striking feature of the data shown in Table 4 is the apparent contradiction between the observed decrease in char porosity and the fact that the mean particle size hardly changes with increasing inertinite content. Bearing in mind that the particle size distributions of all the parent coals were very similar, an increase in the porosity of the particles should be accompanied by a parallel increase in their size, since the amounts of solid residue remaining after pyrolysis are not expected to vary by more than about 15% (as calculated by the values of fixed carbon) from the most to the least volatile coal. This anomalous result could be explained as the consequence of differences in the bulk densities of the carbonaceous matter from vitrinites and inertinites. The fact that vitrinites pass through a more or less plastic stage during pyrolysis allows the densification of these materials, while at least part of the inertinite macerals will not experience any alteration, and so the bulk densities of the corresponding chars could be lower than the vitrinite-derived carbonaceous matter. This would lead to the formation of a microscopically denser carbonaceous matter (less porous particles) in the highest inertinite coals, though this material would be less dense than that of vitrinite chars at a submicroscopic (molecular) level. The optical texture of the carbonaceous matter in these chars was quite difficult to assess because of the fact that it is precisely at these rank values (around 0.6% vitrinite reflectance) where the anisotropy of chars becomes apparent under the microscope, the size of the oriented domains still being too small for an accurate identification. The unequivocally vitrinite-related chars of DR, GE, and GG show a low degree of anisotropy, whereas CE char exhibits the highest development of anisotropy . LE chars typically show a very fine-grained mosaic. Finally, some highly anisotropic regions were found in KR and VD, although the fact that they normally appear associated with unreacted inertinite fields suggests that they could be derived from lowreflecting inertinites (commonly found in close vicinity to fusinites in coals) able to develop anisotropy. In general, the optical texture of inertinite-derived chars shows all the intermediate stages between solid, unfused particles and porous, highly anisotropic structures. No relationships were found between the ability of inertinites to develop anisotropic structures and the maceral composition of the parent coals. Char Reactivity. Figure 2 shows the plots of char reactivity at 50% conversion vs inertinite content of the parent coals, measured at 500 C. A broad trend of decreasing reactivity with increasing inertinite content was found, as indicated by the high reactivities measured in the vitrinite-rich coal chars GG and GE and the low values found in KR and VD, the most inertinite-

706 Energy & Fuels, Vol. 11, No. 3, 1997

Borrego et al.

Figure 2. Reactivity (500 C) of pyrolysis chars versus inertinite content of coals.

Figure 3. Variation of coal combustion efficiency (1100 C) with inertinite content.

rich coal chars. LE and DR show intermediate reactivities, in agreement with their intermediate inertinite contents. However, the reactivity of CE was too low, according to this figure, for its low inertinite content (17%). In fact, the three last chars show very similar reactivities, despite their different inertinite contents (17, 30, and 41%, respectively). Thus, although the observed trend can be broadly explained in terms of the different inertinite contents of the parent coals, other characteristics must also play an important role in char gasification at 500 C. Combustion Experiences. The variations in the combustion efficiency of coals at 1100 C with their inertinite contents are shown in Figure 3. Despite the large scattering of these results, two particularly striking aspects have been found: the highest combustion efficiencies measured were precisely those of the most inertinite-rich coals (DR, KR, and VD); coals with very similar ranks and maceral composition (GG, GE and CE) showed very different combustion performances. The former observation does not, however, offer a general guideline able to explain, on its own, the overall results shown here. Thus, LE has a rather high inertinite content (30%), but its burnout is very similar to that of GG, with only 10% inertinite, and much lower than that of DR (40% inertinite). Also, the fact that CE had burned so inefficiently cannot be attributed to any specific feature regarding its maceral composition. Discussion The variations in the pyrolysis behavior of the studied coals can be explained in terms of their different inertinite (and, accordingly, vitrinite) contents, with the most inertinite-rich coals yielding the denser chars, in agreement with most of the work published on this

subject.26-28 However, not only the amount but also the type of inertinite present in a coal influences its porosity development during pyrolysis. Thus, the porosity of LE char (0.58) is lower than that of DR (0.60), although the latter contains significantly higher amounts of inertinite, and this has to be explained on the basis of the different distributions of inertinite reflectances and/or detailed maceral compositions (Table 1 and Figure 1). The existence of large amounts of low-reflecting inertinite in DR or, alternatively, its high semifusinite content gives rise to more porous chars than expected from its high inertinite content. Combustion data at low and medium temperatures revealed the existence of only some general tendencies linked to the maceral composition of coals, which can be summarized as (i) the reactivity of coal chars at 500 C decrease with increasing inertinite content in the parent coals (Figure 2) and (ii) at higher temperatures (1100 C), inertinite-rich coals burn more efficiently than vitrinite-rich coals (Figure 3). However, the deviations from these trends are so large in some of the coals that they cannot be regarded as having a good predictive ability any longer. Anyway, these observations are themselves valid clues that can help in the understanding of the combustion process. Feasible mechanisms of combustion for both low and medium temperatures, on the basis of these general observations, are set out below. The high combustion efficiencies (1100 C) found in KR and VD, the most inertinite-rich coals and also those yielding the least porous chars, and, conversely, the low burnout achieved by CE despite its low inertinite content and its ability to generate porous char immediately discard any possible diffusional effect as the mechanism controlling the overall reaction. The variations in the intrinsic reactivity of the carbonaceous matter that forms the char samples also have to be taken into account. The reactivity of chars at 500 C could partially depend on the accessibility of their carbonaceous matter to oxygen, thus explaining the observed decrease with increasing inertinite content. But once this effect has been discarded at 1100 C, its relative contribution at 500 C must be even less important. In this specific case, the course of the reaction at 500 C, as measured by TGA, was mainly governed by the intrinsic reactivity of the carbonaceous matter and, accordingly, the variations in the intrinsic reactivity of chars at 1100 C should follow a tendency similar to that found at 500 C, i.e., decreased reactivity with increasing inertinite content. Since the burnout data at 1100 C show precisely the opposite trend, it is concluded that the chemical reaction itself is not the mechanism controlling the combustion at that temperature. Once both the diffusion of oxygen and the intrinsic chemical reactivity have been ruled out as rate-controlling steps and assuming that large variations attributable to heat-transfer effects are unlikely to exist, given the small differences among the calorific values of the coals (Table 2), only one possibility is left, which is the concentration of active sites in the surface of the char
(26) Lightman, P.; Street, P. J. Fuel 1968, 47, 7-28. (27) Street, P. J.; Weight, R. P.; Lightman, P. Fuel 1969, 48, 343365. (28) Shibaoka, M. Fuel 1985, 64, 263-269.

Effects of Inertinite Content

Energy & Fuels, Vol. 11, No. 3, 1997 707

material governing the overall combustion rate. In any carbonaceous structure built up as a three-dimensional arrangement of polyaromatic rings, it is assumed that the active carbon atoms are those located in the edges of these rings because they have free -orbitals with the adequate energy and symmetry for the attack of oxygen to take place. However, carbon atoms in inner positions in a ring lack free -orbitals and, moreover, they are shielded from the attack of oxygen by the electronic cloud over and under the ring. In fact, it is said that carbons in peripheral positions are more reactive than inner carbons by 2-3 orders of magnitude.29 If the active site concentration of chars controls the combustion process, then by use of the existing knowledge about the structure of coals, it should be possible to explain the variations in the combustion efficiency (1100 C) and reactivity (500 C) of the coals on the basis of their theoretical molecular structures. It is established that inertinite macerals come from the same plant materials as the vitrinite macerals, though the former have undergone variable degrees of alteration (oxidation) prior to deposit. For this reason, the most severely altered inertinites remained virtually unchanged throughout the subsequent burial.30 Conversely, vitrinite macerals undergo a number of physicochemical changes (coalification) due to the increase of temperature and pressure that takes place during their burial. Thus, the final molecular structure of vitrinites is the result of gradual changes in the structure of the original plant materials through an evolutive process lasting some hundred millions of years. These changes can be summarized as follows:31 with increasing coalification (i) the degree of substitution in the aromatic rings decreases, (ii) these rings are enlarged through the coalescence of smaller units, and (iii) the planes of neighbor polyaromatic rings tend to adopt an increasingly higher degree of parallel ordering. As far as the inertinites are concerned, the chemical transformations brought about by their oxidation prior to deposit result in an essentially disordered state, its structure being described as a three-dimensional network of small polyaromatic rings with variable degrees of cross-linking depending on the degree of alteration undergone but always higher than the corresponding vitrinites.32 The inertinite-rich coals exhibited the lowest char porosities and reactivities at 500 C (Figure 2) and the highest burnout levels at 1100 C (Figure 3). The latter result can be explained as the consequence of their disordered molecular structure and their accordingly high active site concentration. The low reactivities found for these coals at 500 C could be due to the higher strength of the bonds in their structure compared with those of vitrinite chars. On the other hand, for those coals in which vitrinite is the major component, the progressive coalescence of polyaromatic units as coalification proceeds must decrease the active site concentration through the conversion of edge carbon atoms to
(29) Marsh, H.; Kuo, K. In Introduction to Carbon Science; Marsh, H., Ed.; Butterworth: London, 1989; pp 107-151. (30) Stach, E.; Mackowsky, M. Th.; Teichmuller, M.; Taylor, G. H.; Chandra, D.; Teichmuller, R. Coal Petrology; Gebruder Borntraeger: Berlin, 1982; p 271. (31) Behar, F.; Vandenbroucke, M. Org. Geochem. 1987, 11, 1524. (32) Diessel, C. F. K. Coal-bearing depositional systems; SpringerVerlag, Berlin, 1992; pp 58-59.

Figure 4. Variation of coal combustion efficiency (1100 C) with coal rank.

inner ones, thus leading to a lower combustion efficiency. This effect should not be so important at lower temperatures, where the reactivity of char depends on the nature of the bonds in its structure rather than the active site concentration. Yet, it has to be explained why GG, GE, CE, and LE display such a wide range of combustion efficiencies despite their similar rank and petrographic compositions. These different behaviors cannot be systematized on the basis of their inertinite contents, as shown in Figure 3. According to the above assumptions, these differences have to be due to variations in the rank of the selected coals. Although we had assumed that all our coals had the same rank, their mean random vitrinite reflectances were not exactly equal but expanded over the interval 0.58-0.73%. The variation of coal burnout with vitrinite reflectance for the studied coals is shown in Figure 4. The combustion performance of CE coal is much lower than those of GE and GG at either side of the former in this figure. Also, the rest of the coals seem to draw a minimum of combustion efficiency for vitrinite reflectances of about 0.60-0.65%, though the high burnout levels of KR, VD, and DR were justified by their high inertinite contents. Let us consider now which characteristics of the vitrinite could give rise to the existence of a rank interval with lower reactivities at 1100 C. The molecular structure of a char is mainly conditioned by that of the parent coal and also by the heating rate applied to the coal particles during pyrolysis. The experimental conditions used in our pyrolysis tests were the same for all the coals, so the variations in the structure of these chars have to be attributed to differences in the structure of the parent coals. The number of edge carbons in a vitrinite decreases with increasing rank because of the coalescence of polyaromatic units. Also, a more ordered stacking of these polyaromatic units will take place as coalification proceeds. The mobility of vitrinitederived polyaromatic rings during pyrolysis will allow for a further increase in both their size and parallel ordering, especially if these already had a certain degree of ordering (promoted by coalification) before pyrolysis. In a disordered structure, edge carbons of a polyaromatic ring will face the electronic clouds of the neighboring units, thus being sheltered from the attack of oxygen, whereas in an ordered structure these edge carbons will be easily attacked, provided that they are located in the surface of the char particle. Therefore, the stacking of polyaromatic rings should improve the accessibility of oxygen to their edge carbons. The common figure of the molecular structure of vitrinite changing with coalification through the coa-

708 Energy & Fuels, Vol. 11, No. 3, 1997

Borrego et al.

lescence and parallel ordering of the polyaromatic moieties can explain our results in the rank interval considered here. Thus, if we assume that the disordered polyaromatic rings progressively grow in size as coalification proceeds and that, once they reach a certain mean size, they are not able to pack properly, then the density of the material would decrease, which contradicts the observed increase of density with coal rank. Instead, the weight of the sediments over the coal bed could promote the parallel ordering of these enlarged moieties, and the density of the materials would increase with coal rank, as is known to happen. In other words, it is postulated that the growth and ordering of polyaromatic rings could take place not simultaneously but consecutively. This figure could explain our burnout data as the consequence of a reduction in active site concentration due to the coalescence of polyaromatic rings and leading to a decrease in coal burnout with increasing rank. The subsequent parallel ordering would then give rise to an increased combustion efficiency promoted by the improvement in the accessibility of these active sites, as was explained above. Ongoing work on the effects of coal rank upon coal combustion efficiency will demonstrate how, under the same experimental conditions, coals with vitrinite reflectances around 0.65% burn less efficiently than others with slightly higher or lower ranks.33 This limit (Rr ) 0.65%) coincides with the beginning of hydrocarbon generation for type III kerogens and could be the threshold up to which reorganization of aromatic micelles is possible because of the progressive release of more labile bonded hydrocarbons.34
(33) Alvarez, D. Influence of char structure on coal combustion behaviour. Ph.D. Thesis, The University of Oviedo, Oviedo, Spain, 1996 (in Spanish). (34) Vanderbroucke, M.; Bordenave, M. L.; Durand, B. In Applied petroleum geochemistry; Bordenave, M. L., Ed.; Editions Technip: Paris, 1993; pp 101-121.

Conclusions Our results of char characterization and the combustion data at low (500 C) and medium temperatures (1100 C) discard the diffusional constraints and the intrinsic chemical reactivity as possible features governing the combustion process at 1100 C. The hypothesis that it could be the concentration of active sites in the char material, the key factor for the onset of low or high burnout levels, is able to explain the totality of our results. We have no experimental evidence supporting our assumptions, but most of them rely on wellestablished theories regarding the structure of coals and its evolution with rank. Furthermore, the high levels of unburnt carbon found here for CE coal are in agreement with data on an industrial scale that have been reported for this particular coal. Our proposed explanation is also valid under the conditions prevailing in an industrial boiler, since even at those high temperatures it is unlikely that char gasification could proceed through active sites other than the edge carbons of the polyaromatic moieties. Inner carbon atoms are far less reactive than those in the periphery, and the increase in combustion temperature should be much higher than merely 500 C to mask this effect. Conversely, diffusional and/or chemical constraints would disappear or, at least, be strongly reduced by such an increase of temperature. Acknowledgment. The financial support of EU through the Project 7220-ED/075 is acknowledged. J. Bailey kindly provided the burnout experiments. D.A. and A.G.B. thank the Spanish Ministry for Education and EU TMR program, respectively, for postdoctoral fellowships.
EF960130M

Anda mungkin juga menyukai