Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Coal Gasification and Its Applications
Coal Gasification and Its Applications
Coal Gasification and Its Applications
Ebook694 pages7 hours

Coal Gasification and Its Applications

Rating: 3 out of 5 stars

3/5

()

Read preview

About this ebook

Skyrocketing energy costs have spurred renewed interest in coal gasification. Currently available information on this subject needs to be updated, however, and focused on specific coals and end products. For example, carbon capture and sequestration, previously given little attention, now has a prominent role in coal conversion processes.This book approaches coal gasification and related technologies from a process engineering point of view, with topics chosen to aid the process engineer who is interested in a complete, coal-to-products system. It provides a perspective for engineers and scientists who analyze and improve components of coal conversion processes.The first topic describes the nature and availability of coal. Next, the fundamentals of gasification are described, followed by a description of gasification technologies and gas cleaning processes. The conversion of syngas to electricity, fuels and chemicals is then discussed. Finally, process economics are covered. Emphasis is given to the selection of gasification technology based on the type of coal fed to the gasifier and desired end product: E.g., lower temperature gasifiers produce substantial quantities of methane, which is undesirable in an ammonia synthesis feed. This book also reviews gasification kinetics which is informed by recent papers and process design studies by the US Department of Energy and other groups, and also largely ignored by other gasification books.• Approaches coal gasification and related technologies from a process engineering point of view, providing a perspective for engineers and scientists who analyze and improve components of coal conversion processes • Describes the fundamentals of gasification, gasification technologies, and gas cleaning processes • Emphasizes the importance of the coal types fed to the gasifier and desired end products • Covers gasification kinetics, which was largely ignored by other gasification books
  • Provides a perspective for engineers and scientists who analyze and improve components of the coal conversion processes
  • Describes the fundamentals of gasification, gasification technologies, and gas cleaning processes
  • Covers gasification kinetics, which was largely ignored by other gasification books
LanguageEnglish
Release dateDec 8, 2010
ISBN9781437778519
Coal Gasification and Its Applications

Read more from David A. Bell

Related to Coal Gasification and Its Applications

Related ebooks

Power Resources For You

View More

Related articles

Reviews for Coal Gasification and Its Applications

Rating: 3 out of 5 stars
3/5

1 rating1 review

What did you think?

Tap to rate

Review must be at least 10 words

  • Rating: 3 out of 5 stars
    3/5
    nice book and best regarding point of you and useful book

Book preview

Coal Gasification and Its Applications - David A. Bell

Table of Contents

Cover Image

Front Matter

Copyright

Introduction

Chapter 1. The Nature of Coal

Chapter 2. Non-gasification Uses of Coal

Chapter 3. Gasification Fundamentals

Chapter 4. Gasifiers

Chapter 5. Underground Coal Gasification

Chapter 6. Sulfur Recovery

Chapter 7. Hydrogen Production and Integrated Gasification Combined Cycle (IGCC)

Chapter 8. Hydrogen Adsorption and Storage

Chapter 9. Mercury Removal

Chapter 10. CO2 Sorption

Chapter 11. Ammonia and Derivatives

Chapter 12. Methanol and Derivatives

Chapter 13. Substitute Natural Gas and Fischer-Tropsch Synthesis

Index

Front Matter

Coal Gasification and its Applications

DAVID A BELL

BRIAN F TOWLER

MAOHONG FAN

Amsterdam • Boston • Heidelberg • London • New York • OxfordParis • San Diego • San Francisco • Singapore • Sydney • Tokyo

William Andrew is an imprint of Elsevier

Copyright © 2011 Elsevier Inc.. All rights reserved.

Copyright

William Andrew is an imprint of Elsevier

The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK

30 Corporate Drive, Suite 400, Burlington, MA 01803, USA

First edition 2011

Copyright © 2011 Elsevier Inc. All rights reserved.

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher

Permissions may be sought directly from Elsevier's Science & Technology Rights Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions and selecting Obtaining permission to use Elsevier material

Notice

No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made

British Library Cataloguing in Publication Data

A catalogue record for this book is available from the British Library

Library of Congress Cataloging-in-Publication Data

A catalog record for this book is available from the Library of Congress

ISBN: 978-0-8155-2049-8

For information on all William Andrew publications visit our web site at books.elsevier.com

Printed and bound in Great Britain

11 12 13 14 15 10 9 8 7 6 5 4 3 2 1

Copyright © 2011 Elsevier Inc.. All rights reserved.

Introduction

This is a book about coal gasification and its related technologies. The relationship between these technologies is shown in Figure 0.1. The gasification process begins with a viable feedstock. In this book, we focus on one of those feedstocks that must go through the gasification process, coal. The nature of coal, including its properties and availability, are described in Chapter 1. Petcoke, petroleum coke, a solid, high-carbon byproduct of petroleum refining, can also be gasified. Gasifiers designed for coal, especially high temperature, entrained flow gasifiers, are used for this application. Biomass gasification has a great deal in common with coal gasification, but biomass gasifiers are optimized for biomass feedstock.

The product of gasification is syngas, which is primarily a mixture of carbon monoxide and hydrogen. Most syngas, however, is not currently made by gasification, but rather by the steam reforming of natural gas. In this process, steam and natural gas are fed to catalyst-packed tubes, which are held inside a furnace to provide the endothermic heat of reaction. Figure 0.1 also shows other gases, which can be blended with syngas for further processing. One such gas under consideration is hydrogen, which can be produced by electrolyzing water using off-peak power from a nuclear power plant. In a few cases, carbon dioxide from an external source may supplement the carbon monoxide in syngas.

Just as coal is not the only feedstock for gasification, gasification is not the only use of coal. Most coal is burned to produce electric power. Chapter 2 describes a few of the non-gasification uses of coal.

Gasification is described in Chapter 3, Chapter 4 and Chapter 5. Chapter 3 describes gasification as a chemical reaction system. Although this chapter may look complex, our knowledge of the chemistry of gasification is far from complete. Chapter 4 covers several gasifier designs. These designs were selected because they are now in commercial use or development, or because they illustrate interesting concepts. One gasification approach is sufficiently different that it deserves its own chapter, underground coal gasification, covered in Chapter 5. Instead of mining coal and transporting it to a gasifier, the coal is left in place underground, and the reactant gases are brought to the coal. Deeply buried coal seams, which are uneconomic to mine, may be exploited by underground coal gasification.

Syngas leaving the gasifier contains numerous impurities. The inorganic fraction of the feedstock leaves as solid ash or molten slag. Ash or slag removal is usually an integral part of the gasifier design. If the gasification occurs at relatively low temperatures, then tar will be produced. Tar removal is also an integral part of gasifier design. Higher-temperature gasifiers do not produce significant tar. The syngas also contains sulfur in the form of H 2S, with lesser quantities of COS. Sulfur must be removed from syngas either to prevent emission of SO 2 when syngas is burned, or to prevent catalyst poisoning in downstream reactors. Sulfur removal is described in Chapter 6.

Carbon dioxide removal can occur either as a part of impurity removal, or after water gas shift, as shown in Figure 0.1. The traditional carbon dioxide removal techniques are closely related to sulfur removal, and are described in Chapter 6. The ability to remove carbon from syngas and sequester it in a geological formation is one of the major attractions of coal gasification. This allows coal to be used while minimizing greenhouse gas emissions. A major objection to this approach is that carbon capture and sequestration are expensive. This prompted a great deal of research into new carbon dioxide separation technologies, and which is described in Chapter 10.

Syngas contains a number of minor impurities, and one of the more significant is mercury, a neurotoxin. Removal of mercury is discussed in Chapter 9.

For some applications, a nearly pure hydrogen stream is desired. In others, such as methanol synthesis, a specific ratio of carbon monoxide to hydrogen is required. In either case, the gasifier usually produces a higher ratio of carbon monoxide to hydrogen than desired. This ratio needs to be shifted towards a greater hydrogen content. The usual way to do this is through the water gas shift reaction in which carbon monoxide reacts with steam to form hydrogen and carbon dioxide, as described in Chapter 7. Hydrogen can then be burned in a turbine to generate electric power, an application known as integrated gasification combined cycle. This is a means of producing electric power from coal with minimal greenhouse gas emissions.

Hydrogen is also a potential transportation fuel. The usual approach is to produce electric power from hydrogen in a fuel cell, and then use that power in an electric motor. One of the main technical obstacles is a practical means of storing hydrogen in a vehicle. Chapter 9 explores hydrogen storage for this application.

Nearly all synthetic nitrogen chemicals start as ammonia, synthesized from hydrogen and nitrogen gas. Nitrogen fertilizers are, by far, the largest volume synthetic nitrogen chemicals. Chapter 11 describes ammonia synthesis and some of the more common nitrogen fertilizer compounds.

Methanol is a major commodity chemical made from syngas, as described in Chapter 12. Methanol is an intermediate used to make a wide range of products. One of these, dimethyl ether (DME), is especially interesting. DME can be used as a fuel or converted to hydrocarbons, including gasoline and olefins for polymer production.

Chapter 13 describes the direct conversion of syngas to hydrocarbons, including substitute natural gas (methane) and Fischer-Tropsch liquid, a synthetic crude oil. The Fischer-Tropsch liquid is then refined to meet petroleum product specifications.

Coal is an inexpensive feedstock, but gasification-based plants tend to have very high capital construction costs. In concept, one could build a single plant that would incorporate all of the elements shown in Figure 0.1, but such a complex plant would be extraordinarily expensive to build. Instead, gasification-based plants have a more limited set of features dictated by economics and the regulatory environment.

There are two major trends that prompt current interest in coal gasification. The first is the widely held belief that conventional petroleum supplies are declining, while demand for transportation fuels continues to rise. This has led to heightened interest in alternative energy supplies, including coal. The second major trend is concern about global warming. Gasification offers a relatively cost-effective means of using coal while minimizing greenhouse gas emissions.

Chapter 1. The Nature of Coal

Contents

The Geologic Origin of Coal 1

Coal Analysis and Classification 2

Coal Rank 4

Ash Thermal Properties 5

Coal as a Porous Material 9

Spontaneous Combustion 10

Reserves, Resources, and Production 11

References 15

The Geologic Origin of Coal

Coal is fossilized peat. A peat bog is a marsh with lush vegetation. Plant matter dies and falls into the water, where partial decomposition occurs. Aerobic bacteria deplete the water of oxygen, and bacterial metabolic products inhibit further decomposition by anaerobic bacteria. Plant matter accumulates on the marsh bottom faster than it decomposes, and, over a period of many years, a layer of peat forms. The peat that became today’s coal was laid down millions of years ago.

Buried peat is converted to coal when high pressure and elevated temperature is applied to the buried layer. This process is known as coalification. The physical and chemical structure of the coal changes over time. As shown in Figure 1.1, the youngest (least converted) coal is known as lignite, which can be further converted to sub-bituminous coal, bituminous coal, and finally anthracite. These coal types strongly influence the properties and use of coal, and will be discussed further.

Petrography is the visual inspection of a rock sample to determine the mineral types in the sample. When applied to coal, the different coal types are known as macerals. Table 1.1 lists coal macerals, and shows how they are derived from plant material.

Coal Analysis and Classification

Coal is used primarily as a fuel, so its most important property is its heat of combustion. Gross calorific value, also known as higher heating value (HHV), is determined by measuring the heat released when coal is burned in a constant-volume calorimeter, with an intitial oxygen pressure of 2 to 4 MPA, and when the combustion products are cooled to a final temperature between 20 and 35°C (ASTM D 5865-04). The tests mentioned in this book are primarily based on the American Society for Testing and Materials (ASTM) specifications. ¹ Coal is a variable, widely distributed and widely used material so a wide range of standard tests have been developed by a variety of individuals and organizations.

Coal is a porous medium, and these pores, especially in low rank coals, can contain substantial quantities of water even though the coal appears to be dry. The water is either adsorbed onto hydrophilic surface sites or held in pores by capillary forces. When this moist coal is burned or gasified, a substantial fraction of the combustion heat is required to vaporize water. Since the final temperature in the gross calorific value test is 20 to 35°C, most of the water is condensed, thereby recovering the heat of vaporization.

Water in the HHV test is primarily a non-combustible diluent. For example, a Wyoming Powder River Basin coal typically has an HHV of 19.8 MJ/kg (8500 Btu/lb) and a 28% moisture level. One can then calculate an HHV value for the coal if it is dried:

Eqn. 1.1

If coal is burned or gasified near atmospheric pressure, then the heat of condensation for the water may not be recovered. For example, in a coal-fired power plant, the water contained in the coal may go up the stack as steam. In other situations, the heat of condensation is recovered, but the value of this heat is relatively low because of its temperature. In these cases, a better estimate of coal heat of combustion is the net calorific value, also known as Lower Heating Value (LHV), which assumes that vaporized water remains as steam and that the heat of condensation is not recovered. Water in the coal reduces its heating value by its heat of vaporization, 2.395 MJ/kg water (1055 Btu/lb water). Again, for a typical PRB coal:

Eqn. 1.2

Proximate Analysis (ASTM D 3172-89) involves a series of tests that heat and burn coal. Moisture is measured (ASTM D 3173-03) by determining the weight loss after coal is dried at 104 to 110 °C. Volatiles are then measured (ASTM D 3175-02) by determining additional weight loss when coal is pyrolyzed at 950 °C. Ash is determined (ASTM D 3174-04) by the weight of inorganic materials remaining after coal is burned. Fixed carbon is the fraction of coal that is not moisture, volatiles, or ash. Fixed carbon, which is mostly carbon but can contain other elements represents the combustible portion of the coal char that remains after the volatiles have been removed.

Proximate analysis results are sometimes reported on a dry mineral matter-free basis. Mineral matter is calculated using the following equation:

Eqn. 1.3

Where: Mm = percent mineral matter

A = percent ash

S = percent sulfur (ASTM D 3177 or D 4239)

The 1.08 factor presumes that minerals in the coal are hydrated. This water of hydration is lost when the coal is burned. The 0.55 factor assumes that sulfur is present as pyrites, which in many areas are converted to the corresponding oxides during combustion.

Ultimate analysis (ASTM D 3176) describes coal in terms of its elemental composition. For a dried coal, weight percentages of carbon, hydrogen, nitrogen, sulfur, and ash are measured. The remainder of the coal sample is assumed to be oxygen.

Coal Rank

In the coalification process, the coal rank increases from lignite to anthracite, as shown in Figure 1.1. Coal rank is useful in the market, because it is a quick and convenient way to describe coal without a detailed analysis sheet. A more detailed description of coal rank is shown in Table 1.2 and Table 1.3.

Bituminous and sub-bitumous coals are the primary commercial coals. A relatively small amount of anthracite is available. In the USA, anthracites are produced only in north-eastern Pennsylvania. Lignites are abundant. But the economics of hauling a low-grade fuel long distances are unfavorable; so most lignite is consumed close to where it is mined.

Peat is also mined and generally used close to where it is mined. Peat may be either considered old biomass or very young coal. In nations that regulate greenhouse gas emissions, the difference between the two is more than mere semantics. Carbon dioxide emissions from biomass combustion are not considered a contributor to global warming, because these emissions are offset by carbon dioxide uptake by growing biomass. On the other hand, the same emissions from fossil fuels, are restricted. Emissions from peat combustion are a regulatory gray area.

Some coal, particularly bituminous coal, has the tendency to cake. With increasing temperature, coal particles simultaneously pyrolize and partially melt, causing the coal particles to stick to one another. Some gasification reactors, especially moving bed and fluidized bed gasifiers, are limited to processing coal that does not cake.

Ash Thermal Properties

The melting temperatures of coal ash impose temperature limits for coal gasification. Fluidized bed gasifiers and dry-bottom moving bed gasifiers, such as the Lurgi gasifier, require free-flowing ash. The maximum operating temperature for these gasifiers is the initial deformation temperature. When the temperature rises above the initial deformation temperature the ash becomes sticky. Fluidized bed gasifiers often run near the initial deformation temperature to maximize carbon conversion.

Entrained flow gasifiers and slagging moving bed gasifiers such as the BGL gasifier require a fluid slag, so they must operate at a sufficiently high temperature to completely melt the ash. Operation at significantly higher temperatures increases oxygen consumption.

Ash is a complex mixture of minerals, which will cause the coal ash to melt over a temperature range rather than at a fixed temperature. Temperatures in this range are specified by ASTM D-1857-04. A coal ash cone, 19 mm high and with an equilateral triangle base 6.4 mm on each side, is placed in an oven. Temperatures are reported for reducing or oxidizing gas environments. The initial deformation temperature (IDT) occurs when rounding of the cone tip first occurs. The softening temperature (ST) occurs when the cone has fused to produce a lump which has a height equal to its base. The hemispherical temperature (HT) occurs when the lump height is half the length of its base. The fluid temperature occurs when the fused mass has spread out in a nearly flat layer with a maximum height of 1.6 mm.

A number of researchers have attempted to correlate ash thermal properties with ash composition. The most extensive effort was by Seggiani and Pannocchia, ² who correlated the behavior of 433 ash samples, based on nine elemental concentrations. Note that mineral elemental compositions are reported as if the mineral sample were a blend of simple metal oxides. For example, the fraction of aluminum in a sample is typically reported as the equivalent weight percent of Al 2O 3. Seggiani and Pannocchia’s correlations are based on mole percents, rather than weight percents, on a normalized, SO 3-free basis.

The correlation for initial deformation temperature is given as:

Eqn. 1.4

The correlation for softening temperature is given as:

Eqn. 1.5

The correlation for hemispherical temperature is given as:

Eqn 1.6

The correlation for fluid temperature is given as:

Eqn. 1.7

The temperature of critical viscosity, T cv, is not part of the ASTM D1857 test but it is important for slagging gasifiers because it marks the transition of slag from a difficult-to-handle Bingham plastic, below T cv, to a more easily handled Newtonian fluid, above T cv. The correlation for temperature of critical viscosity is given as:

Eqn. 1.8

Seggiani and Pannocchia report standard deviations for their correlations to be 70 to 88 oC. Table 1.4 compares experimental results for four American coals from Baxter³ to the temperatures predicted by these correlations. The predicted results are very close to the experimental results for the lignite and the sub-bituminous coals. The exception is the predicted temperatures are substantially higher than the experimental values for the bituminous coals.

Inorganic additives have been added to coal gasifiers to modify ash thermal properties. For example; alkaline materials such as sodium, potassium and calcium compounds tend to lower ash melting temperatures. These can be added to an entrained flow gasifier to lower slag viscosity. Care must be taken with refractrory-lined gasifiers, because these compounds may attack the refractory. The opposite approach was taken by van Dyk and Waanders. ⁴ They sought to increase the ash fusion temperature (ISO 540 and 1195E) to allow higher temperature operation in a Lurgi gasifier. Tests with Al 2O 3, TiO 2, and SiO 2 showed that Al 2O 3 was most effective. Addition of 6 weight % Al 2O 3 boosted the ash fusion temperature of a mixture of South African coals from 1,340 °C to greater than 1,600 °C.

Ash is typically land-filled. If the landfill is unlined, then water percolating through the ash pile may affect surface and groundwater quality. Many ashes are alkaline and there is the possibility that toxic heavy metals in the ash may be leached by rainwater. Slagging gasifiers produce glassy, non-leachable slag.

Some coal ashes are pozzolanic, which means that they tend to set up like cement when mixed with water. These ashes are often used as road base. High calcium ash has an analysis that is similar to commercial cement. Pozzolanic ashes are less likely to pose leachate problems than unconsolidated ashes.

Coal as a Porous Material

Coal is a porous material. ⁵ Pores are classified as macropores (greater than 50 nm), which are measured using mercury porosimetry, mesopores (2.0 to 50 nm), measured by nitrogen adsorption at 77 K, and micropores (0.4 to 2.0 nm), measured by carbon dioxide adsorption at 298 K. Micropores are due to the voids from imperfect packing of large organic molecules. Coals typically have surface areas in the range of 100 to 400 m ²/g, which is due almost entirely to micropores. For comparison, a typical atomic diameter is 0.25 nm, so only small molecules may penetrate micropores. Olague and Smith⁶ studied gas diffusion in coal.

Spontaneous Combustion

Coal oxidizes when it is exposed to oxygen at ambient conditions. Low grade coals are especially prone to low temperature oxidation. The effect of long-term air exposure on coal quality is known as weathering. Oxidation at low temperatures is exothermic, resulting in increased temperatures that accelerate the rate of coal oxidation. This sometimes leads to spontaneous combustion of coal.

Itay et al.⁷ studied low temperature oxidation of South African coal. They found that the quantity of oxygen adsorbed was greater than the quantity of oxygen-containing product gasses (CO 2, CO, H 2O), so most of the adsorbed oxygen remains in the coal. Other investigors ⁸. and ⁹. found an increase in carboxylic acids in weathered coal. Itay et al. found that oxygen uptake with repeated oxygen exposures declines. This same effect is shown in Figure 1.2¹⁰. Small particles tend to oxidize faster than large particles, but the particle size effect is not large. This suggests that the rate of oxidation is limited by oxygen diffusion in coal micropores or by surface reaction rates.

As-mined low grade coals typically have high water content, and the water-filled pores tend to block low temperature oxidation. As shown in Equation 1.1, the heat content of these coals can be greatly increased by drying. Unfortunately, the dry coals cannot be safely stored or shipped under ambient conditions due to their tendency to spontaneously combust. Exposure of dry coals to high humidity or liquid water accelerates the rate of low-temperature oxidation, possibly because water adsorption on coal is exothermic. Processes have been developed ¹¹. and ¹². to convert high moisture, low grade coals to low moisture fuels with reduced spontaneous combustion tendencies.

Reserves, Resources, and Production

Throughout history coal has played a very small role in the world’s energy mix. Locally it was a curiosity because it was an interesting rock that could be made to burn. However, commencing in about the year 1500, it began to be mined for small scale energy use in England and Germany. When the Industrial Revolution dawned in England in the eighteenth century coal became a significant energy source that fueled the English factories that were the hallmark of the Industrial Revolution. However, in terms of total energy use biomass (particularly wood) remained the major source of energy for the world until about 1900, when coal overtook biomass as the chief world energy source. In the United States, where coal was abundant and easier to mine, it had become the chief energy source in about 1880. Throughout the first half of the twentieth century coal was the major world energy source, until it was overtaken by oil in about 1960. Even though coal production has continued to increase since then it remains in second place behind oil and just ahead of natural gas. This is illustrated in Figure 1.3.

In the near future it is likely to remain in the second position until oil production peaks and starts to decline. It is conceivable that when this happens coal will again become the most consumed energy source in the world. In terms of energy reserves the world has much more coal than any other energy source. Some might argue that coal production will be restricted because of the amount of CO 2 that it produces. But as we learn to capture and sequester the CO 2 economically, coal production will likely continue to increase.

The entire quantities of coal present, regardless of the cost or practicality of recovery, are known as resources. A 5 cm thick layer of low quality coal buried under 2000 m of overburden contributes to total resources, but it is unlikely that it will ever be mined. A more practical measure of the quantities of coal available are known as reserves, which is the subset of resources that can be mined at current prices using current technology. There are an estimated¹³ 275 billion tons of coal reserves in the United States, compared to approximately 4 trillion tons of coal resources. Table 1.5 shows estimates of worldwide reserves, recent annual production and reserve/production (R/P) ratios, which is the number of years these reserves will last at current production levels. ¹⁴

It does not mean that the world will run out of coal in the next 148 years. For example, Luppens, et al.¹⁵ studied coal reserves and resources in the Gillette Coal Field, a 5,180 km ² portion of the 57,000 km ² Powder River Basin in northeastern Wyoming and southeastern Montana. Coal is abundant throughout the Powder River Basin, but mining is restricted to coal with the lowest mining cost. Most mining in the basin occurs within the Gillette Coal Field, where thick coal seams lie close to the surface. There are multiple coal beds and the thickest of these, the Anderson, has an average 15 m thickness. The beds dip slightly from east to west, so coal is produced from 13 strip mines along a 78 km north–south line along the eastern edge of the district. These 13 mines produced over 42% of the coal produced in the USA in 2007. ¹⁴

Because of the abundance of this coal, and because of its low mining cost, the open market mine mouth price¹⁴ for this sub-bituminous coal in 2007 was $9.67/ton; compared to $47.63/ton for bituminous coal from West Virginia. As shallow coal in the Gillette Coal field is depleted, mining moves to the west, with gradually increasing overburden and gradually increasing mining cost. When the overburden becomes too thick for strip mining, underground mining may be used to further extract coal. Luppens et al. estimated the volume of coal available in the Gillette Coal Field versus price, and these data are shown in Figure 1.4. At the 2008 production rate of 464 million tons/year, the coal reserves in the Gillette Coal Field will last only 21 years if coal is priced at $10/ton. If the price of coal rises to $60/ton, then the coal will last 176 years at the 2008 production rate. Of course, the Gillette Coal Field is only one mining district, amongst many scattered throughout the world. As mining costs in the Gillette Coal field rise, mining will shift to other portions of the Powder River Basin and to other coal provinces. On a worldwide basis, coal will be available for a very long time, but coal prices are expected to gradually increase due to increasing mining costs.

In-situ coal gasification is the process of partially burning a coal seam in place to produce a synthesis gas. This technology may greatly expand coal reserves, because deeply buried coal seams may be exploited at a reasonable cost. In-situ coal gasification has its own set of issues, however, which will be discussed in Chapter 4.

We cannot continue to mine coal until we completely exhaust the resource. Other phenomena will curb mining before then. One concept, popular in peak oil discussions, is called Energy Returned on Energy Input (EROEI). This concept states that if the energy consumed in producing energy is greater than energy produced, then energy production will halt regardless of price. Coal production will probably never reach the EROEI limit. Instead, the limit to coal mining will be set by the principle of economic substitution.

Wooly mammoths were once a major human food source. When mammoths were hunted to extinction during the last ice age, our ancestors did not starve. Instead, they found something else to eat. This is an early example of economic substitution.

Coal initially became popular when the growing human population and increased urbanization made wood scarce. The technology of coal use developed to the point that coal became a major source of chemicals, fuel gas and transportation fuel. These coal uses fell out of favor when natural gas and crude oil became abundant and inexpensive in the mid twentieth century. Current interest in coal technology is largely the result of rising natural gas and crude oil prices and concern about future energy supplies.

At the time this was written, almost all coal consumed in the industrialized nations was burned to produce electric power. In the near future, coal-burning power plants will probably be required to capture and sequester carbon dioxide to reduce global warming. When this happens, coal may no longer be a low cost fuel for power generation. The estimated cost ¹⁶. and ¹⁷. of pulverized coal-fired power production with carbon capture and sequestration is about double the cost of pulverized power production without carbon capture and sequestration. Residential power customers would see a 50% increase in their power bills, assuming that distribution costs would change.

At this price, non-fossil fuel sources of electric power, such as nuclear and wind power, are attractive. Currently, solar power is too expensive, but advancing technology promises to lower costs. A challenge for clean coal technology is to produce power and sequester carbon dioxide at a price that is competitive with alternative power sources. Integrated gasification combined cycle (IGCC) coal plants with carbon capture and sequestration (CCS) have attracted a great deal of interest because the cost of electric power from these plants is only 60% higher than conventional pulverized coal generated electricity. ¹⁶ The cost and complexity of replacing coal-fired power plants is enormous, so coal-fired power plants will be a substantial source of electric power for many years.

Coal has a bright future as a raw material for liquid fuels, fuel gas, and chemicals. Projected crude oil prices suggest that liquid fuels will be produced from coal at a lower cost than the same fuels produced from crude oil. Coal technology, and in particular, coal gasification, will have a large role in energy production for many years.

References

1. American Society for Testing and Materials, Annual Book of ASTM Standards 2006, Volume 5 Part 6, Gaseous Fuels; Coal and Coke.

2. M. Seggiani, G. Pannocchia, Prediction of coal ash thermal properties using partial least-squares regression, Ind Eng Chem Res42 ( 2003) 4919– 4926.

3. Baxter L. Brigham Young University. Coal database, www.et.byu.edu/%7Elarryb/CoalDatabase.htm.

4. J.C. van Dyk, F.B. Waanders, Manipulation of gasification coal feed in order to increase the ash fusion temperature of the coal enabling the gasifiers to operate at higher temperatures, Fuel86 ( 2007) 2728– 2735.

5. H. Gan, S.P. Nandi, L. Walker, Nature of the porosity in American coals, Fuel51 ( 1972) 272– 277.

6. N.E. Olague, D.M. Smith, Diffusion of gases in American coals, Fuel68 ( 1989) 1381– 1387.

7. M. Itay, C.R. Hill, D. Glasser, A study of the low temperature oxidation of coal, Fuel Proc Tech21 ( 1989) 81– 97.

8. Y. Yun, H.L.C. Meuzelaar, Development of a reliable coal oxidation (weathering) index—slurry pH and its applications, Fuel Proc Tech27 ( 1991) 179– 202.

9. J. Hayashi, S. Aizawa, H. Kumagai, et al., Evaluation of a brown coal by means of oxidative degradation in aqueous phase, Energy & Fuels13 ( 1999) 69– 76.

10. Balasubramani R. Calorimetric Investigation of the kinetics of low-temperature oxidation of dry coal, M.S. Thesis, University of Wyoming (2003).

11. Sethi VK, Dunlop DD. A coal upgrading technology for sub-bituminous and lignite coals, < www.westernresearch.org/uploadedFiles/Energy_and_Environmental_Technology/Coal/Upgrading_(Including_Headwaters)/CoalUpgradingFMI.pdf>.

12. Evergreen Energy, K-Fuel and K-Direct, < www.evgenergy.com/fact_sheets/K-Fuel.pdf>.

13. U.S. Energy Information Administration, U.S. coal reserves, 1997 Update, DOE/EIA-0529(97), 1999.

14. U.S. Energy Information Administration, < www.eia.doe.gov/fuelcoal.html>.

15. Luppens JA, Scott DC, Haacke JE, et al. Assessment of coal geology, resources, and reserves in the Gillette Coalfield, Powder River Basin, Wyoming, U.S. Geological Survey, Open-File Report 2008-1202 (2008).

16. Katzer J, Ansolabehere S, Beer J, et al. The future of coal, options for a carbon-constrained world, an interdisciplinary MIT study (2007).

17. Woods MC, Capicotto PJ, Haslbeck JL, et al. Cost and performance baseline for fossil energy plants, volume 1: Bituminous coal and natural gas to electricity, DOE/NETL-2007/1281, < www.netl.doe.gov/energy-analyses/pubs/Bituminous%20Baseline_Final%20Report.pdf> (2007).

Chapter 2. Non-gasification Uses of Coal

Contents

Home Heating and Cooking vs. Industrial Use 17

Coal Combustion Pollutants 17

Pulverized Coal Combustion 19

Supercritical Pulverized Coal Combustion 20

Carbon Capture with Pulverized Coal Combustion Plants 21

Oxy-combustion 24

Sargas 27

Coal-to-liquids 28

ENCOAL 28

Direct Hydrogenation of Coal 30

References 33

Home Heating and Cooking vs. Industrial Use

The simplest use of coal is to burn it for heat. Coal was once used as a household heating and cooking fuel in Western nations; but it was largely replaced by natural gas, propane, electricity and fuel oil. Coal is still used for household heating and cooking in China, where it is a major source of air pollution.

In Western nations, coal is used primarily as a fuel for large industrial boilers, especially for electric power generation. Large users are able to get more complete combustion, which reduces odor and soot, and are able to install complex and expensive air pollution equipment. Since pollution control equipment strongly influences the configuration of a modern coal-burning plant, emissions from coal combustion will be described next.

Coal Combustion Pollutants

The US Environmental Protection Agency developed a list of priority pollutants, which are common air pollutants that are primarily generated by combustion. The following is a partial list of these pollutants.

SO x consists primarily of SO 2 but may also contain small amounts of SO 3. In the atmosphere, SO 2 oxidizes to SO 3. This combines with water to form sulfuric acid, H 2SO 4, the primary acid component in acid rain. Combustion of sulfur-containing fuels creates SO x, and coal typically has high sulfur levels compared to other fossil fuels.

NOx consists of several nitrogen-oxygen compounds that contribute to photochemical smog, ozone depletion and global warming. There are two primary sources of NO x. Fuel NO x forms when nitrogen-containing fuel is burned. Not all nitrogen in the fuel forms NO x. Some of the fuel nitrogen may be converted to N 2. Thermal NO x is created by direct combination of N 2 and O 2 in a flame. Thermal NO x is favored by high flame temperatures and high oxygen concentrations. At ambient conditions, NO x is not thermodynamically stable; but it is very difficult to decompose once formed.

CO is formed when carbon-containing fuels are burned. In a flame, carbon is burned to form CO; which is then further oxidized, at a slower reaction rate, to CO 2. In nearly all combustion processes, some of the intermediate product, CO, escapes into the flue gas. Carbon monoxide emissions are favored by low oxygen/fuel ratios.

Particulates are divided into two categories, PM 10, which consists of particles less than 10 microns in diameter; and PM 2.5, a subset of PM 10 which consists of particles less than 2.5 microns in diameter. When inhaled, these particles, especially PM 2.5, tend to remain in the lungs. This can lead to chronic health conditions such as black lung in coal miners, silicosis in people who have prolonged exposure to dust and smoker’s lungs. Some dust is generated when coal is mined, crushed, and shipped. When coal is used for home heating and cooking, the flue gas can contain significant quantities of soot, which is a fine carbon-rich dust. In industrial boilers, combustion is more complete and little soot is produced. Particulate emissions are primarily due to fly ash, which are the fine ash particles entrained in the flue gas.

Volatile organic compounds, VOCs, are nearly all organic compounds that have a significant vapor pressure at ambient conditions. In home heating and cooking applications, VOCs in the flue gas cause disagreeable odors. Since combustion is more complete in industrial boilers, little odor is produced. VOCs are a major issue in organic chemical plants, including coal-to-chemical and coal-to-liquid fuels plants.

Air toxics include a long list of specific toxic compounds. Coal contains small quantities of volatile heavy metals; which vaporize during combustion and may leave with the flue gas. Mercury¹ has received the most attention. Mercury is a neurotoxin, and tends to accumulate in aquatic systems. Mercury bio-accumulates, meaning that large fish that eat smaller mercury-containing fish do not excrete the mercury. Consequently, mercury concentrations are highest at the top of the food chain, including large fish and the people who eat them. The US Environmental Protection Agency issued the first Clean Air Mercury Rule in 2005.

Greenhouse gasses include CO 2, CH 4, and NO x compounds. Because of the large volume emitted, CO 2 has received the most attention. Most members of the scientific community believe that global warming is largely due to greenhouse gasses released by fossil fuel combustion. Since coal has lower H/C ratios than other fossil fuels, coal combustion releases more CO 2 per unit of energy than other fossil fuels.

Pulverized Coal Combustion

The most common type of coal-fired power plant is pulverized coal combustion (PCC), shown in Figure 2.1. A mixture of pulverized coal and air is blown into a low NO x burner. This burner has an annular arrangement. Coal and a portion of the air are fed to the center tube. The remainder of the air is fed through the space between the inside and outside tubes. The main portion of the flame has a low

Enjoying the preview?
Page 1 of 1