Flue Gas Heat Recovery in Power Plants, Part I

Embed Size (px)

Citation preview

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    1/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 1/12

    Power 101: Flue Gas Heat Recovery in

    Power Plants, Part I

    04/01/2010 | By Nenad Sarunac

    Every power engineer must have a firm grasp of the rudiments of how fuel is

    processed to produce electricity in a power generation facility. With this article,

    we begin a series of Power 101 tutorials that present these fundamentals in a

    clear and concise way. First up are the essentials of recovering heat from flue

    gas.

    The major operating cost of a coal-fired power plant is for fuel. Given the higher

    heating value (HHV), the amount of coal required to generate a desired power

    output depends on unit efficiency (net unit heat rate). Therefore, unit efficiency

    is an important economic factor, and recovering heat from the stack gas will

    further improve the plant economics.

    Some Preliminaries

    A relationship between the net unit efficiency and net unit heat rate (HRnet) is

    presented in Figure 1. [Download this ppt file [1] to view all figures at a legible

    size.] Thermal efficiency, or efficiency (), is defined as the electric energy

    output as a fraction (or percentage) of the fuel energy input. Heat rate is an

    inverse of efficiency (multiplied by the unit conversion factor of 3,412). Both theefficiency and heat rate can be expressed on an HHV or a lower heating value

    (LHV) basis. In the U.S., HHV is used for the coal-fired power plants, while in

    Europe, efficiency calculations are based on LHV. A recent article in POWER[2]

    provided a comprehensive discussion of power plant efficiency. Bottom line: Be

    careful when comparing efficiencies from different data sources. To avoid

    confusion, a note on HHV basis or LHV basis should be added next to the

    numerical value of efficiency or heat rate.

    Reference also is often made in the literature to changes in efficiency by

    percentage points (%-points), which should be distinguished from relative

    changes in percentage. For example (Figure 1), a change of 1%-point in

    http://www.powermag.com/Assets_coal/File/CP%20Apr%2010%20HR%20Figures.ppthttp://www.powermag.com/coal/2432.htmlhttp://www.powermag.com/Assets_coal/File/CP%20Apr%2010%20HR%20Figures.ppt
  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    2/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 2/12

    efficiency (from 36% to 37%) represents a relative change of 2.7%. The

    difference in efficiency between HHV and LHV for bituminous coal is about 2%-

    points (5% relative), while for the high-moisture subbituminous coals and

    lignites, the difference is about 3% to 4 %-points (8% to 10% relative,

    depending on the coal composition).

    Besides lower fuel cost, reduced fuel use results in lower emissions of NOx, SOx,

    Hg, PM, and other pollutants. Efficiency improvement, as the only practical

    option for reducing CO2 emissions in the short term, has become a key

    consideration when choosing technology for new plants and for upgrades of

    existing power plants. A relationship between heat rate improvement and

    reduction in CO2 emissions, presented in Figure 2, shows that CO2 reduction is

    proportional to the heat rate improvement. That is, 1% improvement in heat

    rate results in 1% reduction in CO2 emissions, regardless of the coal type or its

    rank.

    Savings in the fuel and CO2 emissions cost for a typical 580-MW power plant

    firing Illinois coal are presented in Figure 3 as functions of the heat rate

    improvement and cost per ton of CO2 over an assumed range of carbon

    allowance prices. The results show that 1% improvement in net unit heat rate

    (on a relative basis) results in annual fuel savings of $1.6 million, assuming an

    energy cost of $4/million Btu (MBtu) and unit capacity factor of 85%. With aCO2 cost of $30 per ton, annual savings are almost doubled ($2.8 million/year).

    The efficiency of a coal-fired power plant also will have a strong effect on the

    cost of carbon capture; with higher efficiency, the flow rate of flue gas that

    needs to be treated will be lower, resulting in a smaller and less-expensive

    carbon capture and sequestration (CCS) system. A smaller CCS will have smaller

    negative effect on plant efficiency.

    There are numerous opportunities and options for improving the efficiency of

    existing power plants. Utilization of waste heat for boiler efficiency

    improvement, improvement of steam turbine cycle heat rate, and stack reheat

    are described below. Boiler efficiency improvement achieved by using heat

    recovered from the flue gas for drying of high-moisture and washed coals will be

    discussed in Part II. The improvement in steam cycle performance achieved by

    using heat recovered from the flue gas for feedwater heating and preheating of

    combustion air will be discussed in Part III.

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    3/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 3/12

    Heat Recovery from Flue Gas

    The temperature of the flue gas leaving the boiler is commonly reduced in an air

    preheater (APH) when the sensible heat in the flue gas leaving the economizer

    is used to preheat combustion air. Preheating of combustion air has a

    significant positive effect on boiler efficiency. Common practice is to recover

    sensible heat from flue gas until the temperature of the flue gas drops toapproximately 300F. The primary impediment to recovering heat by additional

    cooling is the risk of condensing sulfuric acid on the APH heat transfer surfaces

    and downstream ductwork.

    Acid deposition leads to corrosion of affected surfaces, as well as to fouling and

    plugging of the APH heat transfer passages. The APH fouling increases pressure

    drop across the APH (both on the air and flue gas sides), which increases power

    requirements for the forced draft (FD) and induced draft (ID) fans, resulting in

    higher station service power and higher net unit heat rate (lower net unit

    efficiency). Higher pressure drops also lead to higher pressure differentials

    between the air and flue gas streams, which result in higher air-to-flue gas

    leakage. Higher leakage increases fan power requirements and increases the flow

    rate of flue gas through the pollution control equipment.

    Sulfuric acid in the flue gas is formed in gas-phase reactions of SO3 and H2O

    upstream of the APH. The SO3 is formed from SO2 by homogeneous and

    heterogeneous reactions in the furnace and convection pass of the boiler. The

    presence of SO3 in the flue gas increases the dew point of the flue gas. The acid

    dew point temperature is presented in Figure 4 as a function of the SO3 and

    H2O concentration in the flue gas. Sulfuric acid condenses as temperature is

    decreased bellow the dew point temperature. The condensed sulfuric acid (acid

    and water mixturesulfuric acid is hydroscopic) is corrosive to the inexpensive

    materials used in construction of the APH heat transfer surfaces anddownstream ductwork. Heat rate improvement that could be achieved by

    increasing heat transfer in the APH is typically inadequate to justify higher-cost

    corrosion-resistant materials that would be needed for the APH cold end,

    electrostatic precipitator (ESP), and the ductwork, and to deal with higher APH

    fouling and plugging.

    Assuming the SO3 concentration in the flue gas of 5 ppm and bituminous coal

    (H2O concentration in the flue gas of 8% by volume) gives the acid dew pointtemperature of approximately 263F. It has to be noted that although the flue gas

    temperature at the APH exit is typically 50F or more above the acid dew point

    temperature, metal surface temperatures could be below that, as in the case of

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    4/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 4/12

    the Ljungstrom APH. In a Ljungstrom APH, temperature of the heat transfer

    surfaces located in the cold end layer of the APH are controlled by the inlet air

    temperature and are significantly lower than the acid dew point temperature.

    More details on the APH heat transfer, performance, and fouling and plugging

    will be presented in Part II.

    Besides acid deposition, the other impediment to recovering heat from the flue

    gas by additional cooling in the APH is the ESP performance. As presented in

    Figure 5, resistivity of flyash decreases as the flue gas temperature is reduced

    below 300F. However, in case of the high-resistivity ash (Figure 5), the

    temperature reduction would not be a problem, for the low-resistivity ash low

    flue gas temperatures will have a significant negative effect on the ESP

    performance.

    Steam generators that employ ammonia injection for selective catalytic

    reduction (SCR) or selective noncatalytic reduction (SNCR) of NOx encounter an

    additional challenge in design and operation of low-temperature heat-recovery

    equipment, particularly APHs. Unreacted ammonia combines with SO3 in the

    flue gas stream and SO3 produced on the SCR catalysts to form ammonium

    bisulfate (ABS). The ABS forms in a temperature range between the APH flue gas

    inlet and outlet temperatures. The deposits are sticky and corrosive to steels

    commonly employed in the APHs.

    Upon exiting the ESP, it is common to cool the flue gas by evaporative cooling

    to a temperature close to the adiabatic saturation temperature by spraying water

    into the flue gas stream within a wet flue gas desulfurization (FGD) system.

    According to an FGD manufacturer, the optimal flue gas temperature for a

    desulfurization process is approximately 149F (65C). Cooling of the flue gas to

    the saturation temperature occurs in a spray area, and the flue gas leaves the

    FGD reactor at a temperature close to the saturation temperature. In same cases,the flue gas is leaving the FGD in a supersaturated state with a temperature

    slightly below the saturation temperature. This practice results in significant

    use of water for evaporative cooling. More importantly, the sensible heat of flue

    gas is not beneficially used.

    Saturation temperature is a function of the moisture content in the flue gas,

    which depends on the total coal moisture (TM) content of coal, and humidity of

    inlet air. The moisture content of the flue gas is presented in Figure 6 as a

    function of the total coal moisture content. Calculations were performed for the

    excess air coefficient (E) of 17.2% and humidity of inlet air of 0.01149 mole

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    5/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 5/12

    H2O/mole air.

    The flue gas moisture content and saturation temperature for the bituminous

    and washed Illinois coals, subbituminous (Powder River Basin, PRB) coals, and

    lignites are summarized in Table 1. The saturation temperature is presented in

    Figure 7 as a function of TM. For these coals and combustion conditions,

    saturation temperature varies in the 104F to 134F range.

    Table 1. Saturation temperature for various coals. Source: Energy Research

    Center

    Low-Temperature Heat Recovery

    The amount of heat available in the flue gas is presented in Figure 8 as a

    function of the flue gas temperature for four coals: bituminous, washed Illinois,PRB, and lignite. Washed Illinois coals contain significant amount of moisture

    (18% to 22%, or more), which reduces its HHV value. Most of this moisture is

    surface moisture, which can be removed by drying. As the flue gas temperature

    is reduced, the amount of available sensible heat increases. The sensible heat

    that could be recovered from the flue gas by cooling it in a flue gas cooler (FGC)

    from the APH gas outlet temperature of 310F to the FGD inlet temperature of

    140F is 43 to 46 Btu per pound of flue gas (depending on the coal type).

    As the flue gas is cooled below its saturation temperature, flue gas moisture

    condenses to maintain partial pressure of the water vapor in flue gas that is

    consistent with the flue gas temperature, and water vapor content of the flue

    gas decreases (Figure 9). The amount (mass) of condensed water increases as the

    temperature of the flue gas is reduced and is a strong function of the coal

    moisture content.

    As moisture condenses out of the flue gas steam, the flow rate of flue gasdecreases, causing a slight decrease in the amount of sensible heat in flue gas

    (Figure 8). The kink in the sensible heat curve occurs at the saturation

    temperature of flue gas.

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    6/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 6/12

    As shown in Figure 9, most of the moisture can be removed from the flue gas by

    cooling it to a very low temperature. The chilled ammonia concept, developed

    by Alstom Power, employs cooling of the flue gas to a very low temperature

    using chillers. At the current state of technology development, such low-

    temperature cooling of the flue gas is expensive due to high power requirements

    for the chillers.

    Condensation of the flue gas moisture liberates latent heat. The amount of

    latent heat released is a function of the flue gas temperature and coal type

    (Figure 10). The amount of released latent heat increases as TM content of the

    coal increases and temperature of the flue gas decreases. The latent heat can be

    recovered in condensing heat exchangers (CXEs), but due to the low

    temperature of a cooling fluid, there are practical temperature limits

    (approximately 100F to 110F) that impose limits on the amount of latent heat

    than can be economically recovered from the flue gas. Available heat sinks limit

    the amount of low-temperature heat that can be beneficially used.

    The total (sensible and latent) heat of the flue gas is presented in Figure 11. As

    the flue gas is cooled below its saturation temperature, the amount of total heat

    greatly increases. However, as discussed previously, there are practical

    limitations associated with cooling of the flue gas to low temperatures and

    beneficial use of the recovered low-temperature heat.

    To illustrate total amount of heat available in the flue gas, sample calculations

    were performed for a conventional supercritical pulverized coal-fired power

    plant and four different coals. The gross power output of a 642.18-MW, turbine

    cycle heat rate (HRcycle) of 7,467 Btu/kWh (cycle,gross = 45.69%), APH leakage

    of 10%, flue gas temperature at the boiler outlet of 680F, and coal TM content

    from Table 1 were assumed in the calculations. The results are presented in

    Table 2 and Figure 12. The sensible heat was determined for the case where theflue gas is cooled in a FGC located upstream of the FGD from a temperature of

    310F (APH gas outlet) to 140F (FGD inlet).

    Table 2. Baseline power plant. Source: Energy Research Center

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    7/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 7/12

    The total heat (including latent heat) was determined for the case where the

    flue gas was cooled in the FGC and CXE from a temperature of 310F (APH gas

    outlet) to 110F (inlet to a CO2 absorber). Flue gas cooling down to about 110F

    will be needed for efficient operation of the post-combustion CO2 capture

    system (CO2 absorber). Even deeper cooling is needed for the chilled ammonia

    process.

    The results show that for low-moisture bituminous and washed Illinois coals

    having low saturation temperatures, the benefit of cooling the flue gas down to

    110F is small. As TM content of the coal increases, such for the PRB and

    lignites, the amount of total heat increases significantly, especially for the

    lignites. Therefore, for high-moisture coals it might be economical to recover

    the low-temperature heat. This is not the case for the low-moisture coals,

    where cooling of the flue gas in a FGC upstream of the FGD is most economical

    option.

    Feedwater Heating and Combustion Air Preheat

    The technology to recover low-temperature heat from flue gas originated in

    Europe, where it has been used to improve performance of coal-fired power

    plants and industrial plants for more than 15 years. Utility companies such as

    RWE Power, Vattenfall, and others utilize the low-temperature heat from flue

    gas for feedwater (FW) heating and preheating of combustion air. Several

    different configurations with different commercial names, such as Powerise,

    were developed and successfully used at power plants such as Schwarze Pumpe,

    Mehrum, Niederaussem, Lippendorf, and Werndorf in Germany; Voitsberg in

    Austria; and at other locations, including industrial plants and waste-to-energy

    plants, such as Vestforbraending in Denmark, where recovered low-temperature

    heat is used for district heating.

    Typical configurations for using low-temperature heat from the flue gas include

    configurations allowing FW heating and preheating of the combustion air. The

    low-temperature heat is recovered from the flue gas using the flue gas cooler

    (FGC) located upstream of the FGD. A configuration was developed for post-

    combustion CO2 capture retrofit or new construction, where the flue gas is

    cooled to the 105F to 110F range. This configuration includes the FGC upstream

    of the FGD and a condensing heat exchanger (CXE) downstream of the FGD andupstream of the CO2 absorber. Lets estimate the performance benefits achieved

    by using recovered low-temperature heat.

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    8/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 8/12

    To illustrate the benefits of using heat recovered from the flue gas for the FW

    heating, and preheating of combustion air, analyses were performed for a

    baseline power plant configuration presented in Figure 13 and three coals:

    washed Illinois, PRB, and lignite. The baseline configuration is a conventional

    coal-fired power plant employing a boiler, steam turbine cycle with seven stages

    of regenerative heating of the condensate, and a FGD for SOx control.

    Temperature of the condensate leaving the main steam condenser is, in this

    example, 85.9F. Note that the condenser outlet temperature is highly site-

    specific and depends on the temperature of the cooling water into the

    condenser, condenser cleanliness, and state of maintenance. Temperature of

    the cooling water is subject to seasonal variations and location of the plant. For

    plants equipped with a cooling tower, performance of the cooling tower adds

    another level of complexity, as its performance is affected by the ambient and

    process conditions. Combustion air is preheated in a steam air heater (SAH)

    using steam extracted from the steam turbine cycle. The flow and temperature

    data presented in Figure 13 correspond to lignite. The results for all three coals

    are summarized in Tables 3 through 5.

    Table 3. Increase in power output compared to baseline shown in Table 2.

    Source: Energy Research Center

    Table 4. Improvement in net unit heat rate compared to baseline shown in

    Table 2. Source: Energy Research Center

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    9/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 9/12

    Table 5. Increase in net unit efficiency compared to baseline shown in

    Table 2. Source: Energy Research Center

    The first investigated configurationConfiguration A, for using low-

    temperature heat from the flue gas, involving a FGC upstream of the FGDis

    presented in Figure 14. Instead of using steam extracted from the steam turbine

    cycle for the combustion air preheat, combustion air is preheated by the heatrecovered from the flue gas stream. This increases steam flow through the low-

    pressure (LP) turbine with a resulting increase in the steam turbine power

    output. The increase in the turbine power output results in an improvement in

    turbine cycle heat rate and, ultimately, in the net unit heat rate. One effect was

    that the heat rejected by the condenser and the condensate flow increased.

    Also, the amount of heat supplied by the extraction steam and recovered from

    the flue gas were matched to achieve the same level of combustion air preheat.

    Finally, the feedwater temperature entering the boiler was kept constant for allanalyzed cases.

    The second configuration, Configuration B, uses low-temperature heat from the

    flue gas and includes a FGC upstream of the FGD (Figure 15). One hundred

    percent of the condensate flow leaving the main steam condenser flows through

    the FGC, where it is heated. The heated condensate is circulated back to the

    steam turbine cycle, bypassing low-pressure feedwater heaters (FWH) 6 and 7.

    This arrangement eliminates low-pressure steam extractions, and the steam

    that would normally be used in the FWH6 and FWH7 is expanded in the LP

    turbine. The result is an increase in the steam turbine power output, increase in

    steam flow to the condenser and main condensate flow, and increase in heat

    rejected by the main steam condenser. The increase in turbine power output

    results in an improvement in turbine cycle and net unit heat rates. In this

    example, the flue gas is cooled to a temperature of 135F. Combustion air is

    preheated by steam extracted from the steam turbine cycle.

    Configuration C represents a combination of Configurations A and B, where a

    portion of heat recovered from the flue gas is used for FW heating, while the

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    10/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 10/12

    remaining heat is used for the combustion air preheat. A schematic of

    Configuration C is shown in Figure 16. For clarity, the FGC is divided into two

    parts, where FGC1 is used for the combustion air preheat and FGC2 is used for

    the FW heating.

    Configurations D and E allow cooling of the flue gas to the 105F to 110F range,

    which is required for post-combustion CO2 capture. Configuration D (Figure 17)

    is a variant of Configuration B and incorporates a CXE downstream of the FGD

    and upstream of the CO2 absorber (indicated as CCS). Sensible heat of the flue

    gas is recovered in the FGC. The first stage of the condensing heat exchanger

    (CXE1) is recovering sensible and latent heat from the flue gas. The recovered

    heat is used for the FW heating. The second CXE stage (CXE2) is used to further

    reduce the flue gas temperature and decrease the flue gas moisture content.

    Please note that the flue gas is exiting the FGD in a saturated or supersaturated

    state. Reduced moisture content in the flue gas has a positive effect on the

    efficiency of the CO2 absorption/desorption process. Cooling the flue gas to

    approximately 105F removes about half of the moisture from the flue gas

    stream. The recovered heat is very low in temperature and has limited use, such

    as building heating. The alternative is to use a spray cooler instead of CXE2.

    However, in such cases the flue gas is entering the CO2 absorber in a saturated

    or supersaturated state. The high moisture content of the flue gas has a negativeeffect on the efficiency of the CO2 absorption/desorption process and

    equipment size. A tradeoff analysis is needed to determine the most cost-

    effective configuration, and that is outside of the scope of this article.

    Configuration E (Figure 18) is a variant of Configuration C. Similar to

    Configuration D, it employs a FGC and a two-stage CXE. The recovered latent

    and sensible heat is used for the FW heating and preheating of combustion air.

    The results are summarized in Tables 3 through 5 and are presented in Figures

    19 through 21 for the cycles discussed above. The potential improvement

    depends on the configuration and coal type. The heat rate improvement varies

    from 1.24% to 3.65% relative, considering all configurations. Please note that

    reductions in CO2 emissions are directly proportional to the heat rate

    improvements.

    For configurations not including the CXE, the improvement is lower, from

    1.24% to 3.18% relative, but still significant. The corresponding improvement

    in net unit efficiency is from 0.51%-points to 1.39%-points considering all

  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    11/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/?printmode=1 11/12

    1. http://www.powermag.com/Assets_coal/File/CP%20Apr%2010%20HR%20Figures.ppt

    2. http://www.powermag.com/coal/2432.html

    3. http://www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-

    i/'+String.fromCharCode(110,115,48,49,64,108,101,104,105,103,104,46,101,100,117)+'?'

    configurations. For configurations not employing the CXE, the improvement in

    net unit efficiency ranges from 0.51%-points to 0.90%-points.

    Performance improvements for Configuration A are relatively insensitive to the

    coal. For Configurations B, C, D, and E, potential performance improvement

    typically increases with the increase in coal moisture content and is highest for

    lignites. At another subcritical unit, the higher temperature of the condensate

    leaving the condenser (105.3F vs. 85.9F) for low- and mid-moisture fuels

    showed approximately 0.2% lower improvement in net unit heat rate compared

    to our finding summarized in Figure 20.

    In summary, performance improvements achievable by using heat recovered

    from the flue gas for FW heating and combustion air preheat can be significant

    and should be considered as measures for improving performance and reducing

    emissions for existing and newly constructed power plants. For existing power

    plants, where it is difficult or impossible to raise steam parameters to improve

    performance of the steam turbine cycle, using heat recovered from the flue gas is

    an attractive alternative. Optimization of the system configurationsuch as

    temperature of the preheat air leaving the APH and FW bypass (fraction of the

    FW flow bypassing low-pressure FWHs [100% bypass was used in this work])is

    a necessary part of any robust plant design.

    More to Come

    In Part II, well examine the types of coal-drying technologies available, their

    performance, and operating economics. In Part III, well look at options for flue

    gas reheat, feedwater heating, and combustion air preheating.

    Nenad Sarunac ([email protected] [3]) is principal research engineer andassociate director at Energy Research Center, Lehigh University. The Illinois

    Clean Coal Institute funded a portion of this work.

    Print Mode [4] : ON

    http://void%280%29/http://www.powermag.com/power-101-flue-gas-heat-recovery-in-power-plants-part-i/'+String.fromCharCode(110,115,48,49,64,108,101,104,105,103,104,46,101,100,117)+'?%27
  • 7/27/2019 Flue Gas Heat Recovery in Power Plants, Part I

    12/12

    10/13/13 Power 101: Flue Gas Heat Recovery in Power Plants, Part I | POWER Magazine

    4. javascript:void(0)