Sie sind auf Seite 1von 104

WARNING: Please read the License Agreement on the back cover before removing the Wrapping Material

Small Gas Turbines for Distributed Generation Markets


Technology, Products, and Business Issues A Product of EPRIsolutions, Inc., and GTI
1000768 GTI-00/0219

Effective December 6, 2006, this report has been made publicly available in accordance with Section 734.3(b)(3) and published in accordance with Section 734.7 of the U.S. Export Administration Regulations. As a result of this publication, this report is subject to only copyright protection and does not require any license agreement from EPRI. This notice supersedes the export control restrictions and any proprietary licensed material notices embedded in the document prior to publication.

Small Gas Turbines for Distributed Generation Markets


Technology, Products, and Business Issues 1000768 GTI-00/0219 Technical Progress, December 2000

EPRIsolutions Project Manager K.R. Amarnath

GTI Project Manager W.E. Liss

EPRIsolutions 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10414, Palo Alto, California 94303 USA 800.313.3774 650.855.2121 askepri@epri.com www.epri.com

Gas Technology Institute 1700 South Mount Prospect Road, Des Plaines, IL 60018-1804 847.768.0500 www.gastechnology.org

DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES


THIS REPORT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN ACCOUNT OF WORK SPONSORED OR COSPONSORED BY EPRISOLUTIONS, INC. NEITHER EPRISOLUTIONS, THE ELECTRIC POWER RESEARCH INSTITUTE, INC. (EPRI), ANY MEMBER OF EPRI, ANY COSPONSOR, THE ORGANIZATION(S) NAMED BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM: (A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS REPORT, INCLUDING MERCHANTABILITY AND FITNESS FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTYS INTELLECTUAL PROPERTY, OR (III) THAT THIS REPORT IS SUITABLE TO ANY PARTICULAR USERS CIRCUMSTANCE; OR (B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRISOLUTIONS OR ANY EPRISOLUTIONS REPRESENTATIVE HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS REPORT OR ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS REPORT. ORGANIZATION(S) THAT PREPARED THIS REPORT Steven I. Freedman Onsite Energy Corp.

This is an EPRIsolutions Level 2 report. A Level 2 report is intended as an informal report of continuing research, a meeting, or a topical study. It is not a final EPRIsolutions technical report.

ORDERING INFORMATION
Requests for copies of this report should be directed to the EPRI Distribution Center, 207 Coggins Drive, P.O. Box 23205, Pleasant Hill, CA 94523, (800) 313-3774. Electric Power Research Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power Research Institute, Inc. EPRIsolutions is a registered service mark of EPRIsolutions, Inc. Copyright 2000 EPRIsolutions, Inc. All rights reserved

CITATIONS
This document was prepared by Steven I. Freedman Engineering Consultant 410 Carlisle Avenue Deerfield, IL 60015 S. Freedman Onsite Energy Corp. 701 Palomar Airport Road, #200 Carlsbad, CA 92009 P. Bautista This document describes research sponsored by EPRIsolutions and GTI. The publication is a corporate document that should be cited in the literature in the following manner: Small Gas Turbines for Distributed Generation Markets: Technology, Products, and Business Issues, EPRIsolutions, Palo Alto, CA, and GTI: 2000. 1000768. GTI-00/0219.

iii

EPRIsolutions Licensed Material

ABSTRACT
Small gas turbines (300 kW to 5 MW) offer an attractive way for utilities and energy service companies to generate electric power within distribution grids and for consumers to generate their own power. Distributed generation also benefits utilities by deferring or avoiding costly expansion of the power transmission and distribution system, which could allow them to offer customers lower cost power. Gas turbines process more power-generation cycle air per unit size and weight of machine than do reciprocating engines and, consequently, are lighter weight and more compact. This size and weight advantage could reduce gas turbine cost relative to reciprocating engines if business conditions warrant production of small gas turbines at volumes equal to or greater than reciprocating engines. The characteristics of gas turbines include clean emissions, good efficiency in some models, high reliability, low noise and vibration, compact installed footprint (sq ft/kW), a good match with exhaust-fired steam generation boilers, and infrequent maintenance and overhaul. However, small gas turbines are not generally sufficiently economically attractive at the present time for intermediate-duty, daily cycling applications in competition with deregulated grid prices or reciprocating engine generators. The high technology of the newest, very large gas turbines both heavy-duty industrial and aeroderivative machines I has not yet found its way into the smaller sizes. The cost of applying known technology is still unjustifiably high considering the limited size of the current market and the manufacturing difficulties involved. As deregulation progresses and more commercial and light industrial businesses and electric utilities take advantage of distributed generation options, additional sales of small gas turbines should result in manufacturing economies of scale and price reductions. Very small gas turbines, 3 MW or less, are noticeably higher priced per kilowatt than competing reciprocating engines, due principally to low production volumes and low commonality of parts among multiple turbine models. Reciprocating engines benefit from their modularity of piston and cylinder configuration, allowing multiple power capacities to be delivered from the same basic engine family or design. In the 3-5 MW size range, gas turbines are manufactured in greater volumes, and their compactness and high power density begin to make them more economically competitive. In sizes above 5 MW, gas turbines have an inherent economic advantage over reciprocating engines in that they process more air per unit volume of machine, and their power generation efficiencies begin to approach those of reciprocating engines. As distributed generation applications increase, manufacturing economies of scale could make small gas turbines an economic source of power or combined heat and power to commercial and small industrial customers with high intermediate power prices.

EPRIsolutions Licensed Material

CONTENTS

1 2

INTRODUCTION ...............................................................................1-1

Market Background................................................................................................. 1-1 Competitive Positioning of Small Gas Turbines ...................................................... 1-1 Report Objective ..................................................................................................... 1-4

TECHNOLOGY DESCRIPTION ........................................................2-1

Basics ..................................................................................................................... 2-1 Types of Gas Turbines ...................................................................................... 2-2 Ratings .............................................................................................................. 2-3 Fuels.................................................................................................................. 2-3 Fuel Pressure Requirements ............................................................................. 2-4 Operational Characteristics ............................................................................... 2-6 Thermodynamic Cycles .......................................................................................... 2-7 Simple Cycle...................................................................................................... 2-7 Recuperated Cycle ............................................................................................ 2-8 Intercooled Cycle............................................................................................... 2-8 Intercooled Recuperated Cycle ......................................................................... 2-9 Design Considerations.......................................................................................... 2-12 Recuperated Cycle Design .............................................................................. 2-12 Intercooled Cycle Design................................................................................. 2-13 Combined Intercooled and Recuperated Cycle Design ................................... 2-15 Size Considerations .............................................................................................. 2-16 Gas Turbine Systems ...................................................................................... 2-16 Gas Turbine Component Selection.................................................................. 2-17 Example of a Tradeoff ..................................................................................... 2-18 The Use of Steam to Increase Efficiency .............................................................. 2-19 Steam Injection................................................................................................ 2-19 Combined Cycles ............................................................................................ 2-19 High-Temperature Materials and Component Engineering................................... 2-20 Turbine Blade Cooling ..................................................................................... 2-21 Competing Technology Reciprocating Engines ................................................. 2-22

vii

EPRIsolutions Licensed Material

Heat Recovery and Cogeneration ................................................................... 2-23 Noise ............................................................................................................... 2-23 Emissions and Environmental Concerns .............................................................. 2-24 Advanced Gas Turbine Cycles ............................................................................. 2-25 Recent Developments in Advanced Cycles..................................................... 2-25 Wet Compression ............................................................................................ 2-27 Investment in the Development of New Products ............................................ 2-28 Alternative Fuels.............................................................................................. 2-29

EMISSION CONTROL & CATALYTIC COMBUSTION.....................3-1

Introduction ............................................................................................................. 3-1 Emission Control Technology ................................................................................. 3-1 Catalytic Combustion for Gas Turbines .................................................................. 3-2 Catalytic Combustion System Developers .............................................................. 3-3 Alzeta ................................................................................................................ 3-3 Catalytica........................................................................................................... 3-4 PCI .................................................................................................................... 3-4 Low Emission Gas Turbine Projects ....................................................................... 3-5 Economics .............................................................................................................. 3-6

PROVIDERS .....................................................................................4-1

Company Profiles.................................................................................................... 4-8 Alstom Power .................................................................................................... 4-9 Dresser-Rand .................................................................................................... 4-9 Ebara............................................................................................................... 4-10 GAS Power Systems ....................................................................................... 4-11 Honeywell........................................................................................................ 4-12 Kawasaki Gas Turbine .................................................................................... 4-13 Nuovo Pignone ................................................................................................ 4-14 OPRA Optimal Radial Turbine......................................................................... 4-15 Orenda ............................................................................................................ 4-16 Pratt & Whitney Canada .................................................................................. 4-16 Ramgen Power Systems ................................................................................. 4-17 Rolls-Royce ..................................................................................................... 4-18 Schelde Heron B.V. ......................................................................................... 4-19 Solar Turbines ................................................................................................. 4-20

viii

EPRIsolutions Licensed Material

Stewart & Stevenson ....................................................................................... 4-21 Turbomeca ...................................................................................................... 4-21 Vericor ............................................................................................................. 4-21 Yanmar Diesel Engine Co. Ltd. ....................................................................... 4-22

ECONOMICS & BUSINESS ISSUES ................................................5-1

Introduction ............................................................................................................. 5-1 Regulated Utilities The Previous Era .............................................................. 5-1 Deregulation The Present Era ........................................................................ 5-2 Applications ............................................................................................................ 5-3 Combined Heat and Power ............................................................................... 5-3 Standby Generation........................................................................................... 5-4 Intermediate Duty .............................................................................................. 5-4 Competitive Positioning of Small Gas Turbines ...................................................... 5-5 Capital Costs .......................................................................................................... 5-7 Generator Set Equipment.................................................................................. 5-8 Switchgear and Utility Interconnection............................................................... 5-8 Heat Recovery Equipment................................................................................. 5-9 Emission Control Equipment ............................................................................. 5-9 Total Installed Costs ........................................................................................ 5-11 Operation & Maintenance Costs ........................................................................... 5-11 Non-Fuel O&M................................................................................................. 5-11 Fuel ................................................................................................................. 5-12 Cost of Electricity .................................................................................................. 5-12 Market Outlooks.................................................................................................... 5-13 CHP Market ..................................................................................................... 5-13 Standby Market ............................................................................................... 5-14 Intermediate-Duty Market ................................................................................ 5-14

POWER GENERATION HISTORY......................................................... A-1


Gas Turbine History ................................................................................................A-1 Early Development ............................................................................................A-1 Small Gas Turbines for Automotive and Other Vehicle Applications .................A-2 Microturbines .....................................................................................................A-3

BIBLIOGRAPHY ..................................................................................... B1

ix

EPRIsolutions Licensed Material

1
INTRODUCTION
Market Background Continued market growth is expected for natural gas-fueled prime movers, primarily turbines and reciprocating engines. Gas turbines in both simple- and combined-cycle systems have accounted for the vast majority of power generation capacity added in the last five years in both international and U.S. markets. These are predominantly central station power plants greater than 50 MW. In short, large gas turbines have become the power generation technology of choice. This trend is expected to continue over the foreseeable future. Several factors contribute to the strong position of gas turbine-based power generation: 1. An optimistic outlook for the supply and price of natural gas 2. Technology advances that produced substantial improvements in efficiency and emissions 3. Emissions regulations that could favor gas turbine projects over traditional fossil-fueled steam turbines 4. Attractive initial capital costs and reduced time and cost for power plant permitting and installation, compared to traditional power plants. While the market and business climate is quite favorable for large gas turbines, gas turbines in the distributed generation market face greater challenges. Fundamental market drivers favor large gas turbine power plants because of lower capital costs and shorter construction and permitting lead times than traditional fossil-fueled steam turbines. Large combined-cycle systems have efficiencies in the 50-58% range, based on the fuels lower heating value (LHV). The environmentally clean nature of the these plants is evidenced by their ability achieve 9 ppm of nitrogen oxides (NOx) emissions without exhaust treatment and lower than 3 ppm with postcombustion control technologies. Competitive Positioning of Small Gas Turbines Gas turbines start to lose their clear advantage at smaller sizes. In the industrial market segment (3.5-30 MW), worldwide sales of gas turbines and reciprocating engines are about equal. The size range below 5 MW is dominated by reciprocating engines, both natural gas and diesel fueled. As gas turbines decrease in size, they exhibit higher specific capital costs ($/kW) and lower efficiencies. For example, in the 2-5 MW size range, gas turbines and reciprocating engines have comparable capital costs, but the reciprocating engines have substantially better efficiencies, between 37% and 42% (LHV), versus 32% or lower typical of comparably sized gas turbines (Figure 1-1). The new Mercury 50 gas turbine, a recuperated 4-MW machine manufactured by Solar Turbines Incorporated (San Diego), is the exception to this relationship; this unit has an efficiency of about 40%. In the 2-MW and smaller range, reciprocating engines are priced lower than gas turbines and are considerably more efficient. Figure 1-2 shows the price and efficiency of simple-cycle gas turbine products up to 25 MW. Figure 1-3 compares the

1-1

EPRIsolutions Licensed Material Introduction

capital costs of reciprocating engines and gas turbines (also called combustion turbines, CT in the figure) up to 7 MW. The distinct advantages of small gas turbines over reciprocating engines include higher quality recoverable energy, lower emissions, lower maintenance requirements, and higher power density (kilowatt or horsepower per unit of air flow and machine volume). Recoverable energy refers to the amount of energy that can be recovered from the turbine exhaust stream, usually in the form of clean, high-temperature heat. Higher quality recoverable energy allows for a wider range of thermal energy (for example, high-pressure steam) to be generated if needed. Consequently, many small gas turbines are currently deployed in combined heat and power configurations where this recoverable energy can be used and higher total fuel efficiency can be achieved. In addition, gas turbines often have less frequent requirements for routine maintenance compared to reciprocating engines, which need periodic oil and spark-plug maintenance.

Figure 1-1 Small Gas Turbine Product Positioning Source: GTI

1-2

EPRIsolutions Licensed Material Introduction

Gas Turbine Products


Simple Cycle Units
1000 Price ($/kW) Efficiency (%, LHV) 35 , Price Efficiency , 40

800 , 600

,, , ,, , , , , , , ,

30

,, ,, ,, , , , ,

400

25

200 0 5 10 15 20 25 Size (kW, thousands)


Actual purchase prices may vary due to market conditions and other factors Source: Gas Turbine World/GTI

20

Figure 1-2 Small Gas Turbine Products

1000 800 600 400 200

Price ($/kW)
CT Price RE Price

Size (kW, thousands)


Actual purchase prices may vary due to market conditions and other factors. Prices do not include gas compressors (if required). Source: Gas Turbine World/SFA Pacific/GTI

Figure 1-3 Capital Cost Comparison of Small Gas Turbine Products with Reciprocating Engine Products
1-3

EPRIsolutions Licensed Material Introduction

Manufacturers and providers of small gas turbines have long recognized the competitive positioning of their products and have initiated developments to enhance their position in the evolving power market. R&D and technology improvements for small gas turbines have focused primarily on the testing and integration of new components including recuperators, advanced hot-section materials, and low emission combustion systems. Report Objective This report is intended to serve as a primer on small gas turbine technology, products, manufacturers/suppliers, and business issues affecting the use of gas turbines in power markets. The report focuses on the status of technology, potential performance enhancements, products available from key companies, small gas turbine applications, and market and business issues. The size range investigated for this report is 300 kW to 5 MW. The next four chapters of the report cover: 1. 2. 3. 4. Technology characteristics and developments (Chapter 2) Catalytic combustion systems and developers (Chapter 3) Key companies and providers (Chapter 4) Market and business issues (Chapter 5).

1-4

EPRIsolutions Licensed Material

2
TECHNOLOGY DESCRIPTION
Basics Gas turbines vary greatly in size, from 25-kW microturbines to heavy duty, 250-MW machines in central power applications. Gas turbines serve a wide variety of purposes including electric power generation; aircraft, helicopter, and boat propulsion; water pumping; and other mechanical drive applications such as compressor drives. A gas turbine, also referred to as a combustion turbine, is a continuous flow machine that compresses the working gas (typically air), heats it, and expands it in an expansion turbine to its original ambient pressure. (Turbine can refer to the entire machine or just to the expansion turbine component.) Between the compressor and the expansion turbine, the working gas is heated, raising its temperature enough to increase the systems efficiency and power to economically attractive levels. The usual method of raising the working gas temperature is to burn fuel in the compressed air stream. The fuels used most frequently today are natural gas and several types of liquid fuels. These fuels must be pumped or compressed to higher pressures than the compressor discharge pressure to promote rapid fuel/air mixing in the combustor. Net power is generated when the expansion turbine produces more power than the compressor consumes. This occurs when greater volumes of gas pass through the expansion turbine than through the compressor, due to the higher temperature of the expansion gas. Because gas volume increases with temperature, all other things being equal, gas turbines with higher turbine inlet temperatures have higher efficiency and produce more power per pound of air throughput than gas turbines with lower turbine inlet temperatures. Figure 2-1 shows a cutaway of a typical industrial gas turbine machine. Gas turbines have three key subsections the compressor, combustor, and expander or power turbine. Heat exchangers can be used in several ways to reduce gas turbine fuel consumption and increase power output. For example, fuel consumption can be with recuperators preheating combustion air, while intercooling can reduce compressor power consumption. Gas turbine exhaust is frequently used to raise steam, which can either supply heat or boost the systems power generation capacity. For additional power, the steam can be injected into the expansion turbine or passed through a separate steam turbine in a combined cycle. In large sizes, such combined gas and steam turbine cycles are the highest efficiency power generation systems in commercial service today.

2-1

EPRIsolutions Licensed Material Technology Description

Figure 2-1 Gas Turbine Cutaway Source: Solar Turbines

Types of Gas Turbines Small gas turbines, ranging in size from 300 kW to 5 MW, are frequently categorized as either industrial or aeroderivative. Aeroderivative turbines are engineered for flight services including propulsion via turbojet, turbofan, or turboprop and for helicopter power. Typically, aeroderivative turbines have more high-temperature and high-technology components, such as blades and vanes with complex internal cooling passages, hot-section materials capable of higher temperature operation, and blade and vane profiles that are optimized aerodynamically for maximum performance. These features, along with higher pressure ratios, give aeroderivative engines higher efficiency and power density (kilowatt or horsepower per unit of air flow and machine volume) than their industrial competitors. Because of these design features, aeroderivatives typically cost more than industrial turbines with equal continuous-duty ratings. Because a turbines optimum pressure ratio increases with increasing turbine inlet temperature, pressure ratios of aeroderivative gas turbines are somewhat higher those of industrial types of the same vintage. In the 300 kW to 5 MW size range, aeroderivative turbine pressure ratios range from below 6.5 to above 14. Both aeroderivative and industrial models offer pressure ratios from 6.8 to 12.5; most units of both types have pressure ratios in the 8-10 range. Small gas
2-2

EPRIsolutions Licensed Material Technology Description

turbines on the market today were designed over the past 40 years, as turbine inlet temperatures were advancing appreciably; thus these turbines have pressure ratios that cover a corresponding range. The high-technology components of older aeroderivatives are less advanced and less expensive compared to newer models, due in part due to the dissemination of R&D accomplishments. A new industrial gas turbine could have a higher pressure ratio than an older aeroderivative machine. When comparing gas turbines for stationary applications, many features vary from product to product over the wide range of machine vintages, which complicates generalizations regarding cost and performance. Engine weight per unit of power is the most important variable for aeroderivative turbines, even more important than efficiency, in their primary market applications. Aeroderivative gas turbines enter service in aeronautical applications, after which the manufacturer modifies them for stationary power generation and markets units built around the aeronautical core engine. Aeroderivative turbines also seek small frontal area to minimize drag. For stationary power applications, turbojet and fan-jet engines require fan removal and the addition of a power turbine to drive the generator at proper rotational speed. In contrast, heavy-duty industrial gas turbines are designed for long life, high durability, and low cost. Industrial turbines, which have the longest lives of all gas turbines, have been used since the 1950s and have served as natural gas pipeline compressor drives. Some heavy-duty pipeline compressor turbines have accumulated more than 300,000 operating hours, which would correspond to 10,000,000 miles of service at an average speed of 33 mph. Turbine life and durability in this application are increased by operating continuously at a set point, which minimizes the injurious effects of thermal cycling. Ratings Gas turbines are rated for either continuous duty or emergency service (standby or backup). The different power ratings reflect the number of operating hours before major overhaul. Emergency turbine generators, with their higher turbine inlet temperatures and correspondingly greater power and higher efficiency, are designed for a small number of operating hours between overhauls. Continuous-duty ratings apply to operations with longer overhaul intervals and lower power and efficiency. Ratings correspond to differences in the allowable combustor and turbine inlet temperatures, reflecting the deleterious effects of high temperature on the creep of turbine blades, thermal distortion of the combustor liner, and degradation of other hot-section parts. Ratings for gas turbines assume certain ambient conditions that users should be aware of. Standard ISO rating conditions are 59oF (15oC), sea-level elevation, 50% relative humidity, and no inlet or exhaust pressure losses. Higher ambient temperature or elevation, for example, will result in derating. Fuels Both gaseous and liquid fuels are burned in gas turbines today. Gaseous fuels include natural gas, propane, landfill gas, several types of refinery off-gas, and gasified solid fuels of several compositions, while liquid fuels include kerosene, jet fuel, and No. 2 distillate oil. Nuclear heat sources have been considered, as well as solar concentrators for both terrestrial and space applications. Externally fired primary heaters fueled by the combustion of coal, biomass, and
2-3

EPRIsolutions Licensed Material Technology Description

other fuels (which usually contain ash, sulfur, alkalis, and other impurities) have also been used historically for research or demonstration purposes. Many gas turbines have dual-fuel capabilities. Some can switch fuels after a brief shutdown for changeover of combustor nozzles, while others require that combustor parts be replaced to operate on a second fuel. Some turbines have two sets of nozzles in the combustor and can switch fuels while operating. Not all of these dual-fuel features are available on combustors designed for low NOx emissions, in particular with lean premixed designs. Industrial gas turbines of commercial interest for distributed generation applications are expected to burn only clean fuels including natural gas, various clean petroleum liquids, and purified alternative gaseous fuels. Gas turbines require clean fuels to avoid the deleterious effects of corrosion and deposits. Combinations of sulfur and alkalis (sodium and potassium) in the fuel create sodium and potassium sulfate, which chemically attack (corrode) the gas turbine blades at typical surface operating temperatures and form deposits on the blades. Gas turbine fuel specifications now limit the permissible concentration of these chemicals in the fuel and instruct operators to avoid sites and inlet configurations that result in the intake of salt and dust, which contribute to corrosion and deposits. On an experimental basis, solid fuels such as coal have been burned in gas turbines for brief periods. Even with the best available ash separation systems, sufficient ash entered the turbine that rapid erosion of the blade profile occurred. Large gas turbines have been operated successfully for extended periods of time on gasified coal after the fuel is cleaned of particulate and gaseous contaminants. Special combustors have been developed to burn gasified coal that has been mechanically and chemically cleaned. This technology is applicable in an integrated gasification-combined cycle (IGCC) plant. At present, however, natural gas fueled combined cycles can produce lower cost electric power than coal gasification-combined cycle plants. Gas turbines of all sizes can use gasified coal technology to burn dilute fuel gas with heat contents as low as 90-100 Btu/SCF. Fuel Pressure Requirements The combustors of gas turbines operate at elevated pressures, typically 75-200 psig (517-1379 kPa) or higher depending on machine pressure ratio. The fuel must be at pressures adequately above combustor pressure for it to flow into the machine through control valves, enter the combustor through nozzles, and retain enough momentum to mix properly with the combustor air. This typically requires fuel pressures of 120-350 psig (827-2410 kPa). Table 2-1 lists representative examples of small gas turbines and booster compressors. These machines are typically twin-screw, reciprocating, or centrifugal compressors. Figure 2-2 shows a typical booster compressor skid, based on a Kobelco twin-screw machine. Pressure is not a problem with liquid fuels, because available pumps have the appropriate capacity and pressure capability. However, gas compressors for this application (low flow and somewhat high pressure) are not in high demand. Typically they must be made to order and thus are not inexpensive.

2-4

EPRIsolutions Licensed Material Technology Description Table 2-1 Example Gas Turbine Booster Compressors Source: GTI* Compressor Manufacturer Compressor Outlet Fuel Rate Type Pressure (scfm) (psig) Centrifugal Twin Screw Recip & Centrifugal 360 120 200 625 360-460 490 Prime Mover and Type Prime Mover Size

Gardner Denver Sullair, others General Electric (Nuovo Pignone) Mycom/ Frick Co. Mycom/ Frick Co. Mycom Mycom

Orenda Dresser-Rand General Electric (Nuovo Pignone) Kawasaki

turbine 2500 kW turbine 1450kW/1850 kW turbine PGT-2 - 2000 kW

Twin Screw 200 - 220 331 511

turbine Model M1A 1235 - 2065 kW turbine Model M1T 2950 - 4022 kW turbine Saturn - 1210 kW turbine Centaur 3515-4600 kW turbine Taurus 52006890 kW

Twin Screw 200 - 220 768 - 1,039

Kawasaki

Twin Screw 140 - 180 Twin Screw 205 - 300

305 780-975

Solar Turbines Solar Turbines

Mycom

Twin Screw 205 300 1060-1315

Solar Turbines

* Data are representative examples. Gas turbine suppliers work with packagers and customers to accommodate their preferences on booster compressor suppliers.

Figure 2-2 Gas Turbine Booster Compressor Skid Source: Kobelco

2-5

EPRIsolutions Licensed Material Technology Description

Gas turbine packagers sometimes install two fuel gas compressors to ensure system reliability is not compromised by fuel compressor failure or associated long delivery time for replacement. When high-pressure fuel gas is available, it is used directly. In the broad distributed generation market, however, power generation engineers cannot assume the availability of high-pressure fuel at sites with power demand in the 300-kW to 5-MW range. The cost and reliability considerations of the fuel gas booster compressor might need to be considered in project economic and performance analyses. Operational Characteristics In the small gas turbine size range, most products have efficiencies ranging from 25% to 40%, with a corresponding range in heat rate of 13,650-8,530 Btu/kWh (LHV). Note that reciprocating engine and gas turbine efficiency is rated on an LHV basis, whereas natural gas is sold on a higher heating value (HHV) basis. The difference between these fuel heat content ratings is the heat of condensation of the water vapor in the combustion products. This heat is unavailable for power generation in reciprocating engine and gas turbine machines, hence its exclusion in the machine rating. Economic analyses should distinguish between LHV and HHV in fuel price and engine performance. Factors for converting HHV to LHV vary slightly with fuel composition. Typical values are: Natural gas: HHV = 1.106 x LHV Diesel fuel: HHV = 1.06 x LHV

In addition to the effects of derating caused by high ambient temperatures or other factors (see the previous section, Ratings), users need to account for pressure losses and the impact of parasitic system loads such as pumps, fans, and compressors. The exhaust of industrial gas turbines typically ranges from 932 to 1112F (500 to 600C), enabling steam to be generated at medium pressures. The exhaust of recuperated turbines (see Recuperated Cycle below) and aeroderivative gas turbines is lower in temperature due, respectively, to the use of recovered heat for preheating combustion air and to higher pressure ratios. When generating steam, these lower exhaust temperatures result in a somewhat lower amount of heat recovered and a lower heat recovery efficiency. During periods of peak power demand, operators may choose to overfire a gas turbine (fire at slightly higher temperature than continuous-duty design point) to generate more power and improve efficiency. However, overfiring reduces the interval between scheduled maintenance. The tradeoff is between the higher energy prices being paid during peaking periods and the cost of shortening the maintenance interval. Gas turbine power can also be increased by inlet air cooling. Lower compressor inlet temperature means that less power is needed by the compressor, resulting in greater output and higher efficiency. Gas turbine components, including the electric generator, are usually sized for operation in the expected service temperature range, so the extent of power augmentation has a practical upper limit because the generator output cannot safely exceed its rating. As mentioned earlier, the standard temperature for rating gas turbine output is 59F (15C). Gas turbine output and efficiency are quite sensitive to compressor inlet temperature, so appreciable derate occurs when the ambient temperature rises. During peaking periods, increased output is often valued
2-6

EPRIsolutions Licensed Material Technology Description

higher than increased efficiency. As air-conditioning loads increase during hot weather, gas turbines and other peaking equipment are called upon to meet the load. At these times, the value of electricity is greatest, so it often pays to augment turbine output by cooling inlet air. Inlet air cooling can be accomplished either by a heat exchanger and a cooled fluid or by evaporative cooling. Using a heat exchanger can be expensive, in both capital and operating cost, because it requires a source of cooling, typically a chiller (electric, turbine exhaust-driven absorption, or natural gas-fired absorption). Water at a temperature appreciably cooler than air can also be obtained by operating a small cooling tower at night, or whenever the ambient temperature is adequately low, and storing the cooled water for later use during hot-weather peak demand. Inlet air evaporative cooling uses an array of spray nozzles in front of the air inlet to create a mist in the flow before it enters the compressor. The spray system should not leave sizable drops in the air entering the compressor, or compressor blade erosion can result. Adequately demineralized water must be used to avoid deposits in the gas turbine that, over time, would reduce its output power and efficiency. Thermodynamic Cycles The thermodynamic cycle on which the gas turbine operates is called the Brayton cycle. Several variations in the Brayton cycle are described below. The basic cycle, referred to as the simple cycle, consists of compression, temperature rise by heat addition, and expansion. Variations in the cycle involve adding components for efficiency improvement and power increase. In various compound cycles, the gas turbine exhaust is used as heat input to a second cycle, often referred to as the bottoming cycle. In such cases the gas turbine becomes the topping cycle. Simple Cycle The simple cycle consists of a compressor, combustor, and expansion turbine. This configuration is most commonly known as a gas turbine. In a jet engine, the gas turbine exhaust is discharged at high velocity by an exhaust nozzle. Simple cycles may consist of one or more rotating shafts with expansion turbines and compressors on them. Jet engines are made into stationary power turbines by adding either a generator and an additional turbine stage to the shaft or adding a second shaft with an expansion power turbine and a generator on it. The simple cycle is shown schematically in Figure 2-3.

Figure 2-3 Schematic of Simple Cycle

2-7

EPRIsolutions Licensed Material Technology Description

Recuperated Cycle In the recuperated cycle (Figure 2-4), turbine efficiency is raised by adding a recuperative heat exchanger, which uses the hot exhaust gas of the expansion turbine to preheat the air flowing into the combustor, thereby reducing the fuel required. This cycle is also sometimes referred to as a regenerated cycle. There is no difference between these two designations from a thermodynamic viewpoint. A recuperator is a heat exchanger with passage walls through which heat flows by virtue of the temperature difference between the two fluids on either side of the wall. The fluids in a recuperator do not mix at all. A regenerator is a periodic heat exchanger in which hot and cold gas flow alternately in opposite directions through a matrix of fine passages. In a regenerator, the two fluids mix to a small degree, and leakage can occur from the high-

pressure, compressor discharge side to the low-pressure, expansion turbine exhaust side.

Figure 2-4 Schematic of Recuperated Cycle

Conventional (solid boundary) recuperative heat exchangers are used most frequently in heating and air-conditioning applications and for industrial heating. Periodic (rotary wheel) regenerative heat exchangers have been tested since the 1950s for use on automotive gas turbines. Regenerators have been researched because they could be compact enough for the gas turbine to fit under the hood of a car. However, the high-pressure seals required in the regenerator have not yet achieved adequate life for this application. The recuperated turbine cycle produces about 10% less power than a simple cycle of the same compressor pressure ratio and turbine inlet temperature. This is because an inherent pressure drop is associated with the recuperator and with its connections to the engine and gas turbine exhaust. The design of a practical recuperated cycle involves balancing the tradeoffs among the parameters of efficiency, power, and cost. This is accomplished by analyzing various heat exchanger sizes, dimensions, and configurations to obtain a desired level of pressure drop on each side of the recuperator and interconnecting ducting, as well as analyzing recuperator cost. Similar tradeoffs apply to the regenerative cycle. Intercooled Cycle The intercooled cycle (Figure 2-5) splits the compressor into two or more sections and uses externally cooled heat exchangers to cool the air flowing between the sections. Intercoolers are in widespread use for boosting the power of turbocharged diesel and spark-ignited reciprocating engines. Intercooling reduces compressor work by lowering the temperature of the air in later
2-8

EPRIsolutions Licensed Material Technology Description

stages of the compressor, thereby enabling them to act on a cooler gas and reducing power consumed in the compressor. This increases net machine output and results in a higher power density.

Figure 2-5 Schematic of Intercooled Cycle

However, the compressor discharge temperature is lower with an intercooled cycle, so additional fuel must be burned in the combustor to reach the expansion turbine inlet temperature. In spite of its added power, the intercooled cycles efficiency is basically the same as that of a simple cycle with the same pressure ratio, due to the increased fuel required and the losses associated with the intercooler pressure drop. Intercooled cycle design involves pressure drop tradeoffs similar to those of recuperated cycles (see Intercooled Cycle Design). Intercooled Recuperated Cycle In the intercooled recuperated cycle (Figure 2-6), both a recuperator and an intercooler are used, achieving the high efficiency advantage of the recuperated cycle and the high power of the intercooled cycle. The U.S. Navy is currently sponsoring development of an intercooled, recuperated gas turbine with higher fuel efficiency and turbine power. Tradeoffs exist involving heat exchanger pressure drop, with the increases in cycle performance being balanced against the losses caused by the heat exchanger pressure drop and corresponding sacrifices in power and efficiency. In the range near optimum conditions, these sacrifices are typically quite small.

Figure 2-6 Schematic of Intercooled Recuperated Cycle


2-9

EPRIsolutions Licensed Material Technology Description

Figures 2-7 and 2-8 show the efficiency and specific power (another term for power density) for the four cycles of concern: simple, recuperated, intercooled, and intercooled and recuperated. The data assume typical component performance (turbine efficiency of 88%, compressor efficiency of 83%, recuperator and intercooler heat exchanger effectiveness of 85%, and recuperator and intercooler pressure drop of 2% on each side) and a turbine inlet temperature of 2100F (1150C), typical of new gas turbines in the 3-5 MW size range. The graphs cover the pressure ranges of interest to the point where the cycle reaches a diminishing return or crossover point.

Figure 2-7 Efficiency of Selected Cycles

Figure 2-8 Specific Power of Selected Cycles

The gains in efficiency found with recuperators are apparent for both the recuperated cycle and the intercooled and recuperated cycle, as are the gains in specific power with intercooling alone
2-10

EPRIsolutions Licensed Material Technology Description

and with intercooling and recuperation. The slightly diminishing advantage at very low and very high pressure ratios is due to the compromising effects of the heat exchanger pressure drops. The efficiency and specific power of the simple cycle rise with pressure ratio. Increase in efficiency diminishes at pressure ratios above 12, which is the point of maximum specific power. The efficiency of the recuperated cycle reaches a maximum at a pressure ratio of 5; however, the specific power continues to increase up to a pressure ratio of 10. At pressure ratios between 5 and 10, only a small drop in efficiency occurs, so a practical machine would balance the benefits of efficiency and power by operating in the 5-10 range, perhaps at a pressure ratio where the turbine exhaust temperature becomes low enough to allow the use of a lower grade stainless steel in the recuperator. At low and moderate pressure ratios, the intercooled cycles efficiency is not up to that of the simple cycle because of losses due to pressure drop in the intercooler. However, the specific power of the intercooled cycle is head and shoulders above that of the simple cycle for all pressure ratios. The intercooled, recuperated cycle has the best of both improvements high efficiency and high specific power once the pressure ratio is high enough to overcome the consequences of heat exchanger pressure drop. Figures 2-9 and 2-10 show the efficiency and specific power of these cycles as a function of turbine inlet temperature for pressure ratios that represent a balance between high efficiency and high specific power (12-17 for the simple cycle, 8-12 for recuperated, 20-30 for intercooled, and 15-30 for intercooled, recuperated. As expected, these pressure ratios increased with turbine inlet temperature.

Figure 2-9 Locus of Efficiency for Optimized Cycles

2-11

EPRIsolutions Licensed Material Technology Description

Figure 2-10 Locus of Specific Power for Optimized Cycles

Design Considerations Recuperated Cycle Design Recuperators are heat exchangers used for preheating the compressor discharge air to reduce fuel consumption in the combustor, thereby increasing efficiency. The technology, economics, status, and industrial suppliers of gas turbine recuperators are reported in Gas Turbine Recuperators: Benefits and Status (EPRI TR-113745, GRI-99/0263, December 1999). Practical considerations in the design of recuperated gas turbines include flow ducting, diffusion of flow to low velocity, flow turning, and viscous pressure drop in heat exchanger cores. Recuperators by themselves reduce net power due to the pressure drop incurred in the heat exchanger. The pressure drop in the recuperator core is related to the heat exchanger effectiveness in a fundamental manner.* Recuperated cycle efficiency is highest when the gas turbine pressure ratio is low, resulting in high temperature differences between the expansion turbine exhaust and the compressor discharge. This enables the recuperator to have a high fuel savings and high heat exchanger effectiveness with somewhat modest surface area (and cost and pressure drop).
* Readers interested in the relationship between heat transfer and pressure drop in heat exchanger passages are referred to the sections on Reynolds analogy in standard college texts on heat transfer. The Prandtl Number (diffusivity of momentum divided by the diffusivity of heat), cp/k for air and all simple gases, is near unity (0.7 in the case of air), so that experimental confirmation of Reynolds analogy in recuperators is not surprising. To reduce pressure drop losses, the intercooler should be operated at low velocity (actually low Mach Number) which, however, adds cost and size to the gas turbine and incurs additional pressure drop in the diffuser. The optimum velocity in the intercooler involves consideration of cost and performance, which are application-specific.

2-12

EPRIsolutions Licensed Material Technology Description

Microturbines and the Solar Turbines new Mercury 50 are the only gas turbines currently in production that are designed with recuperators and intended for service with high annual capacity factors. The TF-1500 recuperated gas turbine, which serves as the main engine on the M1A1 Abrams tank, is an expensive engine whose recuperator (and resulting fuel economy) was justified by the limited space for fuel onboard a battle tank and by the need to operate under extreme military service conditions. Many natural gas pipeline compressor stations employ recuperators, which were retrofits on older, low pressure ratio gas turbines. Pipeline compressor stations typically have high annual capacity factors on the most efficient units, so the economics of recuperation works out favorably. Recuperators in pipeline compressor service are known to develop problems with leaks, believed to be associated with differential thermal expansion accompanying thermal cycling. Whether new recuperators will develop leaks after experiencing significant thermal cycling is not known yet, due to their limited experience in the field. Recuperator manufacturers, learning from prior experience, have designed and manufactured their products for cycling duty. However, durability under repeated cyclic operation can be determined only on the basis of experience, because it is largely affected by manufacturing quality control and details (especially the welding or brazing of metal joints). The long-term durability and performance of microturbines and Mercury 50 units are of interest to intermediate-duty gas turbine suppliers who want to offer reliable, high-performance, low-cost products. Intercooled Cycle Design Typically two-thirds of the gross power produced by gas turbines is used to drive the compressor. Consequently, reducing compressor power consumption can significantly increase net power production and efficiency. Thermodynamically, intercooling appears very attractive for increasing net power by cooling the air flowing in the middle of the compressor and thereby reducing the work absorbed in later stages of compression. The optimum pressure level for intercooling an ideal gas turbine is at the square root of the pressure ratio, essentially halfway through the compressor. This balances the advantage of having a substantial amount of heat to remove (which was put into the air by the first part of the compression process) and having a substantial extent of compression work remaining to be done in the second half of the compressor. Detailed studies are required for actual machines in specific cases. Intercooling becomes increasingly attractive as the overall pressure ratio increases. Intercooling by itself has a substantial effect on increasing net power from the machine, but only a modest effect on efficiency, because the lower temperature air flowing into the combustor requires additional fuel to deliver the combustion products at design temperature. Also, cycle losses are incurred by pressure drop in the heat exchanger and connections to the cycle.* A price in terms of performance and economy must be paid for the power gained by intercooling. Intercooling heat exchangers are of appreciable size, especially ones designed for ambient and low-pressure air. Consequently the intercooler, with one side handling air at modest pressure and the other side often handling ambient air, is large compared to the compressor. For the air being cooled to flow through the intercooler with reasonable pressure drop, it has to be taken off the
* Refer to previous footnote, page 2-12.

2-13

EPRIsolutions Licensed Material Technology Description

compressor and diffused to low velocity, but diffusers create pressure loss. The flow passages remove air from the compressor and then return it, which changes the direction of flow and creates pressure loss. The intercooling heat exchanger itself incurs pressure drop, because the heat transfer and wall shear forces are coupled by the diffusion of heat and momentum from the wall into the air stream. Intercooling can also be accomplished by spray cooling with water at an intermediate condition in the compressor. Injected water must be fully evaporated to avoid carryover of drops into later sections of the compressor. Droplet carryover could damage compressor blades upon impact. Also, the spray water cannot contain substantial mineral matter, which could result in deposits in either the compressor or the expansion turbine, either of which would reduce output power and efficiency. General Electric (GE) offers spray intercooling (Sprint) on its LM 6000, which increases output power by 20%. The LM 6000 is a two-shaft aeroderivative gas turbine with adequate space for a spray intercooler between the low-pressure and high-pressure compressors. The spray is controlled to very fine droplet size so that most of the evaporation occurs between the compressor stages. Large droplets lose enough mass that the risk of carryover and blade erosion is acceptable. Independent engineers have added inlet spray cooling with chilled water. Some systems chill the water during peak periods for power augmentation, using ice that was made and harvested during off-peak hours. The chilled-water spray cooling technique is used on other large, new gas turbines, such as the W 501 B5 and the GE Frame 7EA. Inlet air cooling and spray intercoolers are new technologies being applied first to large machines where the payoff is highest, but they could be applied to small gas turbines as well. The same care must be taken to avoid large, massive droplets (over 20 microns) and excessive mineral matter in the water, which would deposit in the compressor and turbine sections. With these engineering precautions, the power output of small gas turbines can be increased on hot days with inlet air cooling and on a more routine basis with wet compression. Water quality for inlet air cooling and wet compression are based on manufacturers guidelines for the total amount of sodium plus potassium, lead, vanadium, calcium, and total dissolved and undissolved solids permissible for introduction into the gas turbine, whether they are introduced by the water or the fuel. To increase power with little increase in net engine cost, intercoolers are frequently used commercially on turbocharged reciprocating engines in off-highway vehicles, boats, trucks, and automobiles. However, the only commercial gas turbine with an intercooler was the Stahl-Laval GT120, which is not in production anymore. This machine, made in Sweden during the 1950s60s by a predecessor of ABB, took advantage of the low-temperature cooling water available in Swedish lakes and rivers. Its unusual configuration consisted of three shafts. The machine had a split compressor, with the low-pressure compressor on one end of the machine and the highpressure compressor on the other end, leaving plenty of space for the diffuser, interconnecting ductwork, and intercooler. The added cost of the intercooler and turbine design was presumably justified by the systems economic benefits, which resulted from the turbines intermediate-duty application (high number of operating hours per year) and from high oil prices in interior Sweden at the time.
2-14

EPRIsolutions Licensed Material Technology Description

Most likely, intercooling will be applied first on gas turbines with high pressure ratios, because the benefits of intercooling increase with increasing pressure ratio. However, gas turbine pressure ratio must not be so high that, despite the lowering of the compressor discharge temperature by intercooling, the turbine exhaust temperature is not above the compressor discharge temperature. Studies of advanced cycles by three international manufacturers of large aeroderivative gas turbines were performed as part of the Collaborative Advanced Gas Turbine program in the early 1990s, sponsored by a group of international electric and gas utilities (see Recent Developments in Advanced Cycles). These studies examined the use of new, hightechnology, high-performance, large aeroderivative gas turbines in combined cycles and with intercooling, as possibilities for developing new products for the commercial power generation sector. The program focused on an intercooled aeroderivative product, which incorporated an intercooler and appropriate gas turbine modifications to result in a moderate-cost product of medium efficiency (not quite as high as a combined cycle). However, development never began because the manufacturers did not see an adequate market for the product to justify their investment. Combined Intercooled and Recuperated Cycle Design The combination of intercooling and recuperation is not used on commercial stationary gas turbines at the present time. No gas turbine has yet been developed for an application requiring the performance features of such a machine. Currently, the power generation business is focused on two applications, baseload and peaking power. The industry is interested in using gas turbines for intermediate-duty applications, but few new products (other than the Mercury 50 and microturbines) have been developed specifically for the distributed generation market. Highefficiency, recuperated and intercooled industrial gas turbines that are attractively priced could provide power to intermediate-duty customers at lower cost than they are currently paying to regulated distribution companies and could also reduce greenhouse emissions appreciably. One project is underway to develop an intercooled, recuperated gas turbine. This research is being conducted for the U.S. Navy (with some European involvement and interest) by Northrop Grumman Marine Systems (Sunnyvale, CA). The turbine would serve as the main propulsion engine on ships requiring about 30,000 hp of shaftpower. In power generation service this engine, the WR-21, would produce 24.4 MW at its continuous-duty ISO operating condition. The Navy is interested in fuel conservation to extend the ships range with limited onboard fuel storage. Fuel cost savings and extended fuel supply during periods of international crises provide incentive for applying the intercooled, recuperated cycle. The systems high efficiency offers the major advantage of reducing fuel consumption while cruising at part power, yet retaining the ability to go to high power when necessary. The poor efficiency of conventional gas turbines at part power caused the U.S. Navy to seek a better gas turbine cycle for performing longer range missions without refueling. Figure 2-11 compares the part-load efficiency of a simple-cycle gas turbine, a diesel engine, and an intercooled, recuperated gas turbine. As shown, little efficiency is sacrificed by the intercooled, recuperated turbine at part load in contrast to the efficiency loss of a simple cycle. The intercooled, recuperated gas turbine is smaller than a diesel engine of comparable power. The cost, durability, and commercial fate of this gas turbine are unknown at this time.

2-15

EPRIsolutions Licensed Material Technology Description

Figure 2-11 Comparison of Part-Load Efficiency of Diesel Engines and Gas Turbines

Size Considerations Gas Turbine Systems Historically, both industrial and aeroderivative gas turbines have increased in size from the small units built in the 1950s-60s to the larger ones of recent vintage. Several small gas turbines are old models that have recently found a market niche. As sales of gas turbines increased over the years, manufacturers had cash to develop new products, and larger units were built to compete favorably with diesel engines. The newer, larger turbines employ advanced technology to reach higher efficiency and specific power and lower cost per kilowatt. The progression from small to larger sizes illustrates a combination of three phenomena: Increasing turbine inlet temperature (via improved high-temperature materials and the introduction of internal blade and vane cooling) with corresponding increases in power and efficiency Increasing compressor and turbine component efficiencies (due to research on the flow of air in turbomachinery passages) over time Increasing turbine efficiency as a consequence of an increase in physical size.

This latter improvement is due to a more favorable ratio of volume to surface phenomena, such as decreased percentage of blade tip leakage, the decreasing effects of Reynolds number-related phenomena such as wall frictional pressure drop, and the ability to produce blade and vane profiles with less aerodynamic flow losses in larger sizes.

2-16

EPRIsolutions Licensed Material Technology Description

Gas Turbine Component Selection Very small gas turbines, typically below 250 kW in size, use centrifugal (radial-outflow) compressors and radial-inflow turbines, while larger sized gas turbines, typically over 2 MW, use axial-flow compressors and turbines. These designs are most efficient at their respective sizes for a number of technical and economic reasons. At low volumetric flow rates, the blade heights of axial-flow compressors and turbines become very small, incurring large viscous surface pressure loss, proportionately higher tip leakage, and other size-related phenomena affecting compressor efficiency. Additionally, the relative cost of a single-stage radial-flow compressor or turbine is much less than the cost of a multi-staged, multi-bladed, axial-flow compressor or turbine of comparable overall pressure ratio. In the transition size range, from 250 kW to 2 MW, the performance differences of these components compete, and designers must consider numerous factors including manufacturing and development cost, output shaft speed, and the cost and weight of gearboxes for turboshaft and helicopter engine applications. Component selection is also affected by experience with specific components and by part availability. In this transition size range, centrifugal compressors are often combined with axial-flow expansion turbines, which can then operate with higher pressure ratios per stage than possible with axial-flow compressors. Other combinations, such as a few stages of axial-flow compression followed by a centrifugal or mixed-flow stage, have also been used in this size range. Commercial users should focus mainly on the cost, efficiency, and durability required of a turbine in their application. Gas turbine flow path design tradeoffs are complex, and manufacturers evaluate these tradeoffs based on long-term, broad market considerations, not on sales of a few units. Numerous texts exist on the topic of turbomachinery and turbine system design. Gas turbines have an optimum pressure ratio for maximum efficiency and another optimum pressure ratio for maximum specific power. The specific power maximum occurs at a higher pressure ratio than that for maximum efficiency. These optimum pressure ratios increase with increasing turbine inlet temperature, which, in terms of gas turbine products on the market, typically increases over time as technology advances. Figure 2-12 shows efficiency as a function of machine power rating for commercially available gas turbines in the 300 kW to 5 MW size range. Most of the plot points fall in a broad, central band sweeping upward and to the right, to increased efficiency with increased size. Points above the band are modern aeroderivative engines (signified by an A) and the recently introduced recuperated industrial gas turbine (R). Points below the band are low-cost, typically older, gas turbines made for emergency power, where efficiency is not a major product selection criteria. For each gas turbine presented in the figure, the manufacturer presumably has optimized the pressure ratio for the turbine inlet temperature and business application.

2-17

EPRIsolutions Licensed Material Technology Description

Figure 2-12 Efficiency of Small Gas Turbines Source: Turbomachinery International Handbook

Users concerned with the most advantageous gas turbine for their application should focus on their particular economics. Generally, higher efficiency equipment costs more to manufacture and carries a premium for the increased benefits to the user. In selecting a small gas turbine, a very important parameter is the annual full power capacity factor. As the power generation machine gets more use per year, it accumulates more hours over which to spread its fixed carrying charges and is better able to pay a higher price for a higher efficiency. Example of a Tradeoff As a simple tradeoff, a decrease in heat rate of 1000 Btu/kWh results in the same electricity cost reduction as a decrease of $112.50/kW of gas turbine cost. This calculation assumes a fuel cost of $4.50/million Btu (MMBtu), 5000 hr/yr of operation, and a capital charge factor of 20% (combined interest, taxes, insurance, and depreciation). This 1000 Btu/kWh heat rate decrease, or increase in efficiency from 28.4% (12,000 Btu/kWh) to 31.0% (11,000 Btu/kWh), and the corresponding variation in gas turbine cost straddle the price and performance range of many competitive products. The formula for this tradeoff is: dC = d(HR) x (FP) x (D)/i Where d = differential operator C = Cost ($/kW) HR = heat rate (Btu/kWh)
2-18

EPRIsolutions Licensed Material Technology Description

FP = fuel price ($/Btu) D = annual duration of use (hr/yr) i = capital charge factor (as a decimal, not as a percentage). Fuel prices are usually quoted in $/MMBtu, so care must be taken to convert the fuel price to $/Btu, or the results will be off by a factor of one million. The Use of Steam to Increase Efficiency Steam Injection Gas turbine exhaust heat can be used to generate steam in a boiler, often called a heat recovery steam generator (HRSG). The steam can be injected into the combustor or turbine, generating additional power by increasing mass flow through the turbine. However, steam injection into the combustor requires additional fuel to maintain turbine inlet temperature. Using the unfired gas turbine exhaust heat source, larger amounts of steam can be raised in a humidifier than in an HRSG, which is more constrained by the heat transfer pinch point consideration. In humidification, compressed air and hot, pressurized water are passed in opposite directions in a combined heat and mass exchanger called a humidification tower (see Recent Developments in Advanced Cycles). Using such large amounts of steam, however, is not possible in existing gas turbines because they are designed for approximately equal mass flows in the compressor and expansion turbine sections. Such a turbine could be built, with much larger mass flow in the expansion turbine than through the compressor, but to date, manufacturers have not invested in it because it lacks the power generation industrys interest. Combined Cycles Combined cycles use the turbine exhaust to raise steam in a separate boiler, or HRSG. Additional power is then produced by a steam turbine. More complex cycles, such as steam addition by adiabatic saturation (humidification) and reaction of fuel and steam in a catalytic heat exchanger (chemical recuperator) have been considered in analytical studies but have not resulted in commercial development. Due to their complexity and cost, chemical recuperators and humidifiers are generally considered by gas turbine manufacturer and power generation business executives to be candidates first for large, central station systems before their complexity and cost is reduced enough for them to be economically advantageous in small gas turbines for distributed generation and industrial cogeneration. Today the vast majority of new baseload power plants use the combined cycle. With its very high efficiency and modest cost, it is the most economic system for large units. Combined-cycle plants are being installed in great numbers in the United States, Europe, and Japan not only for economic reasons, but also for their environmental advantages (extremely low emissions of NOx, carbon monoxide [CO], and carbon dioxide [CO2] per MWh generated). These turbine plants are expected to help industrial nations to meet the Kyoto protocol regarding CO2 emissions. Switching from coal-fired steam to natural gas-fueled combined-cycle plants reduces CO2 emissions by more than a factor of two.
2-19

EPRIsolutions Licensed Material Technology Description

Combined-cycle plants use large, modest pressure ratio gas turbines (the same as those used for peaking applications, only as continuous-duty machines), with the exhaust ducted into HRSGs, which provide steam to a separate steam turbine. The turbines modest pressure ratio results in a high enough gas turbine exhaust temperature that the steam is generated at high enough pressure and temperature to yield very low-cost electricity. High pressure ratio, aeroderivative gas turbines have been considered for combined-cycle applications, but because of their substantially lower gas turbine exhaust temperature, they generate less steam at lower pressures, with lower overall combined-cycle efficiency. Simply put, in the temperature range of 1050F (570C) down to 750F (400C), the Rankine (steam) cycle is more efficient in generating electricity than is the Brayton (gas turbine) cycle. Steam cycles cannot economically utilize cycle temperatures over 1050F (570C), while the gas turbine can, so each is preferred in its own most advantageous temperature range. High-Temperature Materials and Component Engineering Both power output and efficiency increase with turbine inlet temperature. In the early days of gas turbines (1950s and 60s), much effort was directed at increasing turbine inlet temperature. Advances included materials with high strength and creep resistance at increasing temperatures and the development of internal cooling of the turbine blades and vanes. Internal cooling technology permits higher gas temperatures in the expansion process while keeping the blade material below the levels dictated by its metallurgy. Performance increases have been obtained with compressor blade coatings that provide a smooth surface, reducing wall shear forces and pressure drop. These coatings prevent or delay fouling of the compressor blades and losses in compressor and gas turbine efficiency. Turbine blade coatings for protection against metal degradation and overheating were developed and are in use on blades of newer stationary gas turbines. High-temperature oxidation and corrosion also limit turbine durability. Fuel chemical purity specifications regarding fuel sulfur and alkalis have been getting tighter as more is learned about the details of oxidation, diffusion in metals, sulfidation, and grain boundary phenomena. Alloys used for gas turbine blades and vanes include IN 100, Udimet 500, 700 and 710, INCO 713 and 713 LC (low carbon), and Waspaloy. These materials, in one form or another, have been around for over 30 years. The U.S. Department of Defense (DOD) and jet engine developers have been pushing materials temperature capabilities higher and higher. Extensive research has been sponsored by the Air Force and the Navy for military jet engines, as well as decades of proprietary development by gas turbine manufacturers and blade and vane suppliers. Gas turbine users should be aware of the extent to which they should specify low sulfur and low alkali content in their fuel purchases and the need to select sites and filter inlet air to reduce salt and dust ingestion into the gas turbine. Recent progress in high-temperature capability has been made in the field of controlled solidification of gas turbine blades, using current metallurgy and blade coatings. Directionally solidified alloys and single crystal blades, each produced by a different controlled solidification technique, are obtaining increased creep strength at slightly higher temperatures from available materials. Controlled solidification avoids the weaknesses that exist at grain boundaries due to the polycrystalline nature of conventionally produced materials. Note that such high-temperature blades become more expensive as the technology for their manufacture becomes more complex
2-20

EPRIsolutions Licensed Material Technology Description

and sophisticated. Directionally solidified and single crystal blades appear to be too expensive for todays industrial gas turbines, because manufacturers are aware of the technology but have not yet offered such products. Thermal barrier and corrosion protection coatings have been developed for turbine blades and vanes. Thermal barrier coatings provide a thin insulating layer on top of the blade and vane surfaces, permitting higher temperatures in the gas stream without exceeding the temperature limitations of the blade material itself. Thermal barrier coatings are most effective with the combination of high firing temperatures and internally cooled blades, because the coatings reduce temperature more effectively in the underlying material when high heat fluxes occur through the blade surfaces. Corrosion-resistant coatings provide a ceramic layer on the turbine blade, which drastically reduces the rate at which alkali metal sulfates and other corrosive species attack blades. Ceramic coatings must be kept thin, however, because of their much lower thermal expansion than blade metal. If made too thick, the differential thermal expansion can cause the coating to flake off. Turbine Blade Cooling The high-temperature sections of current aeroderivative gas turbines, as well as medium to large stationary gas turbines, use internal cooling. Compressor discharge air is passed through the base of the blade and through the core of the active portion of the blade to cool the blade material. The cooling air flows through carefully engineered internal passages inside the blades and is discharged into the exhaust from that stage. These large gas turbines typically use axialflow expansion turbines, whose turbine blades are large and basically straight in the radial direction, which facilitates their manufacture with internal flow passages. High-temperature capability also entails internal cooling of the turbine vanes or nozzles and combustor liners. These cooling flows reduce the amount of air available for power generation via expansion in the turbine section. This reduction in power is the price paid, in terms of performance, for the higher temperature capability. Gas turbine manufacturers are always trying to develop more effective internal heat transfer passages inside the blade core to reduce the consumption of cooling air and increase the flow through the turbine, thereby obtaining more power and higher efficiency. Many small gas turbines use radial-inflow expansion turbines with a more complex internal flowpath. Internal cooling involves casting cooling passages in the blades, some as fine as 0.050 inch in heavy-duty and industrial gas turbines and 0.020 inch in aeroderivative models. Designed surface roughness features are contained inside these coolant passages to augment heat transfer. To date it has remained impractical to use this technology in blades for radial-inflow turbines; however, small, aeroderivative gas turbines exist that have axial-flow expansion turbines rather than radial-flow turbines and have been equipped with cooled blades. The economic tradeoffs in developing internal blade cooling for small gas turbines include higher manufacturing costs and low production volumes, insufficient to recover investment in developing, testing, and tooling cooled blades and accompanying disk and vane cooling. The manufacture of turbine blades, vanes, and disks with internal cooling passages for very small gas turbines with radial-inflow turbines would require new passage configurations, finer holes, and more precise manufacture than is considered to be commercially worthwhile today. The additional manufacturing cost, the limited market at present, the cost competitiveness of the product, and the difficulty in recovering development costs for gas turbines with only modest
2-21

EPRIsolutions Licensed Material Technology Description

levels of sales mitigate against investment in such technology. In contrast, the development cost of technology advances in large machines can be written off against many more kilowatts. These advances, particularly in the turbine hot section, have resulted in lower machine cost per kilowatt delivered and increased machine efficiency. Competing Technology Reciprocating Engines Reciprocating engines compete against gas turbines in the small to intermediate size ranges. Both natural gas-fueled, spark-ignited and diesel oil-fueled, compression heat-ignited engines are found in this market. Reciprocating engine manufacturers typically use the same basic engine block for both fuel types. However, as the natural gas goes into the cylinder as a gas at inlet manifold pressure, it displaces its own volume in air and leaves less air for combustion. Liquid fuel is injected in diesel engines after the inlet valve has closed, so the fuel does not displace combustion air. The reduced air charge per cylinder stroke with gaseous fuel, along with derating due to knock limitation concerns (premature ignition), reduces the power capability of the engine and consequently increases the specific cost ($/kW) of spark-ignited engines over that of diesel engines. Reciprocating engines are available in sizes from under a kilowatt up to about 35 MW. Below about 25 kW, engines are typically used as emergency generators. Intermediate sizes, 25 kW to about 300 kW, are often low-cost, high-volume automobile and truck engines modified for emergency and portable power and for commercial and industrial cogeneration. From about 300 kW to 1 MW, reciprocating engines are typically large versions of truck engines. From 1 to 35 MW, they are uniquely designed stationary engines. The largest bore cylinders are just over 3 feet (1 meter) in diameter. Large ship engines technically are not stationary; however, their design and purchasers selection criteria favor high durability, long life, and high efficiency performance, so they resemble stationary engines more than they resemble vehicular engines. Reciprocating engines operating on the diesel cycle with oil as fuel employ higher compression ratios than can be used in spark-ignited engines, because premature ignition (knock) is not a concern. Diesel cycle engines consequently have somewhat higher efficiencies than sparkignited engines operated at wide-open throttle condition and built with the same engine block and cylinder size. When operating at part load, most natural gas-fueled and other spark-ignited engines have lower efficiency than diesels of the same size, because output is reduced by throttling the air intake rather than by delivering less fuel. Such throttling creates an irreversible pressure drop in the cycle and dissipates available work, which lowers efficiency. Diesel engines typically have greater emissions of NOx and particulates than do spark-ignited engines due to several factors, including the higher pressure and temperature of combustion and their need for rapid burning to avoid emission of unburned hydrocarbons. Diesel emissions are being regulated to a greater degree as time progresses, making siting is more difficult than with spark-ignited natural gas engines. Diesel engine manufacturers have been developing staged combustion to reduce NOx emissions. Industrial gas turbines have to compete with both natural gas and diesel oil-fueled reciprocating engines. In sizes below about 300 kW, reciprocating engines have the advantage of being manufactured in large volumes, with corresponding economies of scale. Their production equipment and tooling is amortized over large numbers of engines, and their spare parts are widely available, on a highly competitive basis, from competing aftermarket manufacturers.
2-22

EPRIsolutions Licensed Material Technology Description

Reciprocating engines also benefit from a broad technology base resulting from their widespread use over the last 100 years. Reciprocating engine repair and maintenance personnel are more readily available than are gas turbine repair and maintenance personnel. These factors present a severely competitive environment for industrial gas turbines. In intermediate sizes, up to a few megawatts, the shared technology and production system continues to benefit reciprocating engines through economies of scale, but at sizes of several megawatts, reciprocating engines become purpose designed and manufactured in modest quantities. This results in higher cost per unit of power and more competitive pricing of reciprocating engines and gas turbines in the 3-15 MW size range. Although larger reciprocating engine are being built today, their market niche appears to be based on the use of low-cost fuel, including residual oil, which gas turbines cannot use without compromising reliability or adding cost for fuel cleanup. Typically intermediate and large gas turbines process more air through a smaller volume of machine and are more compact and lighter weight than reciprocating engines. Also, engines consume oil for cylinder lubrication, while gas turbines do not directly consume oil. Large, marine cathedral diesel engines have separate systems for bearing lubricating oil and piston ring sealing oil, which increases their scheduled maintenance interval. Long-life, platinumtipped spark plugs are now available commercially for spark-ignited engines. In automotive service, these plugs have lives in excess of 100,000 miles, or about 3000 hours of operation. Reciprocating engines, particularly stationary diesel engines, have operated with grossly oversize oil sumps and timed additive injection systems so as to permit long service intervals, possibly as infrequently as twice or even once per year. Both gas turbines and reciprocating engines are getting to the point where maintenance intervals could become dictated by the need to inspect for degradation of parts due to inherent wear and reliability considerations, rather than due to oil consumption, ignition system tune-up, or surface wear. Heat Recovery and Cogeneration The exhaust stream of a gas turbine is the single source of heat for cogeneration purposes, as distinguished from reciprocating engines with two sources of recoverable heat (low-temperature jacket cooling water and high-temperature exhaust). When steam at moderate pressures is the heat recovery medium for cogeneration, gas turbines typically provide about double the recovered heat than that obtained from reciprocating engines. The single source of recovered heat usually makes the heat recovery system installation simpler and less expensive per kilowatt added. In many applications, the value of the recovered heat makes the economics favorable for the installation of onsite power generation. However, very small cogeneration systems frequently use hot water for heat recovery and distribution; the gas turbine exhausts high temperature and high mass flow are not of major value for water heating. Numerous reports have been written on the details, markets, and benefits of cogeneration. Interested readers are encouraged to pursue such knowledge from commercial system vendors and research reports. Noise Gas turbines have different noise characteristics than reciprocating engines. Reciprocating engines create noise when the exhaust valves open with a large pressure differential between the
2-23

EPRIsolutions Licensed Material Technology Description

combustion products remaining in the cylinders and the atmosphere. This sound, like the roar of racing car and airplane engines, requires the use of mufflers, which reduce net power somewhat. Gas turbines do not have anything like the pressure pulses created when exhaust valves open, so they require less muffling than do reciprocating engines to achieve specified levels of noise. Because gas turbine noise is of an aerodynamic nature rather than that of pressure waves, less power is sacrificed in mufflers than is needed to achieve the same end level of quiet. Typical gas turbines produce 85 dB noise at 3 meters, which can be attenuated further at modest cost. Gas turbine noise comes from the whine of blades passing flow openings in both the compressor and turbine. Aircraft jet engine exhaust noise is created by shearing action which exists between exhaust jets and the surrounding (stationary) air. Additional noise is created by the flow in the compressor and turbine as it is periodically disrupted by the passage of bladed wheels. Inlet air filters are usually used to remove dirt and mineral matter, but they also attenuate noise created by the compressor that would otherwise be radiated from the air inlet. Turbine exhaust is frequently pointed upwards to reduce the noise level reaching people nearby. Additionally, turbine exhaust can be muffled when required for use in commercial applications where noise would be objectionable. Recuperators and heat recovery steam generators muffle turbine exhaust, because the sound is broken up while passing through the internal passages of the heat exchange surfaces. Emissions and Environmental Concerns The first gas turbine combustors were designed to obtain flame stability in an extremely compact flame zone. In these diffusion flame combustors, fuel was injected as a jet and mixed with combustion air by turbulent diffusion into the air stream. This worked well, but in some regions the fuel was burning in near-stoichiometric proportions, creating high-temperature zones that produced high levels of NOx. In the atmosphere in the presence of sunlight, NOx and unburned hydrocarbons react to produce ozone, the active irritant in smog. Old, diffusion-flame combustors with high-temperature hot spots are no longer used in most areas of the United States with air quality problems. NOx emissions were first reduced by lowering hot spot temperature and size by controlling the fuel/air mixing and by injecting water or steam into the hot spots. More recently, further reductions were achieved by mixing the fuel and air prior to ignition (lean pre-mixed combustion) so that combustion occurs in zones that are substantially lean with lower adiabatic flame temperature. These newer designs are called lean pre-mixed as well as dry low-NOx. When emissions must reach levels below those readily achievable with existing combustors, gas turbine users often turn to selective catalytic reduction (SCR) for additional NOx reduction. SCR uses a catalyst and ammonia, or a urea-based reagent, to reduce the NOx to molecular nitrogen and water vapor. Care must be taken in trying to achieve ultra-low levels of NOx emissions with SCR, as imperfect combustion gas and reagent mixing can result in the slip of unreacted reagent together with the release of a remnant of the original NOx. In practice, obtaining such perfect mixing is difficult, yet much pollution reduction chemistry is based on perfect mixing. Catalytic combustion is an alternative to SCR for NOx reduction. Catalytic combustion, a promising technology, offers the possibility of practical operation at high temperature with extremely low emissions levels. Catalytic combustion achieves low NOx by reacting lean, premixed air/fuel mixtures on the surface of a catalyst. Typically, catalytic combustors are larger
2-24

EPRIsolutions Licensed Material Technology Description

than diffusion-flame or lean premixed combustors and often require modification of the gas turbine housing. Laboratory evaluation catalytic combustion units have shown that in steady operation for modest duration, emissions close to 1 ppm can be achieved at 15% oxygen (O2) and at turbine inlet temperatures around 2600F (1430C). However, this level was reached at least five years ago, and the gas turbine world has since been awaiting the release of commercial high- and medium-temperature catalytic combustors for general use. The entire catalytic products field is extremely secretive, and little information has been released for industry-wide use. Issues facing practical catalytic combustion include durability and catalyst life, degradation and the need for periodic catalyst replacement, mechanical integrity during transient conditions, startup operation, and the development cost of a product for a geographically limited market. NOx formation becomes appreciable at local temperature conditions of 2800F (1540C). With lean pre-mixed combustion, the temperature of the combustion products can be controlled so that this limit is not exceeded. Only large, new aeroderivative turbines (those used on new military and new jumbo passenger jets) and large (250-MW) central station power plants have turbine blade cooling technology capable of utilizing combustion products as hot as 2600F (1430C). Thus, for now, the turbine and combustor technologies of large, high-performance gas turbines appear to have converged with turbine inlet (average) temperature in the range of 2600F (1430C). Small aeroderivative and industrial gas turbine hot-section technology typically lags behind that of the leading large, new products by about a decade. Manufacturers of small gas turbines, both aeroderivative and industrial, have been busy developing lean pre-mixed combustors that can deliver low NOx combustion products in the 2100-2400F (1150-1320C) temperature range for their products. Turbine products with the greatest sales volumes have been the first to benefit from such development, as that is where the greatest profits lie. As manufacturers experience matures, it becomes easier and less expensive to apply the technology to products with smaller sales volumes, as progressively less investment is needed to develop them. It is expected that products with low emissions (9 ppm NOx at 15% O2) in most size ranges will enter the market very soon, if they are not there already. Older, more polluting products will still be made for sites with minimal air pollution control requirements, such as third world and offshore applications. Advanced Gas Turbine Cycles The earliest improvement in the Brayton cycle, upon which gas turbines are based, was the addition of a recuperator, which uses turbine exhaust heat to preheat combustor inlet air. Other early advancements were the intercooler and the combined cycle. Later advancements included steam and water injection. As technology in chemical engineering and other fields developed, chemical recuperation, humidification, and other applications became possibilities for power generation as well. Recent Developments in Advanced Cycles In 1992 GRI contracted with Fluor Daniel for an evaluation of advanced gas turbine cycles. The results were reported in Evaluation of Advanced Gas Turbine Cycles (GRI-93/0250), which evaluated what could be accomplished with technology then under development using the newest, most advanced, large gas turbines. The gas turbines considered as starting points were those under development by industry, some with DOE support, and the newest large
2-25

EPRIsolutions Licensed Material Technology Description

aeroderivative machines. Additionally during the early 1990s, a private group consisting mainly of electric and gas utility companies, called the Collaborative Advanced Gas Turbine (CAGT) group, sponsored research on how to adapt these large, new aeroderivative turbines (the type used for propulsion of the Boeing 777 and 747 aircraft) for more attractive commercial use. After evaluating aeroderivative gas turbine-based combined cycles, the CAGT group focused on the Intercooled Aeroderivative (ICAD) engine. This engine, with an increase in pressure ratio, the addition of an intercooler, and other modifications, had a power output about double that of the stationary version of the basic aircraft propulsion engine alone. The studies took as their point of departure actual engines, rather than rubber engines (idealized engines of an arbitrary size to enable comparisons to be made at specific power levels), which had the newest technology. This decision, to base the analysis of potential future engines on improvements in existing engines, was made because gas turbine development is quite expensive, and most new models are based on improvements in existing models. These improvements include increasing the pressure ratio and mass flow by adding stages ahead of the first stage (zero staging) and improving the hot-section materials and blade and vane cooling by using recent technology advances, thereby upgrading the turbine inlet temperature capability. Consequently, the plants in the study delivered power outputs of varying capacity. The range of gas turbine size was 75-250 MW, typical of central station equipment. The technical conclusions apply to smaller sizes, but the actual performance (efficiency and power) and economics would be less favorable in small gas turbines due to economic and engineering advantages of scale. In all, 15 cycles were analyzed for efficiency and cost. The cost model was that of the architectengineering profession that designs and builds large central station power plants. The four most promising advanced cycles identified were 1) the intercooled cycle based on an aeroderivative gas turbine, 2) the intercooled and steam-injected cycle, 3) the intercooled and chemically recuperated cycle, and 4) the humidified air cycle. (The recuperated, intercooled, and intercooled and recuperated cycles were described in earlier sections.) Of these cycles, the simplest configurations are the intercooled cycle and the intercooled and steam-injected cycle, which might be considered for use on small gas turbines. The other advanced cycles are applicable mainly to large, central station baseload power generation. Readers interested in these cycles are referred to the cited report. Steam injection has been used with gas turbines as small as 5 MW in industrial applications for well over a decade. An interesting fit for steam injection exists in areas where both heating and air conditioning are important applications. Steam can be used for heating during the winter, then during the summer it can be injected into the gas turbine to generate additional power for the air conditioning load. Intercoolers are used in industrial air compressors and on truck and some automobile turbochargers, as well as on the Navys intercooled, recuperated gas turbine (24.4 MW continuous-duty ISO rating) being developed for ship propulsion. Intercooling heat exchangers are not believed to require substantial development for stationary power applications. Seven years have passed since the advanced gas turbine report was published, yet turbine manufacturers have expressed little apparent interest in developing any of the options for advanced cycle, large central station generation machines. Current stationary gas turbine development, partially in conjunction with DOEs advanced turbine system (ATS) program,
2-26

EPRIsolutions Licensed Material Technology Description

appears to be concentrated in the field of higher turbine inlet temperature to reach lower specific cost and higher efficiency. Recuperated cycles have been practiced for over 50 years in the retrofit of pipeline compressor station gas turbine drives. In addition to the Mercury 50, the recuperated cycle is used in the TF-1500 tank engine and in recently commercialized microturbines. Solar manufactures recuperators for microturbines and other turbines. Perhaps the emerging distributed generation market will provide an impetus for developing improved machines of higher power density and increased efficiency, which could also reduce cost. Development of humidifiers to operate with gas turbines has been under way, as has work on a gas turbine specifically engineered for operation with a humidifier. Humidified gas turbine cycles have been discussed and analyzed for at least a decade. Humidification adds water vapor (steam) to the compressed air flowing into the combustor. Humidifiers add a greater quantity of water vapor to the compressed air than do HRSGs. Although this increases cycle efficiency, it requires a special expansion turbine engineered for the greater mass flow. Although this is not particularly difficult, it does requires engineering and testing, so a sufficient market must be perceived to recover product development costs. It has been recently reported that a 12-MW cascaded humidified advanced turbine (CHAT) system is being considered for development. In the CHAT system, the compressor is split into three sections by the addition of two intercoolers. The compressed air is sent to a humidification tower where water contacts and evaporates into the air, thereby augmenting the mass flow into the combustor and turbine. A recuperator is used ahead of the combustor. The combustion products then pass through an expansion turbine designed for the augmented mass flow. This is similar to steam injection, only with a greater amount of water vapor in the expansion gas, and consequently more power and higher efficiency. These benefits have to be considered in light of the added cost of such a system. The reported estimated cost for such a CHAT cycle gas turbine is between $700 and $750/kW, and the efficiency is claimed to be 46.4%. Such a product could well be attractive for distributed generation applications. It could be operated at a high annual load factor because it can economically compete with grid power at most times of the day. The exception is during deep night hours (typically 11 p.m. to 6 a.m.), when baseload plants would be operating at little over marginal cost of production and the transmission lines have plenty of unused capacity. Wet Compression Recently Siemens Westinghouse has offered wet compression, in addition to evaporative cooling, as an option on its large, new, 501 gas turbines. As an example, on a W 501 D5A, when the ambient (dry bulb) temperature is 100F (38C), the power increases from 102 MW to 109 MW with evaporative cooling and to 126 MW with evaporative cooling and wet compression. Dow Chemical has practiced wet compression on Westinghouse W 501 gas turbines for over five years, without deleterious effects on the machine. A slight decrease in heat rate (increase in efficiency) of 1-2% results from the implementation of such combined processes. As in conventional inlet air cooling, the wet compression option provides fog cooling of the inlet air, but the amount of water sprayed is greater than what is needed to cool the inlet air to saturation. The inlet fog is controlled so few droplets (measured by mass) are larger than 20 microns in diameter. With conventional size distribution of droplets, this means few droplets
2-27

EPRIsolutions Licensed Material Technology Description

over 10 microns when the size distribution is counted on a number basis. This difference in how droplets are counted is important, because the large, heavy droplets are the ones that cause the damage to compressor blades, and the mass of the droplets increases as the cube of the diameter. The fine fog carries over into the compressor itself, evaporating as the air increases in pressure and heats up. Substantial decreases in compressor work result from such continuous intercooling of the first few stages of compression. Siemens Westinghouse offers wet compression only on its largest gas turbines. Based on successful operation and lack of problems, six W 501 D5s have recently been to converted to wet compression. Investment in the Development of New Products Engineers have been inventing and analyzing innovative power generation cycles since the science of thermodynamics was born. Most of this effort has concentrated on improving the cycle efficiency of large, baseload, central station plants and the efficiency and power of engines (automotive, truck, boat, and aircraft). Because of the large opportunity for profit in baseload power plant and mobile engine markets, these applications attracted large investment in advanced technology and new products. Fuel cost is the largest variable cost in power production, and historically, fixed power plant costs were recovered through the regulatory process in the rate base. Consequently, engineers concentrated on increasing efficiency to reduce fuel cost. Additionally, engineers tended to have an intellectual dislike for wasting resources. In the historic market, utility power equipment was purchased mainly for baseload and peaking applications. While a modest market (in comparison to the baseload market) has existed for industrial power and cogeneration, to date there have been insufficient sales to provide major incentive for developing new products for intermediate duty. The exception to this is the Mercury 50, a 3.8 MW (continuous-duty rating), 42% efficiency (LHV), recuperated gas turbine recently developed and introduced by Solar Turbines. Aeroderivative gas turbines benefit directly from DOD R&D programs, which are focused on more effective hot-section (blade and vane) cooling, hot-section materials development, and improved compressor flow path efficiency. These DOD gas turbine advancements pass down into the utility and industrial gas turbine products via professional society technical papers and the migration of engineers from one gas turbine manufacturer to another as business activity in these sectors waxes and wanes. Since the energy crises of the 1970s, the U.S. Department of Energy (DOE) has supported R&D on gas turbines for utility and industrial applications and in several other energy fields that were previously the exclusive domain of the private sector. In fact, the Mercury 50 enjoyed substantial DOE support. This machines operating characteristics (efficiency, price per kW, and concern for minimum maintenance and quick repair) were focused on the requirements of the intermediate-duty market. The decision to develop a new or substantially improved piece of power generation equipment for the industrial (and commercial) power market depends on the economics of obtaining an adequate return on investment. This economic criterion involves comparing the development cost with the expected increase in profits that can be attributed to the introduction of the new product. If the market for the new product encroaches upon the existing market, the new product would merely displace the manufacturers existing product, and the circumstances typically do not favor such an investment. The decision to invest in the development of a new gas turbine
2-28

EPRIsolutions Licensed Material Technology Description

involves market research, engineering planning of the development activity, and financial analyses of the time profiles of development and market entry costs and future sales and profits. Only when such business analyses show adequate return on investment are development projects begun. Not all new industrial products receive the market success of the Xerox dry photocopier machine. The power equipment and generation industries are traditionally conservative. It is expected that when adequate market expansion indications are perceived, the development of new power generation products will be undertaken. Due to the high cost of both technology and product development, gas turbine technology advancement usually follows a path of initiation for aeronautical application, followed by adaptation for large central station gas turbines, and then diffusion into the medium and small industrial gas turbine products. Alternative Fuels During the energy crisis several alternative gas turbine fuels, such as heavy (residual) oil and direct and indirect firing of coal and biomass, were tested. These experiments found that gas turbines work most reliably with very clean fuels. Development work on the use of solid fuels has essentially ended because the cleanup cost of solid-fuel combustion products, or the cost of high-temperature heat exchangers to operate in a dirty, hot gas environment, was found to be excessive for commercial purposes. Gas turbine blade coatings were developed and are in use, which protect the blade surface from corrosion and deposits for an appreciable period of time. Gas turbines being made today exclusively burn very clean fuels. During 2000, the high prices of oil and natural gas are a concern, and their effect on fuel choice for gas turbines is being discussed. Prior price spikes and supply shortages caused major inconveniences to individuals and businesses, leading to much discussion of alternative fuels and new energy conversion technology. Today, after the energy crises of 1973 and 1979, only one synthetic fuel plant is operating, and none is expected to be built in the foreseeable future. No commercial solar electric plants (as distinguished from subsidized demonstration plants) are being built for commodity power generation, and only a few commercial fluidized-bed clean coal combustion steam plants have been built in the United States. Many of those fluidized-bed plants that have been built recently are economic principally because they burn a low-grade, quite inexpensive, solid fuel. In the field of alternative energy resources, major energy companies are now determining what their options are, should a real need for a replacement of oil as the major energy resource develop. Petroleum prices of $38/barrel, and consequential high gasoline, diesel fuel, heating oil, and jet fuel prices, have increased public ire but are insufficient to trigger business decisions to switch to alternatives. Apparently the business community views the current circumstances of petroleum price volatility as temporary phenomena, resolvable by political means or by additional petroleum production. Although the Organization of Petroleum Exporting Countries (OPEC) exerts partial control of petroleum production, substantial amounts are now produced by non-OPEC countries, and petroleum discoveries are increasing in non-OPEC locales. Extreme petroleum supply variations are mitigated somewhat by OPECs concern with the effect that major alternative energy investments might have on petroleum demand. Business leaders appear to expect that political and free-market forces will restore a comfortable equilibrium in petroleum price and supply. Deregulation of domestic natural gas production accomplished a
2-29

EPRIsolutions Licensed Material Technology Description

return to market forces in the 1980s, and business leaders most likely expect this to continue. Unfortunately, during this transition from one era to the next, painful price spikes can be expected from time to time. In the near future major corporations cannot be expected to make substantial investments in alternative energy resources or to develop highly efficient but expensive technology unless they expect economic benefits at petroleum prices below $40/bbl.

2-30

EPRIsolutions Licensed Material

3
EMISSION CONTROL & CATALYTIC COMBUSTION
Introduction The use of stationary gas turbines for power generation has been growing rapidly, a trend predicted to continue well into the future. Factors contributing to this growth include advances in turbine technology, operating and siting flexibility due largely to gas turbines clean emissions profile and low capital cost in medium and large units. Restructuring of the electric utility industry will provide new opportunities for onsite generation. In a competitive market, it could be more cost-effective to install small distributed generation units like gas turbines within the grid rather than constructing large power plants in remote locations requiring expansion of the transmission and distribution system. Beyond the issue of cost, public opposition to visible signs of industrial activity manifests itself in community resistance to distribution lines and large power plants. Small, onsite generation can be permitted along with the building itself, the equipment is not visible externally, fuel is readily available, and the environmental impact is minimal in terms of air emissions. For the customer, onsite generation can provide added reliability as well as leverage in negotiating the cost of purchased power. Emission Control Technology Gas turbine emission control technologies are continuing to evolve, with older technologies being gradually phased out as new technologies are developed and commercialized. It has been recognized that add-on emission control technologies are cost-prohibitive in small gas turbine sizes; nonetheless, such controls have been mandated by stringent regional air quality regulations in many parts of the country. In the approaching competitive power market, the opportunities for small gas turbine installations will grow, but the project economics will more than likely be negatively impacted by such regulations. One of the key issues addressed in virtually every gas turbine application is emissions, particularly NOx emissions. Decades of R&D have significantly reduced gas turbine NOx emissions from uncontrolled levels. The evolution of cost-effective, low emission gas turbine combustion systems includes diluent injection (water or steam) and lean premixed (LPM) approaches. LPM combustors, also referred to as dry low emissions (DLE) or dry low NOx (DLN), can differ in hardware from manufacturer to manufacturer, but the principle is the same. LPM combustors reduce peak flame temperatures by mixing fuel and air before combustion and by keeping the fuel-to-air ratio as low (lean) as possible. This avoids hot spots regions where the fuel/air mixture burns in near-stoichiometric proportions, creating high-temperature zones that produce high levels of NOx. Avoidance of local hot spot conditions inhibits NOx formation through thermal fixation of atmospheric nitrogen. LPM combustors are now offered on nearly all new gas turbines and as retrofits for existing popular models.

3-1

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

Ratcheting emissions limitations are driving further improvements in low emission combustion technology. Gas turbine manufacturers are pursuing ultra-lean premix, which pushes the theoretical limits of LPM techniques without compromising efficiency improvements or turbine durability, which is affected by flame stability (undesirable vibrations and acoustics that tend to occur as the lean limit is approached). The primary benefit of LPM gas turbine combustors, both conventional and catalytic, is to minimize local combustion temperatures. This reduces thermal NOx production and, by reducing combustion instability and acoustics, allows combustion to proceed well below the stability limit of diffusion flames. Catalytic Combustion for Gas Turbines In catalytic combustion, the presence of a catalyst allows fuel to be very lean throughout the combustion mixture, so that combustion occurs at local temperatures below which significant NOx quantities are formed. Catalytic combustion is a flameless process, allowing fuel oxidation to occur at temperatures approximately 1800oF (980oC) lower than those of conventional diffusion-flame combustors. Catalytic combustors are being developed to control NOx emissions down to 3 ppm. Preliminary test data indicate that catalytic combustion exhibits low vibration and acoustic noise, only one-tenth to one-hundredth the levels measured in the same turbine equipped with DLN combustors. Compared to noncatalytic, lean premixed combustion, catalytic combustion has the advantage of less severe flame stability problems (less vibration and noise at extremely lean chemical conditions). This means that leaner mixtures become practical to burn, with less NOx resulting at specified combustor outlet conditions. Gas turbine catalytic combustion technology is being pursued by developers of combustion systems and gas turbines and by government agencies, most notably the U.S. DOE and the California Energy Commission (CEC). Catalytic combustion at atmospheric pressure has a long history of development for industrial, commercial, and residential applications. The potential benefit for gas turbine catalytic combustors is to allow a lean premixed combustor to achieve ultra-low NOx (low single-digit ppm), stable combustion stability, and low acoustics without requiring high-cost exhaust clean-up methods. One problem with catalytic combustors is the potential for autoignition of the fuel upstream of the catalyst. Although the air/fuel ratios are well below the lean flammability limit, and in theory should not be susceptible to autoignition, local pockets of rich fuel mixtures can exist near the fuel injector and ignite. Mixing must be achieved quickly to prevent fuel-rich pockets from forming. Optimum catalyst performance also requires the inlet air/fuel mixtures temperature, composition, and velocity profile to be completely uniform. This assures effective use of the entire catalyst area and prevents damage to the substrate due to local high gas temperatures. Past efforts at developing catalytic combustors for gas turbines have achieved low, single-digit NOx ppm levels but have failed to produce combustion systems with suitable operating lifetimes. This was typically due to high-temperature and cycling damage and to the brittle nature of the materials used for catalysts and catalyst support systems. An additional issue with catalytic combustors is the durability of the catalyst. Research suggests that the catalyst will deteriorate during prolonged operation at high temperature. Thermal degradation results from loss of surface area caused by sintering and volatilization of active metals, such as platinum, which oxidizes at temperatures above 2010oF (1100oC).

3-2

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

Currently, catalytic combustor developers and gas turbine manufacturers are testing nonbrittle metal catalytic and partial catalytic systems that are intended to overcome the problems of past efforts. Although catalytic combustion has yet to be fully commercialized in the marketplace, these efforts have evolved to the point where practical application of catalytic combustion in gas turbine systems has become feasible. Catalytic combustors capable of achieving NOx levels below 3 ppm are in full-scale demonstration. Durability of the catalyst module has not yet been proven and must be achieved in a practical and economic manner for commercialization to be achieved. Periodic catalyst replacement might be acceptable to turbine operators, but the cost, the complete catalyst replacement interval, and the turbine reliability between catalyst replacement must not compromise the turbines availability. Small gas turbines can be expected to be turned on and off frequently perhaps daily or even more often and catalyst durability must remain uncompromised by the accompanying thermal cycling. Should periodic catalyst replacement become the norm, replacement costs would be handled as an operation and maintenance (O&M) cost. While each catalytic combustor design is unique, most are larger than conventional combustors and probably would require enlargement of the gas turbine housing. Although this might be technically feasible, the cost of such special housings adds to the total to be considered. For gas turbine manufacturers to offer catalytic combustion commercially, the engineering development costs of gas turbine modification must be acceptable in terms of the expected market size and purchaser action timetable. Despite lower emissions being mandated in California and other areas with severe air pollution, such levels are not required in many segments of the world market where small gas turbines are sold. Consequently, a higher cost, catalytic combustion small gas turbine would be regarded as a special, higher priced, product for a market niche, and investment in its development has to be examined in the perspective of its limited market. Catalytic Combustion System Developers Leading developers of ultra-low emissions systems for gas turbines include: Alzeta Corporation, Santa Clara, CA (Alzeta) Catalytica Combustion Systems, Inc., Mountain View, CA (Catalytica) Precision Combustion, Inc., New Haven, CT (PCI)

Alzeta Since the mid-1990s, Alzeta has been working with support from CEC and other industry partners on developing ultra-low emission combustors for gas turbine systems. Previously, with support from GRI, Alzeta had developed fiber matrix radiant burners for industrial applications that provided many of the beneficial features of catalytic combustors without the high cost of conventional precious metal catalysts. Gas turbine partners throughout this development have included Honeywell (previously AlliedSignal), Solar Turbines, and Westinghouse. Alzetas approach, called the surface stabilized combustor, builds upon Alzetas industrial radiant burner. The technology has potential application to large central station gas turbines, industrial size gas turbines, and microturbines. Recent tests funded by CEC and DOE have shown simultaneous emissions of less than 2 ppm for NOx, CO, and unburned hydrocarbons.
3-3

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

Catalytica Catalytica has apparently evolved the farthest toward a practical catalytic combustion system that is nearing commercial readiness. Catalytica has formed working relations with General Electric (GE), Solar Turbines, Rolls-Royce, Pratt & Whitney Canada, and Kawasaki. Catalytica has developed its Xonon catalytic combustion system, an all-metal catalyst substrate that eliminates the potential problems associated with the limitations of high-temperature ceramic substrates. To avoid damaging the metal substrate, the maximum temperature reached in the catalyst is limited to approximately 1700oF (930oC). All fuel and air is added upstream of the catalyst. Approximately 50% of the fuel is oxidized in the catalyst, which limits the temperature rise. The remaining 50% of the fuel is oxidized downstream of the catalyst. Catalytica performed a successful 1000-hour test of the Xonon catalytic combustor in a 1.5-MW Kawasaki gas turbine that concluded in mid-November 1997. Another 1.5-MW Kawasaki turbine at a cogeneration plant in Santa Clara, CA, has been equipped with a catalytic combustor that began operation in October 1998 and operated through mid-2000. Xonon catalytic combustors have also been tested in large GE turbines at the GE test facility in Schenectady, NY. NOx averaged less than 3 ppm and CO less than 5 ppm (corrected to 15 percent O2) during a test on a Frame 9E turbine. GE recently announced a Memorandum of Understanding with Catalytica to develop catalytic combustors for all GE turbine models through Frame 7E (78 MW). PCI Applications for gas turbines have been under development around the world since Dr. William Pfefferle, PCIs chief scientist, invented the first gas turbine catalytic combustor in 1970. A primary PCI thrust is developing clean, efficient, ultra-low NOx catalytic combustors for turbines used in central station plants and distributed power generation. These catalytic combustors are expected to combine clean emissions with efficient high firing temperatures and wide turndown stability, operability, and reliability. Catalytically stabilized thermal combustors (generally called catalytic combustors) use catalytic surface reactions to extend the lean stability limit of gas-phase combustion, allowing stable and complete combustion to occur at cooler temperatures and shorter average residence times than otherwise required. Multiple approaches are feasible, with PCI technology tailoring product development to individual application requirements. PCI is developing gas turbine catalytic combustors that enable low single-digit ppm emissions along with high combustion stability and a wide operating range. These devices can achieve the low emissions levels of exhaust treatment technologies such as SCR or SCONOx without the associated cost and reduction in gas turbine efficiency. Current PCI catalytic combustors at pressure have demonstrated sustained, natural gas-fueled combustor operation at air inlet temperatures less than 750F (400C) with emissions of NOx below 3 ppm and CO below 5 ppm at burner outlet temperatures above 2800F (1540C). PCIs advanced technology catalytic combustor is attracting interest from government agencies as well as from power-generation and industrial gas turbine manufacturers. Siemens Westinghouse Power Corporation and PCI have a long-term business agreement to develop, manufacture, and sell low single-digit NOx catalytic combustors for Siemens Westinghouses F, G, and ATS class combustion turbines. PCI is also working with other gas turbine manufacturers under shorter term agreements.
3-4

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

PCI has other development programs related to catalytic-based enhancements to gas turbine combustion. PCI is developing catalytic pilot burners to allow conventional state-of-the-art natural gas-fired DLN gas turbine engines to achieve lower NOx emissions with high combustion stability, wide low emissions turndown, and good operability. These small devices are being developed to fit into existing DLN combustor envelopes, providing a low-cost design upgrade for new machines and the potential for a retrofit product for installed machines. DOE and other agencies, as well as other gas turbine manufacturers, are supporting this development. PCI catalytic combustor technology is also being developed for both liquid-fueled and natural gasfueled gas turbines. Applications include distributed power and mobile sources. PCI sells catalytic liners for new combustion turbines and for retrofitting existing gas turbines to reduce CO and other emissions. Catalytic combustion must be integrated with the combustor design of individual gas turbine models. Depending on the combustor design, development costs and subsequent modification costs to the turbine can vary significantly among models and the various gas turbine manufacturers. Low Emission Gas Turbine Projects All three of these companies have been awarded contracts under a recent DOE solicitation for Low Emission Gas Turbine Technologies for Distributed Generation. A summary of awards is listed below. The companies are also participants on awards to gas turbine manufacturers under this solicitation. Demonstration of the Surface Stabilized Combustor for Advanced Industrial Gas Turbines, led by Alzeta, with DOE funding of $2 million. (Partners: CEC, Chevron, Honeywell Power Systems Manufacturing Resources Inc., National Energy Technology Center, Siemens Westinghouse, and Solar Turbines.) The team will develop a novel stabilized combustion technology for industrial gas turbines. Early data indicate high potential for low emissions and cost-effectiveness. Component Development to Accelerate Commercial Implementation of Ultra-Low Emissions Catalytic Combustion, led by Catalytica, with DOE funding of $1.6 million. (Partners: Rolls-Royce Allison and Solar Turbines.) Catalytica is commercializing catalytic combustion technology on a 1.5-MW, grid-connected industrial turbine at Silicon Valley Power in Santa Clara, CA. The project partners will research extending the catalyst longevity for use with other turbines and lowering the cost of emissions prevention. Fuel-Flexible Ultra Low-Emissions Combustion System for Industrial Gas Turbines, led by Honeywell Engines & Systems, Phoenix, AZ, with DOE funding of $660,000. (Partners: PCI, Texas A&M University, and Vericor Power Systems.) Honeywell seeks to develop an innovative, fuel-flexible, air-staged, catalytic gas turbine combustion system with closedloop control. The system will offer low emissions, fuel flexibility, and increased market acceptance of gas turbines. Catalytic Combustor for Ultra-Low NOx Advanced Industrial Gas Turbines, led by PCI, with DOE funding of $1.4 million. (Partners: Haynes International, Honeywell Engines & Systems, Siemens Westinghouse, Solar Turbines, and University of Connecticut). PCI proposes to advance the development of a new catalytic combustion technology that
3-5

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

promises many advantages over existing systems. Near Zero NOx Combustion Technology for Advanced Mercury 50 Gas Turbine, led by Solar Turbines, with DOE funding of $200,000. (Partners: Catalytica, Combustion Systems Inc., PCI, and University of California-Irvine.) Solar Turbines proposes a fully integrated catalytic combustion system for the new Mercury 50 gas turbine.

Economics This section compares the projected catalytic combustion costs with the costs of the principal emission control technologies being employed or nearing commercialization for reducing NOx in stationary gas turbines. Cost data (Table 3-1) are expressed in dollars per ton of NOx removed ($/ton) and cents per kilowatt-hour (/kWh) for gas turbines in the 5-MW, 25-MW, and 150-MW output ranges. The $/ton value indicates the typical cost of a control technology to remove a ton of NOx from the exhaust gas. This value is determined by dividing the ownership (capital and O&M) cost of the control technology by the tons of NOx removed. A relatively lower $/ton value means that the technology is more cost-effective in removing NOx than alternative control technologies.
Table 3-1 Cost Impact Factors for Selected NOx Control Technologies Source: ONSITE Energy
Turbine Output Median value NOX EMISSION CONTROL TECHNOLOGY DLN (25 ppm) Catalytic Combustion (3 ppm) Water/Steam Injection (42 ppm) Conventional SCR (9 ppm) High Temperature SCR (9 ppm) SCONOx (2 ppm) 5 MW Class $/ton /kWh 25 MW Class $/ton /kWh 150 MW Class $/ton /kWh

260 957 1,652 6,274 7,148 16,327

0.075 0.317 0.410 0.469 0.530 0.847 1.060

210 692 984 3,541 3,841 11,554 2,202

0.124 0.215 0.240 0.204 0.221 0.462 0.429

122 * 371 476 1,938 2,359 6,938

0.054 * 0.146 0.152 0.117 0.134 0.289

Low Temperature SCR (9 ppm) 5,894 * 9-25 ppm "/kWh" based on 8,000 full load hours

The $/ton value is a useful comparative indicator when the inlet and outlet concentrations are the same for each group of technologies being evaluated. NOx can be controlled to within a feasible limit for a specific control technology. This limit is largely independent of a gas turbines uncontrolled NOx emission rate. Therefore, to obtain a meaningful comparison, the uncontrolled NOx exhaust concentrations must be considered when evaluating the costeffectiveness in terms of $/ton applied to different turbines makes and models.

3-6

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

The /kWh value indicates the electricity cost impact of a particular NOx control technology and is independent of the tons of NOx removed. This value represents a unit cost for NOx control that must be added to other ownership costs associated with a gas turbine project. The /kWh is determined by dividing the ownership cost of the NOx control technology by the amount of electricity generated by the gas turbine. A comparison between /kWh values is most meaningful for technologies that control NOx to an equivalent ppm level. When performing cost comparisons among technologies that control NOx with different inlet/outlet emission rates, capital and operating costs might need to be adjusted to perform the analysis on an equivalent basis. In this study, capital and operating costs provided by manufacturers were restricted to turbine projects readily available at the time of the inquiry, which resulted in various gas turbine models and inlet/outlet NOx emission rates. Also, some manufacturers considered certain cost numbers as proprietary, which prevented an equitable comparison in some cases. Figure 3-1 compares the cost of power impact (/kWh) for various control technologies. Controlled NOx concentrations are indicated below each technology in the figure. In general, the results shown are ordered from highest cost to lowest cost impact. The cost impact factors are based on 8000 full-load operating hours. The majority of base-loaded gas turbines typically operate at lower full-load hours, so actual cost impacts could be significantly higher.

1.2 5 MW Cost of Power Impact (/kWh) 1 0.8 0.6 0.4 0.2 0


Low Temp SCONOx SCR (2ppm) (9ppm) High Temp SCR (9ppm) Conv. SCR (9ppm) W/S Inj (42ppm) Catalytic (3ppm) DLN (9-25ppm)

25 MW 150 MW

Figure 3-1 Cost Impact of Emissions Control Technologies Source: ONSITE Energy

3-7

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

Prospective catalytic combustion technologies exhibit lower cost impacts than add-on controls for both small and large gas turbines. Projected costs for catalytic combustors indicate that the /kWh cost is 2 to 3 times higher than for a DLN combustor alone. The catalytic combustor can achieve NOx levels of less than 3 ppm, while the most advanced DLN combustor reaches down to only 9 ppm. To achieve NOx levels below 5 ppm, the DLN-equipped turbine requires postcombustion NOx control devices such as SCR or SCONOx. Catalytic combustion is not yet fully commercialized, and catalyst durability is still being proven through demonstrations. The longterm (8000-hour) durability test of Catalyticas Xonon technology on a 1.5-MW gas turbine at Silicon Valley Power (Figure 3-2) is intended to address this issue. In addition, the capital cost of adding catalytic combustion to a turbine combustor will be a strong function of individual turbine designs and therefore will vary significantly. Catalytic combustion cost estimates presented in Table 3-2 are based on the anticipated performance of the Catalytica Xonon catalytic combustion technology. The cost estimates assume catalyst replacement on an annual basis; however, catalyst life is currently being tested at several gas turbine installations. Catalyst durability is an important milestone towards commercialization that has not been currently demonstrated.

3-8

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion

Figure 3-2 Catalytica Installation at Silicon Valley Power Source: Catalytica

3-9

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion Table 3-2 Catalytic Combustion Cost Comparison (Incremental Annual Cost Compared to Conventional Uncontrolled Diffusion-Flame Combustor) Source: Onsite Energy 5 MW Class Gas Turbine Output (MW) Heat Rate (Btu/kWh) Heat Content (Btu/lb) Fuel Flow (Btu/hr) Hours of Operation (hr) Fuel Flow (MMBtu/yr) CAPITAL COST ($) ANNUAL COSTS Equipment Life (yrs) Interest (%) Capital Recovery Factor Capital Recovery ($) Catalyst Replacement ($) Other Parts and Repairs ($) Annual Maintenance Contract ($) Major Failure Impact ($) Taxes and Insurance ($) TOTAL ANNUAL COSTS ($) Uncontrolled Emissions (ppm) Uncontrolled Emissions (tons/yr) Controlled Emissions (ppm) Controlled Emissions (tons/yr) NOx Removed (tons/yr) COST EFFECTIVENESS ($/ton) ELECTRICITY COST IMPACT (/kWhr) 15 10 0.131 28,543 66,100 8,320 5,000 15,293 8,684 131,940 150 140.6 3 2.8 137.8 957 0.317 15 10 0.131 68,867 253,740 42,080 5,000 61,052 20,952 451,691 130 688.5 3 15.4 653.0 692 0.215 15 10 0.131 189,799 1,193,676 271,840 5,000 265,425 57,745 1,983,486 210 5,426 3 77.5 5348.3 371 0.146 Representative 5.2 11, 240 20,610 2,836 8,000 467,584 217,100 25 MW Class Frame 26.3 12,189 20,610 15,554 8,000 2,564,626 523,808 150 MW Class Frame 169.9 9,481 20,610 78,157 8,000 12,886,575 1,443,629

Note: O&M costs for LM2500 used for large Frame as default. Costs based on Catalytica Combustion Systems Xonon technology

Catalytica provided production-run cost estimates of its catalyst module, including an allowance for turbine package modifications, in a recent study by ONSITE Energy for DOE. Catalyst life is estimated at one year, based on a guarantee offered by Catalytica. The cost does not include development costs, which could be substantial for turbine manufacturers depending on specific turbine and combustor designs. The costs provided by Catalytica do not imply that its technology will be applied to the turbines represented in the comparison.
3-10

EPRIsolutions Licensed Material

4
PROVIDERS
This chapter provides an overview of the predominant small gas turbine manufacturers. Their products, partners, and market positioning are discussed. Specific information that is generally not published for competitive reasons includes annual sales by gas turbine type, size range, applications, and market shares. The overall company strategies are reflected in their R&D, market emphases, acquisitions, and alliances. The somewhat universal strategy for most companies appears to include: Maintain and grow market share in their traditional market sectors Gain more of a presence in markets where their products have a competitive advantage Maintain or increase R&D Create strategic alliances Be receptive to potential acquisitions that can add expertise or enhance product offerings. The power generation market represents just a portion of targeted applications for small gas turbine providers. Virtually all gas turbines in this size range are also used as prime movers in other applications including aviation, oil and gas production and distribution, other mechanical drive purposes, and marine propulsion. The current demand for gas turbines for the power generation market exclusively is relatively small in the gas turbine size range reviewed for this report (300 kW to 5 MW). Other applications are needed to justify the costs of production, R&D, and product service. This chapter covers providers of gas turbines in the 300-kW to 5-MW size range. The manufacturers are discussed in alphabetical order. Original equipment manufacturers (OEMs) of generator sets (gensets) and genset packagers are also briefly discussed. Because packagers are somewhat ubiquitous and packaging is also in some cases handled by regional partners, this section focuses on selected leading North American packagers. A summary of the predominant small gas turbine providers, both manufacturers and packagers (Table 4-1), emphasizes predominantly those providers with a U.S. presence in the 300-kW to 5-MW market.

4-1

EPRIsolutions Licensed Material

Table 4-1 Key Providers of Small Gas Turbines (300 kW-5 MW) Company Alstom Power (ABB, EGT, Ruston) Services Design, Manufacture, Package, Installation, and Operations/Maintenance GT Type Industrial Models and Size TB-5000: 3.9 MW Typhoon: 4.3-5.3 MW Background, Strategic Partners, JVs, etc. Alstom recently acquired ABBs 50% share of their joint company ABB Alstom Power. The new company includes the previous ABB power generation capabilities with the exception of nuclear power plants. The gas turbine product line includes models from 1.6 to 265 MW. The product line includes the former European Gas Turbine (Ruston) gas turbines in the size range of interest for this report. Alstom also packages larger GE aeroderivatives. Former joint venture between Halliburton and Ingersoll-Rand. Now 100% owned by Halliburton. Markets are primarily oil and gas. The KG2 is an older model (Kongsberg) that is employed primarily in as a standby generator. Packager of GE aeroderivative engines. Past and current strategic partners include engineering firms Kellogg and Brown & Root for facility management in the oil and gas market. Japanese conglomerate that manufactures, packages, and distributes various power generation equipment. In this size range, packager of Pratt & Whitney Canada aeroderivative gas turbines. Recently completed transaction making Elliot Corporation a wholly-owned subsidiary.

Dresser-Rand (Halliburton)

Design, Manufacture, Package, Installation, and Operations/Maintenance

Industrial

KG2: 1.4-1.8 MW DR990: 4.4 MW

Ebara

Design, Package, Installation, and Operations/Maintenance

Aeroderivative

PW4: 420 kW PW6: 585 kW PW7: 715 kW PW12: 1.2 MW PW14: 1.4 MW PW18: 1.6 MW PW36: 3.3 MW

4-2

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion Table 4-1, contd. GAS Power Systems

Design, Package, Installation, and Operations/Maintenance

Aeroderivative

Avco Lycoming T-53: 1.2 MW

GAS Power Systems is a division of G.A.S. (General Aviation Services) Capital, a holding company of international aviation companies. The product concept utilizes an overhauled helicopter turboshaft engine, proprietary control system, and a 3.32:1 gearbox in 1.2 MW Innovator Genset package. The genset is reported to be multifuel capable. The first unit is being field tested at a rural cooperative in Lubbock, TX. GE has announced planned acquisition of Honeywell. The current Honeywell company was created with the merger of AlliedSignal and Honeywell Inc. Created Vericor Power Systems, a joint company between Honeywell Engine Systems and MTU Aero Engines (DaimlerChrysler), in 1999. Mitsubishi Nagoya Works (Japan) packages AlliedSignal gensets and mechanical drive systems in Japan. Major Japanese manufacturer of industrial duty gas turbines for standby and continuous-duty. US presence historically through dealers. Recently opened US operations to market, package and service Kawasaki gas turbine gensets in the US. Over 170 units in North America. Sizable market for standby gas turbines (both mobile and stationary) of this size in Japan. Developer of 300 kW ceramic engine that has demonstrated 42.1% efficiency. Provider the test-bed engine for catalytic combustion testing by Catalytica.

Honeywell (AlliedSignal)

Design, Manufacture, and Operations/Maintenance

Aeroderivative

ASE8: 548 kW ASE40: 3.3 MW ASE50: 3.8 MW

Kawasaki Gas Turbine

Design, Manufacture, Package, Installation, and Operations/Maintenance

Industrial

S1A-03: 222 kW S1T-02: 430 kW S2A-01: 646-681 kW M1A-01: 1.2 MW M1A-03: 1.4 MW M1A-06: 1.5 MW M1A-11: 1.2 MW M1A-13: 1.4-2.3 MW M1A-23: 2.0 MW M1T-01: 2.3 MW M1T-03: 2.9 MW M1T-06: 3.0 MW M1T-13: 2.9 MW M1T-23: 4.1 MW
4-3

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion Table 4-1, contd. Nuovo Pignone (GE)

Design, Manufacture, Package, Installation, and Operations/Maintenance

Industrial

PGT2: 2 MW PGT5: 5.2 MW PGT5B: 5.9 MW

Subsidiary of GE since 1993. Manufacturers industrial gas turbines in the 2-10 MW range. These same gas turbine models are sold and packaged by GE Stewart & Stevenson in the US (under the model designation GE2, GE5, and GE10 rather than PGT). Packager of GE aeroderivatives. Dutch company developing and manufacturing a 1.6 MW gas turbine. OPRA has a low emission combustion development program based on combustion system for vehicle applications. Has licensed combustor to AlliedSignal (Honeywell/GE). OPRA has plans to partner with Daihatsu Diesel Manufacturing Company (Japan). Orenda (Canada) is an aerospace firm that historically manufactured various gas turbine models, under license, to the Canadian government for aircraft engine applications. Orenda has over 90 gas turbines in oil and gas and electric generation applications. It partners with Ukrainian turbine manufacturer SPE Mashproekt to package OGT2500 generator set. They no longer produce, but still service OT2, OT3, and OT5 gas turbines.

OPRA Optimal Radial Turbine

Design, Manufacture

Industrial

OP-16: 1.6 MW

Orenda

Design, Manufacture, Package, Installation, and Operations/Maintenance

Aeroderivative

OGT2500: 2.8 MW

4-4

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion Table 4-1, contd. Pratt & Whitney Canada

Design, Manufacture, and Operations/Maintenance

Aeroderivative

ST6: 500-560 kW ST18: 1.2-2.3 MW ST30: 3.3-4.9 MW ST40: 4.0-5.4 MW

A subsidiary of United Technologies that is a leading manufacturer of small gas turbines for aviation applications. Works with a network of specialized international packagers, Ebara (Japan), Nedalo (Netherlands), Tecnoenergia (Italy), Bazan Motores (Spain). Until recently UNC Industrial Power was its US OEM packager. Now works with several regional packagers in the US. PW Canada has in the recent passed worked with Catalytica on a catalytic combustion system for its ST18 and ST30 models. Part of a development team with DTE Energy and TurboGenset (U.K.) for a 400-kW mini-turbine system based on the ST5 engine. Ramgen is in the design, development, and test phase of a power generation system based on ramjet technology. It is the only company working with ramjets for stationary applications. Rolls-Royce acquired Allison and US Turbine in 1995-96. Acquired Vickers (including the former Ulstein group of companies). In 1999, completed purchase of the rotating products interests from Cooper Energy Services, part of Cooper Cameron Corp. Primary markets are industrial power generation, oil & gas production and distribution, and marine applications. US Turbine had been a packager of Allison and Kawasaki gas turbines. Allison has a network of worldwide packagers (Centrax (U.K), Ishikawajima-Harima Heavy Industry (Japan), and Hitachi Zosen (Japan)). Japan has had Stewart & Stevenson as past packager of their Allison product line.

Ramgen Power Systems

Design, Manufacture, and Testing

Aeroderivative

Ramgen Prototype: TBD (0.75-2.0 MW)

Rolls-Royce (Allison & Rolls Royce Energy Systems)

Design, Manufacture, Package, Installation, and Operations/Maintenance

Aeroderivative and Industrial

501KB5: 3.9 MW 501KB7: 5.3 MW

4-5

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion Table 4-1, contd. Schelde Heron B.V.

Design, Manufacture

Industrial

Heron: 1.4 MW

Schelde Heron (Netherlands) was formed in 1996 as a result of joint venture between Heron Exergy and Royal Schelde. Current development efforts are focused on ultrahigh efficient (43%) two-shaft, intercooled, recuperated 1.4 MW gas turbine. Major design, manufacturer, and packager of industrial gas turbines for power generation, oil &gas applications, and marine applications. Product line ranges from 1.2 to 15 MW. Recently developed the Mercury 50 recuperated gas turbine under the US DOE Advanced Turbine System Program. Manufactures recuperators and has a business arrangement wit Capstone Turbines to provide recuperators for its microturbine. Has a vertically integrated packaging, sales, distribution, and service network. Works with packagers in overseas markets (Tuma Turbomach in Europe and Nigata in Japan). Subsidiary of GE. Packagers the Nuovo Pignone line of small industrial gas turbines and GE aeroderivatives for industrial power generation, mechanical drive and marine applications. Prior to acquisition by GE had been a packager of Allison and Ruston gas turbines at various times. Part the Snecma Group. Packages the RollsRoyce Allison 501KB5 and 501KB7 under the names TM4000 and TM5000 respectively. Partnered with Ulstein (Norway) and Volvo (Sweden) to develop the Eurodyn gas turbine

Solar Turbines

Design, Manufacture, Package, Installation, and Operations/Maintenance

Industrial

Saturn: 1.2 MW Centaur: 3.5-4.7 MW Mercury: 4.0-4.2 MW

Stewart & Stevenson (GE)

Design, Package, Installation, and Operations/Maintenance

Industrial and Large Aeroderivative)

GE2: 2 MW GE5: 5 MW

Turbomeca

Design, Package, Installation, and Operations/Maintenance

Aeroderivative

TM1000 Mikala T1: 1.1 MW TM2500 Eurodyn: 2.5 MW 501KB5: 3.9 MW 501KB7: 5.3 MW

4-6

EPRIsolutions Licensed Material Emission Control & Catalytic Combustion Table 4-1, contd. Vericor

Design, Package, Installation, and Operations/Maintenance

Aeroderivative

ASE8: 548 kW ASE40: 3.3 MW ASE50: 3.8 MW

A jointly owned company of Honeywell Engines and MTU (DaimlerChrysler). The company is a marketing, sales and services organization to market the AlliedSignal line of aeroderivative gas turbines. US packaging partner is Mendenhall Technical Services. International partners include MTU (Germany), Industua de Turba Propulsores (Spain), Mitsubishi (Japan), and Korean Aerospace Industry (Korea). For VOC destruction packages Vericor teams with North American Energy Systems (Canada). A major Japanese manufacturer of diesel engines, Yanmar also producers a line of gas turbines primarily for standby.

Yanmar Diesel Engine Co.

Design, Package, Installation

Industrial

AT270S: 0.2 MWAT360S: 0.3 MWAT600S: 0.5 MW AT900S: 0.9 MW AT1200S: 1.1 MW AT1800S: 1.7 MW AT2700S: 2.4 MW

4-7

EPRIsolutions Licensed Material Providers

Company Profiles In terms of market activity, Solar Turbines and Rolls-Royce Allison are the leading suppliers of gas turbines in the North American market. Both companies have a strong presence in nonpower generation markets such as oil and gas production. Solar Turbines is the dominant U.S. manufacturer, with a large majority of its sales in oil and gas markets. The relative split between oil and gas and power generation markets shows a recent increase in power generation, but oil and gas is still larger. GE, with its pending acquisition of Honeywell coupled with its current subsidiaries Stewart & Stevenson and Nuovo Pignone, seems to be in prime position to compete with Solar Turbines if development of the small gas turbine market for power generation accelerates. From a worldwide perspective, GE/Nuovo Pignone, Alstom, Dresser-Rand (Kongsberg), Kawasaki Gas Turbines, and Honeywell (through its international partners) all have notable market share. Mechanical drive, marine, and oil and gas production and distribution applications combined outnumber power generation installations in this size range. Kawasaki has a line of standby gas turbines that has been successful in Japan. Eastern European manufacturers are just starting to emerge. The competitive nature of the gas turbine business and the capital required to develop and produce successful products have resulted in several noteworthy acquisitions, mergers, and strategic alliances, including: 1. The successful development of a new generation of low emission combustion technology by virtually all manufacturers 2. Rolls-Royces acquisitions of Allison Engine Company and US Turbine 3. Rolls-Royce and Allison jointly developing and transferring low emission combustion technology 4. GE, Solar Turbines, and Allison evaluating catalytic combustion technology with Catalytica 5. GEs acquisitions of Stewart & Stevenson and full ownership of Nuovo Pignone 6. Solar Turbines introduction of the recuperated Mercury 50 gas turbine, which resulted from the cooperative U.S. DOE Advanced Turbine System Program 7. ABB and Alstom forming ABB Alstom Power and Alstoms subsequent purchase of ABBs interest in the venture 8. Ebara acquiring full ownership of Elliott Company 9. A joint venture, Vericor Power Systems, being formed between Honeywell Engine Systems and MTU Aero Engines (a subsidiary of DaimlerChrysler Aerospace) 10. Kawasakis demonstration of 42% efficiency on a 300-kW ceramic gas turbine 11. Honeywells merger with AlliedSignal in 1999 12. Pratt & Whitney Canada teaming with DTE Energy Technologies and TurboGenset to jointly develop a miniturbine system 13. The announced acquisition of Honeywell by GE.

4-8

EPRIsolutions Licensed Material Providers

The following company profiles discuss selected leading providers of gas turbines in the 300-kW to 5-MW range. The providers are discussed in alphabetical order. Alstom Power Alstom Power, headquartered in France, is one of the worlds largest providers of power generation equipment. In May 2000, Alstom acquired ABBs 50% share of the joint company ABB Alstom Power. The new company includes the previous ABB power generation capabilities with the exception of nuclear power plants. Alstom Power provides virtually every type of power generation and air pollution control equipment and service. These includes gas turbines, steam turbines, hydro turbines, boilers, fluidized-bed systems, generators, heat recovery steam generators, fossil fuel handling systems, low-NOx burners, and turnkey power plants. It claims to have the worlds largest installed base of power generation equipment approximately 640,000 MW, or 20% of all capacity installed worldwide. The gas turbine product line includes models from 1.6 to 265 MW. Gas turbine sales provided 24% of Alstoms $8.6 billion (10 billion euros) of revenues in 1999. The gas turbine product line includes the former European Gas Turbine (Ruston) and ABB gas turbines in the size range of interest for this report. This product line includes the TB-5000 and the Typhoon. Both of these gas turbines are available in single-shaft (power generation) and twin-shaft (mechanical drive) packages. The Typhoon single-shaft gas turbine is shown in Figure 4-1. Alstom also packages larger GE aeroderivatives. U.S. offices are in Midlothian, VA, and Windsor, CT.

Figure 4-1 Alstoms Single-Shaft 5-MW Typhoon Gas Turbine

Dresser-Rand Dresser-Rand, headquartered in Olean, NY, has been a leader in energy conversion equipment for almost a century. Its primary markets are the oil and gas (production, processing, storage, and transportation), chemical, and petrochemical industries. Dresser-Rand, the former joint

4-9

EPRIsolutions Licensed Material Providers

venture between Dresser Industries (Halliburton) and Ingersoll-Rand, is now 100% owned by Halliburton. In 1985, Dresser joint-ventured with Kongsberg Vapenfabrikk (Norway) to form Kongsberg Dresser Power, which is the current Kongsberg operations of Dresser-Rand. Its markets are primarily oil and gas. One of Kongsbergs primary products was the KG2 gas turbine, a singleshaft, all-radial turbine with a single can combustion. A KG2 genset package is shown in Figure 4-2. The KG2 gensets have found international acceptance as standby power systems. DresserRand notes that the radial design allows for high operating and starting reliability.

Figure 4-2 Kongsberg Dresser Powers KG2 Genset

Dresser-Rand has discontinued production of its DR-990 gas turbine but continues to service it. Dresser-Rand is also a packager of GE and Rolls-Royce aeroderivative gas turbines. Past and current strategic partners include engineering firms Kellogg and Brown & Root for facility management in the oil and gas market. The KG2 is manufactured and packaged in Norway. A Houston office of Dresser-Rand provides sales and support for KG products in North America. Ebara Established in 1920, the Japanese Ebara Corporation is one of the worlds major manufacturers of fluid and gas transfer systems. Its market includes strong positions in pumps, compressors, fans, and refrigeration equipment. Ebara is also a prominent provider of environmental systems, in particular for the treatment of water and waste as well as waste utilization. Ebara is actively developing new energy sources, including fuel cells, renewables (wind and solar), and other environmentally clean energy systems. Ebara also provides precision machinery for various industries.

4-10

EPRIsolutions Licensed Material Providers

Gas and steam turbines have been a part of Ebaras Machinery and Equipment Group. This group also includes pumps, fans, turbocompressors, and refrigeration and cooling products. This group provided $1.6 billion of sales revenues, approximately 30% of Ebaras total sales revenues. In this size range of gas turbines, Ebara is a packager of Pratt & Whitney Canada aeroderivative gas turbines. Ebara recently acquired Elliott Corporation (Jeanette, PA) as a wholly owned subsidiary. Ebara had previously owned 50% of Elliott and in April 2000 purchased all remaining shares. Elliott has been a supplier of turbocompressors to the oil and gas industry. Ebara will spin out its compressor and turbine operations and integrate them with Elliott. Ebara hopes to accelerate new product development through the integration of R&D activities, including development of microturbine products (not emphasized in this report). GAS Power Systems GAS Power Systems, LLC (Lake Zurich, IL) is a division of G.A.S. Capital Inc., which is a holding company of international aviation corporations. General Aviation Services is probably the best known of its current holdings. GAS Power Systems has packaged a 1.2-MW generator set based on an overhauled Avco Lycoming helicopter turboshaft gas turbine engine (Figure 4-3). The Innovator Genset package uses the T-53 engine, a Cotta Transmission gearbox, and a proprietary control system. The gas turbine has a speed of 6300 rpm. The gearbox, which provides a reduction of 3.32:1, is used to produce low-speed (1800-rpm) output. The T-53 engine has more than 40 years of service and is used extensively by the U.S. military. The engine is overhauled and packaged at a GAS production facility in Bonners, ID. The rebuild process includes modifications to fuel manifolds and combustion chambers, and new coatings are applied to the compressor and turbine sections.

Figure 4-3 GAS Powers 1.2-MW Innovator Genset

4-11

EPRIsolutions Licensed Material Providers

This unique product concept was pursued by GAS under the premise that the overhauled engine could be packaged and mass-produced more quickly and at a lower capital cost (projected by GAS to be $400/kW), much less than other products in the 1-MW size range. Initially targeting an exclusively international market, GAS has seen opportunities open up domestically due to deregulation and to what GAS considers to be an obvious gap in competitive gas turbine products in this size range. The first Innovator package is being testing in a project to accommodate peaking conditions in Lubbock, TX. The project is being conducted with participation from South Plains Electric Cooperative, GTI, and Texas Tech University. Honeywell Honeywell Engines and Systems is part of New Jersey-based Honeywell. In 1999, Honeywell Inc. and AlliedSignal merged. Honeywell Engines and Systems, headquartered in Phoenix, claims to be the worlds largest manufacturer of small and medium gas turbines. Its gas turbine applications are predominantly military, commercial, and business aircraft engines and auxiliary power units. Honeywells stationary gas turbines in the 300-kW to 5-MW size range the ASE8, ASE40, and ASE50 are aeroderivatives of popular turboshaft helicopter engines. These machines have both axial and centrifugal turbomachinery. The 548-kW ASE8 is a single-spooled gas turbine with a two-stage centrifugal compressor and three-stage axial turbine. The power section of the ASE8 is shown in Figure 4-4. The larger ASE40 and ASE50 gas turbines are two-spool models with a seven-stage axial compressor plus a single-stage centrifugal compressor.

4-12

EPRIsolutions Licensed Material Providers

Figure 4-4 Power Section of the Honeywell ASE8

Vericor Power Systems, a joint venture between Honeywell Engine Systems and MTU Aero Engines (DaimlerChrysler), was formed in 1999 to market and sell power generation sets with Honeywell gas turbines. Mitsubishi Nagoya Works (Japan) packages AlliedSignal gensets and mechanical drive systems in Japan. Another part of the Honeywell, Honeywell Power Systems, develops and manufactures microturbine packages and has been awarded a contract with the U.S. DOE under the Advanced Microturbine System Program. An advanced, high-efficiency system at the low end of the 300kW to 5-MW size range had been proposed. In October 2000, GE announced its planned acquisition of Honeywell, including both Honeywell Engines and Systems and Honeywell Power Systems. This acquisition will augment GEs already strong Power Systems and Aircraft Engine businesses. Kawasaki Gas Turbine The Kawasaki Gas Turbine Division of Kawasaki Heavy Industries produces standby generator sets in the 220-kW to 4-MW size range and cogeneration systems ranging from 640 kW to 30 MW. Kawasaki has over 5000 units installed worldwide. Standby units include a line of truckmounted mobile packages. Kawasaki states that gas turbine standby systems are popular in Japan due to their power density and resistance to damage in the event of an earthquake. Kawasaki has always been a leader in gas turbine research and development. For example, Kawasaki offered the first Cheng Cycle cogeneration system on the M1A-13 in 1988. By injecting superheated steam into the combustor of a gas turbine, electrical output is variable from 1.3 to 2.3 MW. In addition, process steam produced can be variable. A cross-section of an M1A-13CC gas turbine is shown in Figure 4-5.

4-13

EPRIsolutions Licensed Material Providers

Figure 4-5 Cross Section of Kawasakis M1A-13 Cheng Cycle

In 1999, Kawasaki demonstrated a 42.1% efficiency on a 300-kW ceramic turbine. This is the highest simple-cycle efficiency achieved. The M1A-13 was the test-bed engine for catalytic combustion testing by Catalytica in a long-term (8000-hour) test at a utility in northern California. The U.S. presence of Kawasaki has historically been accomplished through American packagers and distributors. There are over 170 units in North America. Kawasaki recently opened a U.S. office (Kalamazoo, MI) to market, package, and service Kawasaki gas turbine generator sets in the U.S. Nuovo Pignone Nuovo Pignone (Florence, Italy), a part of GE Power Systems since 1993, designs and manufactures equipment for the oil and gas, petrochemical, and power generation markets. Compressors, steam turbines, and gas turbines are its main products. Nuovo Pignones gas turbine product line ranges from 2 to 124 MW. The gas turbines in the size range of this report
4-14

EPRIsolutions Licensed Material Providers

are the 2-MW PGT2, the 5-MW PGT5, and the 5.9-MW PGT5B. These gas turbines are also sold and packaged by Stewart & Stevenson under the names GE2 and GE5. The PGT2 is a single-shaft machine with two stages of centrifugal compression, a single combustion can, and a two-stage axial turbine. It is used primarily in cogeneration (5 MW of thermal power available), mechanical drive, and emergency standby power applications. The PGT5 has a 15-stage axial compressor (9.1:1 pressure ratio) and a single combustion can; it comes in single-shaft (power generation) and two-shaft (mechanical drive) versions. The 5.9MW PGT5B (Figure 4-6) is significantly different from the two smaller engine designs. It is much closer to an aeroderivative than the previously mentioned heavy-duty PGT class turbines.

Figure 4-6 Nuovo Pignone PGT5B

The eleven-stage axial compressor has three stages of variable geometry stators, the combustor is annular, and the first stage of the turbine is cooled. It also comes in single-shaft and two-shaft versions. Nuovo Pignone is reported to be working with Catalytica on developing an ultra-low emission catalytic combustion option for the 5-MW PGT5. Nuovo Pignone also packages the LM aeroderivative product line and the frame gas turbines of GE. OPRA Optimal Radial Turbine OPRA (Netherlands) is developing plans to manufacture a line of completely radial gas turbines. Initiated by Jan Mowill, formerly of Kongsberg (Norway), OPRA has built and demonstrated its first gas turbine, the 1.6-MW OP16. Development efforts have focused on low emission technology development. Honeywell has licensed its low emission technology for utilization on the 10-MW ASE120. OPRA has plans to partner with Daihatsu Diesel (Japan) for manufacturing and packaging.

4-15

EPRIsolutions Licensed Material Providers

Orenda Orenda Aerospace (Ontario, Canada), an operating division of Magellan Aerospace Corporation, is a manufacturer of gas turbines for aviation and industrial applications. Its roots stem back to the former Hawker Siddeley. Orendas business is comprised of four elements: repair and overhaul of industrial and aviation gas turbines, component manufacturing, research and development, and the production of aviation reciprocating engines. The company precisionmanufactures parts for other engine manufacturers such as GE, Pratt & Whitney, and RollsRoyce. Orenda has packaged stationary gas turbine packages under the model names OT2, OT3, and OT5. Approximately 90 Orenda-designed industrial units remain in service today in electrical generation, pipeline and process gas compression, and oil pumping applications, with major customers worldwide. Orenda introduced the 2.5 MW-OGT2500 Power Generation Package in cooperation with SPE Mashproekt (Ukraine). Addressing a potential increasing demand for waste-to-energy technology, Orenda has modified the OGT2500 to fire biomass fuel derived from potential feedstocks such as wood or wood residues, lignin and cellulose, agricultural residues, cane trash, palm shells, straw and grasses, and waste paper. Orenda also supports other models of SPE Mashproekt gas turbines up to 27 MW. Pratt & Whitney Canada Pratt & Whitney Canada (Longueuil, Quebec), a subsidiary of United Technologies, has been one of the leading producers of small aircraft propulsion engines. It operates independently from Pratt & Whitney (Hartford, CT). In the mid-1960s, Pratt & Whitney Canada entered the market for industrial and marine gas turbines as a packager of large aeroderivative gas turbines (the Pratt & Whitney FT4 and FT8). By the late 1990s, market conditions led to the decision to leave the large aeroderivative packaging/project development business and concentrate on the Pratt & Whitney product line. The Industrial and Marine Division of Pratt & Whitney Canada has more than 600 units in the field worldwide for electric generation, cogeneration, gas compression, and marine applications. It markets its products internationally with a network of specialized packagers including Ebara (Japan), Nedalo (Netherlands), Tecnoergia (Italy), Bazan Motores (Spain), and several regional packagers in the U.S. Until 1998, UNC Industrial Power was the licensed U.S. packager of Pratt & Whitney Canada industrial gas turbines. The ST series of industrial and marine gas turbines are derivative of established PT6 and PW100 aircraft engines. The product line includes several versions of ST6 (500-kW), ST18 (1-2 MW), ST30 (3-4.9 MW), and ST40 (4-5.4 MW) gas turbines. The ST18 is shown in Figure 4-7. All of the gas turbines are derivatives of popular turboprop aircraft engines; have axial and centrifugal compression stages and free power turbines; and can be packaged for power generation, marine and mechanical drive systems. In power generation applications, a constant output speed is generally required. Because the engines incorporate a free power turbine, the high-pressure gas generator rotors are free to maintain their own optimum speed regardless of output power demand. As a result, the part-load efficiency remains high.

4-16

EPRIsolutions Licensed Material Providers

Figure 4-7 Pratt & Whitney Canadas ST18

In late 1999, Pratt & Whitney Canada launched development of a stationary, aeroderivative gas turbine, the 250-500 kW ST5, derived from the recently announced PW6xx aircraft engine. The new engine will be offered in both simple and recuperated versions for peaking and baseload applications. The new engine is called a miniturbine as it is aimed at power demands slightly largely than those offered by current microturbine manufacturers. In August 2000, Pratt & Whitney Canada announce it had signed an agreement with DTE Energy Technologies for the development of a 400-kW gas turbine package for the distributed generation market. The ST5 gas turbine will drive a high-speed generator supplied by TurboGenset (U.K.). The new package, called the ENT 4000, is targeted at medium-sized commercial and industrial customers and mini-grids serving new residential and commercial developments. The ENT 4000 is planned to be commercially available in 2002. DTE will be responsible for packaging and marketing the ENT 4000. TurboGenset has designed a unique, high-speed generator that is reported to weigh just one-tenth as much as traditional generators. The companys expertise has resulted in the development of power electronics that enable the high-speed generator to deliver power with less voltage variation and better harmonics than conventional designs. TurboGenset has an affiliation with Imperial College in London. DTE is investigating the potential development of other packages with non-gas turbine-based distributed generation technologies using the TurboGenset technology. Ramgen Power Systems Ramgen is a privately held company that has been engaged in the design, development, and testing of a power generation system based on ramjet technology. The company appears to be the only one currently conducting R&D with ramjets in stationary applications. Ramjets, which are internal combustion engines, have been used quite extensively in aerospace applications. In
4-17

EPRIsolutions Licensed Material Providers

particular, they have been used in linear flight applications such as missiles and highperformance aircraft. In the Ramgen engine concept, well-known compressible gas dynamics phenomena (oblique shock wave and related supersonic processes) are used to compress and expand the working fluid, rather than the traditional mechanical methods. In the Ramgen, the thrust of the ramjet is used to create a rotary force. The prototype engine uses a 6-foot diameter, tapered rotor approximately 9 inches thick at it outer diameter. The engines two thrust modules are mounted in opposition to each other on the perimeter of the enclosed rotor, which reaches localized speeds exceeding a Mach number of 2.5. This creates significant centrifugal forces and requires advanced aerodynamic design and highperformance materials. One potential advantage of the Ramgen is that it may not require a separate fuel gas compressor. The fuel and air mixture is compressed as it enters the thrust modules. Ramgen has benefited from continued support from the U.S. DOEs National Energy Technology Laboratory in proof-of-concept testing of the Ramgen engine. The potential to burn a variety of fuels (for example, coalbed methane) is one of the main drivers for DOEs interest. Until recently, all development had proceeded internally to Ramgen. Now that proof of concept has been demonstrated, Ramgen is soliciting partners, in particular companies with existing manufacturing, distribution, marketing, and service channels, to participate in commercialization of its product. Rolls-Royce Rolls-Royce (U.K.), a leader in aeroderivative gas turbines, offers gas turbine products ranging from 4 to 150 MW and also sells diesel engines down to 1 MW. Rolls-Royces current strategy has been to grow its global energy business by providing a full line of products and services for customers in power generation and oil and gas markets. This has been done primarily through acquisitions. Rolls-Royce acquired Allison Engine Company and US Turbine in 1995 and 1996, respectively. Rolls-Royce Allison produces industrial gas turbines in the 3-8 MW size range. US Turbine had been a packager of Kawasaki and Allison gas turbines. Now called Rolls-Royce Energy Systems, it packages the full line of Allison and Rolls-Royce aeroderivative gas turbines and reciprocating engines. In 1999, Rolls-Royce completed the purchase of the rotating products interests from Cooper Energy Services, part of Cooper Cameron Corporation. Rotating compression and control equipment is widely used in the oil and gas industry. The business, based in Mount Vernon, OH, is now part of Rolls-Royce Energy Systems. The 501K series of engines is the only one in the size range reviewed for this report. A 501Kbased electricity generation package is shown in Figure 4-8. The 501K engines are available in both single-shaft (power generation) and two-shaft (mechanical drive) configurations and range from 2.5 to 5 MW in power output. The most recent addition to the 501K line is the 501KB7. This is a boosted (high pressure ratio) version that produces just over 5 MW.

4-18

EPRIsolutions Licensed Material Providers

Figure 4-8 Rolls-Royce Allisons 501KB5 Gas Turbine Generator Package

Gas turbine development, design, and manufacturing is done at Rolls-Royce Industrial and Marine Gas Turbines Ltd. (Ansty, Coventry, U.K.) and Rolls-Royce Allison (Indianapolis, IN). Marketing and packaging are conducted by Rolls-Royce Energy Systems. North American Offices of Rolls-Royce Energy Systems are in Maineville, OH; Houston; Oakland, CA; and Markham, Ontario, Canada. Schelde Heron B.V. Schelde Heron (Netherlands) was formed as a joint venture between Heron Exergy and Royal Schelde, a Dutch industrial group, in 1996. Schelde Heron has focused on developing the 1.4MW Heron gas turbine, a two-shaft unit with intercooling, heat recuperation, and two-stage (reheat) combustion. The machine has reportedly achieved, at ISO conditions, an electrical efficiency of 42.9% with NOx emissions below about 10 ppm (20 g/GJ) at 15% O2. Relatively low turbine inlet temperatures and low pressure have resulted in simplicity of design, in that no blade cooling is required. A robust, conservative design is intended to reduce maintenance frequency and cost and increase availability and reliability. This is important because the cycle approach is notably innovative. The turbine consists of modular components that can be easily replaced. Installation of the unit requires a footprint of about 29 ft by 7 ft 7- in. (9.5 m by 2.5 m) for the complete system. The standardized package can be installed outdoors and is easily transportable, facilitating relocation if site economics change. Targeted applications include standalone power generation and combined heat and power. Schelde Heron plans to build five preproduction units and is soliciting investment partners. The company projects commercial availability in 2002.
4-19

EPRIsolutions Licensed Material Providers

Solar Turbines Solar Turbines, a wholly owned subsidiary of Caterpillar Inc. located in San Diego, CA, is the largest domestic producer of industrial gas turbines. It offers fully factory-assembled and tested gas turbine packages, which include gas compressor sets, pump-drive packages, and electric gensets. The electric generator sets range from 1.2 to 13.5 MW. Gas turbines in the size range reviewed for this report include the 1.2-MW Saturn, the 3.5-4.7 MW Centaur, and the 4.0-4.2 MW Mercury. In international markets, Solar works with licensed packagers including Tuma Turbomach in Europe and Niigata in Japan. The Saturn engine is the most widely used industrial gas turbine, with over 4800 units in 80 countries, primarily in oil and gas applications. The Centaur family of engines entered commercial service in 1968 and is pervasive in pipeline compression installations throughout the world. In 1997, Solar unveiled the recuperated Mercury 50 gas turbine (Figure 4-9). This is the first new centerline gas turbine designed by Solar in over a decade. The Mercury 50 was developed in cooperation with the U.S. DOE under the Advanced Turbine System Program. The recuperated cycle allows for 40% efficiency to be achieved. The Mercury is revolutionary in that it was designed primarily for intermediate-duty cycle operation. It has been called the only industrial gas turbine optimized from its conception for the distributed generation market.

Figure 4-9 Solar Turbines Mercury 50 Recuperated Gas Turbine

A Mercury 50 unit has been installed at the municipal utility of Rochelle, IL. Six to eight units will be installed by mid-2001.

4-20

EPRIsolutions Licensed Material Providers

With regard to R&D, Solar continues to invest in advanced component technology for its smaller gas turbines. In 1992, Solar introduced its low emission SoLoNOx combustion technology. SoLoNOx equipment is available for all Solar gas turbines but the Saturn. In 1997, Solar field tested the first industrial gas turbine equipped with ceramic components in place of standard metal components, which require internal cooling. The test engine is a Centaur 50 installed at an oil company cogeneration facility. Ceramic components, if proven durable, can allow for higher efficiency and power output. Currently Solar continues to invest in ultra-low emission technologies such as catalytic combustion. It has participated with Catalytica, PCI, and Alzeta in an advanced combustion development program. Stewart & Stevenson Stewart & Stevenson Energy Products, a subsidiary of GE located in Houston, provides power generation packages ranging from 2 to 50 MW. Its primary business is packaging GEs larger aeroderivative LM line of gas turbines. In the size range reviewed for this report, Stewart & Stevenson packages the Nuovo Pignone PGT2 and PGT5B gas turbines under the names GE2 and GE5 (see the previous section on Nuovo Pignone). Prior to being acquired by GE, Stewart & Stevenson had packaged European Gas Turbine/Ruston gas turbine models and Allison gas turbines in this size range. In their primary U.S. market, Stewart & Stevenson has seen much more market activity in its larger packages, specifically the LM6000, than the smaller packages that are the subject of this report. To meet a perceived growing need for larger packages, Stewart & Stevenson now packages a spray-intercooled version of the LM6000 designated as the LM6000 Sprint, which generates 47.3 MW. Turbomeca Turbomeca (France), part of the Snecma Group, is a world leader in small aviation gas turbines. It produces a complete range of turboshaft gas turbine engines, and the company packages industrial gas turbines for marine and land applications. Power generation applications include standby generation and combined heat and power. Turbomeca currently packages the RollsRoyce Allison 501KB5 and 501KB7 under the product names TM4000 and TM5000 (see the previous section, Rolls-Royce). Turbomeca partnered with Volvo (Sweden) and Ulstein (Norway) on the design of the 2.5-MW Eurodyn gas turbine engine, a two-spool unit used initially in high-speed marine applications. When integrated with a gearbox, it is suitable for electricity generation and combined heat and power applications. Vericor Vericor Power Systems, headquartered near Atlanta, is a joint company of Honeywell and MTU Aero Engines (a subsidiary of DaimlerChrysler Aerospace). Vericor markets, sells, and services gas turbines and genset packages using Honeywell gas turbines. Gas turbines currently offered in packages include the ASE8, ASE40, ASE50, and the ASE120 (see the previous section on Honeywell). The joint venture was formed to help ensure investment for product development and growth, leverage the technology competencies of Honeywell and MTU, and focus resources
4-21

EPRIsolutions Licensed Material Providers

on highly competitive and changing markets. In effect, the venture is intended to reduce risk to both Honeywell and MTU in highly competitive industrial power markets. Strategic partners associated with the venture include Mendenhall Technical Services (Fairfield, OH), Motoren and Turbinen Union Munchen (Germany), Industria de Turbo Propulsores (Spain), Mitsubishi Heavy Industries (Japan), Super Precision Heat Treatment Co. (Taiwan), Korean Aerospace Industry (KAI), AIDC Aero Engine Factory (Taiwan), and Shimadzu (Japan). Distribution partners include North American Energy Systems (Canada), MTU (Germany), Mitsubishi Heavy Industries (Japan), and Tiveni Engineering & Industries (India). North American Energy Systems designs and packages volatile organic compound (VOC) destruction packages. A VOC destruction installation is shown in Figure 4-9. This installation consumes output gases from carbon vapor deposition furnaces used to manufacture aircraft carbon brakes.

Figure 4-9 North American Energy Systems VOC Destruction Installation with Five Gas Turbine Units

Vericor recently started up a new ASE40 gas turbine-based cogeneration plant in Nevada. QDI INC developed the 3.1-MW plant. All steam and electricity generated from the cogeneration plant is provided to Quebecor Printing Nevada by QDC under a power purchase contract. The total energy efficiency of the plant is over 90%. It is unknown at the time of writing this report how GE acquisition of Honeywell will impact the Vericor joint venture. Yanmar Diesel Engine Co. Ltd. Yanmar (Japan) is a leading producer of diesel engines, which are used in all applications marine, industrial, mechanical drive, and power generation. In 1983, Yanmar developed and began marketing its AT product line of centrifugal gas turbines for standby power generation, which ranges from 280 kW to 2.4 MW.
4-22

EPRIsolutions Licensed Material Providers

The gas turbines are made in Japan and sold and serviced in North America through Yanmar offices and partners in Illinois, California, Florida, New York, and Vancouver.

4-23

EPRIsolutions Licensed Material

5
ECONOMICS & BUSINESS ISSUES
This chapter provides an overview of representative equipment, installed plant, and O&M costs for small gas turbines in the 300-kW to 5-MW size range. The cost information is based on published information and contacts with manufacturers, distributors, packagers, and project developers. Representative equipment costs are presented for both combined heat and power and standby applications. Equipment costs are presented as budgetary prices. As might be expected, these prices can vary considerably depending on competitive pressure among manufacturers, current market supply and demand conditions, and other factors. Installation costs are discussed in ranges of representative estimates. Finally, the cost impact of emission control equipment, specifically catalytic exhaust gas treatment methods, is presented. In the case of small gas turbines, this has negatively impacted project economics and could inhibit R&D in combustionbased pollution prevention technologies. Introduction Historically, small gas turbines in power generation applications have been used primarily in continuous-duty combined heat and power applications and to a lesser extent standby generation. Not included in this review of small gas turbine economics is power generation at oil and gas exploration and production sites. In those applications the footprint, power density, and capability to run on various fuels are factors in the equipment decision. Regulated Utilities The Previous Era In the history of power generation, a regulated utility owns generation and distribution systems and recovers its investment on fixed assets through the regulatory process. Traditionally, in the era of regulated utilities, there have been two markets for stationary gas turbines and, more generally, for power generation equipment: baseload and peaking. Older baseload plants that have been superseded by newer, higher efficiency plants traditionally served intermediate-duty applications. Peaking is a nearly separate market from the baseload and intermediate-duty markets, because peaking plants operate for the purpose of system reliability. Peaking power is generated for only a few hundred to perhaps a thousand hours each year, so peaking plant economics is dominated by equipment cost and recovery of investment, with little regard for efficiency. Traditionally, regulated utilities provided peaking capacity because of their obligation to serve and typically used their oldest, low-efficiency units for such service because these required the least investment. Because the economics of peaking plants is dominated by capital recovery, installed price is their most important economic attribute. Recovery of the investment in peaking plants is highly uncertain, as their use is largely determined by uncontrollable circumstances such as the intensity and duration of hot weather and the vagaries of the business cycle.

5-1

EPRIsolutions Licensed Material Economics & Business Issues

The emergence of distributed generation and microturbines at this time appears to be initiating an era of new technology and new applications. Electricity deregulation seems to be resulting in shortages and price spikes in peaking power. However, investment in peaking power generation is risky, depending on variables beyond the investors control, so it is not surprising that shortterm shortages, like transportation space during rush hour, will be very expensive. This shortage of peaking power is viewed by some as forcing consumers to think about the economic consequences of their use of electricity at instants of high demand. High peaking prices for electricity are the systems way of handling the equivalent of the market for freeway space during rush hour and for tickets to rock concerts or Super Bowl games. The economics of baseload plants is based on obtaining the lowest cost of power for the annual duration of operation expected from the unit. Plants move from baseload to intermediate-duty application when baseload demand grows and an additional plant is built with better economic performance. The usual case is that demand for intermediate-duty power has also risen, so that the old baseload plant becomes available for intermediate duty, with most of its investment already recovered. Such use of existing, older, baseload plants has been the most economic option for intermediate-duty operation. Deregulation The Present Era In the deregulated electric utility environment now upon us, power generation is like an a-lacarte menu. Purchasers can buy exactly what they want and need not buy the complete mix of baseload, intermediate load, and peaking that had been bundled together in historic utility rates. With modern meters and computer control via telephone and radio links, generators and certain loads can be turned on and off automatically as determined by business agreements. Customers with the interest and dedication to make the effort can reduce their power costs. Similarly for electricity suppliers, both utility and non-utility, deregulation allows them to provide power for specific applications. This could lead to investment in power generation equipment intended for intermediate-duty distributed generation, typically 1000 to 4000 hours per year. Small gas turbines are attractive candidates for such applications. Electricity can also be purchased from an energy service company, which, in turn, would own and operate distributed generation equipment. With rising interest in distributed generation and the advent of deregulation, the question of how to supply intermediate-duty power needs to be examined in light of the economics and competitive nature of a deregulated power supply. Power economics is local or regional, not national, because at present power can be economically moved only a modest distance. Often, constraints exist on individual power distribution lines and not on adjacent lines. Long-distance transmission lines were built primarily for power supply security. Long-distance transmission for economic reasons is a relatively recent development, and limited capacity exists for longdistance transmission of bulk power. The excess capacity in power generation equipment built in the 1970s and 1980s led to less power plant construction in the 1990s, so that the excess capacity could be economically used. This reduction in power plant construction, together with business preparation for electricity deregulation, resulted in current local shortages of peaking and intermediate-duty power. Thus

5-2

EPRIsolutions Licensed Material Economics & Business Issues

both peaking and intermediate-duty electric energy are commanding unusually high prices in several areas. With this in mind, the question arises whether or not it is a worthwhile investment to build intermediate-duty plants. The extent to which intermediate-duty plants (and even more so peaking plants) that supply power to the grid will be utilized by distribution utility customers is uncertain, and power industry investors disfavor such business uncertainties. In contrast, when a dedicated power generation unit is built to supply a well-defined load such as an industrial plant, office building, or shopping center, the investment uncertainty is greatly reduced. Power generation economics can appear differently to utility grid power providers and commercial and industrial onsite power consumers. When a cogeneration plant is economically advantageous, the electric utility sometimes offers to build, finance, operate, and maintain it. Uncertainty still remains regarding forced outage of the generation unit and the means and cost of backup power and utility grid interconnection. In a case where an intermediate-duty power plants installed cost is $500/kW, finance charges are 20%, and the plant is used 5000 hr/yr, the capital recovery on the plant amounts to 2/kWh. With a heat rate of 10,000 Btu/kWh (an efficiency of approximately 34% HHV), fuel costs amount to 4.5/kWh with natural gas delivered at $4.50/MMBtu (HHV). With an additional 1 cent/kWh for operation and maintenance, the total cost of electricity amounts to 7.5/kWh a modest amount compared to the cost of intermediate-duty power in some regions, but considerably greater than the cost of baseload power from a large gas turbine combined-cycle plant. This simple calculation makes the case for more exact economic analysis of directly serving intermediate-duty loads rather than letting them fall to the successful bidder in hourly supply auctions. Risk can be averted by contracting for the sale of the power, either for certain hours or for specified amounts of electrical energy, with appropriate provisions for use. In the arena of intermediate-duty power, gas turbines compete with reciprocating engines and the grid for the market. Various combinations of price and efficiency can compete. Lower cost, lower efficiency products are more attractive for peaking loads, while higher cost, higher efficiency products are favored at higher annual capacity utilization. Care must be taken to examine whether the unit will operate at constant or varying load, as simple-cycle gas turbine efficiency typically decreases at decreased load. Similarly, care must be taken to operate the gas turbine at power levels consistent with the manufacturers recommendation for a specific life. Operation at the high emergency and standby ratings can severely shorten gas turbine life. The cost of maintenance and periodic overhaul should be allocated over the operating hours at the corresponding power level so that operating and maintenance costs can be planned for, scheduled, and budgeted. Applications Combined Heat and Power Power generation systems create large amounts of heat in the process of converting fuel into electricity. For the average power plant, over two-thirds of the energy content of the input fuel is converted to heat and wasted. As an alternative, a customer with significant thermal and power needs can generate both thermal and electrical energy in a single combined heat and power system located at or near its facility. Combined heat and power (CHP), also called cogeneration,
5-3

EPRIsolutions Licensed Material Economics & Business Issues

can significantly increase the efficiency of energy utilization, reduce emissions of criteria pollutants and CO2, and lower the users operating costs. CHP is typically a baseload application with high annual hours of operation. Total fuel efficiency and operating costs are the primary economic parameters with baseload CHP. The outlook for CHP systems in the restructured environment has changed. Large installations depending on excess power sales (commonly referred to as PURPA machines, which resulted from the Public Utility Regulatory Policies Act of 1978) will have to compete with other wholesale generators that might have advantages in terms of dispatchability or cost, and competition may reduce retail electricity rates for large industrial users, decreasing the value of power generated onsite. At the same time, small to medium sized industrial plants and commercial/institutional facilities could see their electricity rates increase, increasing the value of CHP. With technology improvements and pending policy initiatives aimed at encouraging CHP due to its overall efficiency and environmental benefits, the customer base for economic, within-the-fence (onsite) CHP systems has the potential to expand considerably as an important subset of distributed generation. Standby Generation Some energy customers operate critical processes that are sensitive to electric service outages. These customers have standby generators to supply power when utility service is interrupted. Some standby installations are required by law to maintain public health and safety in such applications as hospitals, elevators, and water pumping stations. For other customers such as the telecommunications, retail, and process industries, the installation of standby generators is an economic choice based on the high costs of outages to their operations. Installed standby capacity in the United States is estimated to be over 40,000 MW. In contrast with CHP, standby applications require very few hours of operation per year. The primary equipment selection criteria are low capital costs and starting reliability. In this market, with short duration usage, low capital cost is much more important than operating costs and other variable costs. In certain areas, utilities recruit customers with standby generation for peak load reduction programs, offering payments or rate relief for limited operation during peak periods (typically fewer than 150 hours per year). In other areas, utilities are providing standby generators to customers for a fee, not only to ensure continued electric service during system outages but also to dispatch the generators for system peak needs. Customer choice among competitive power suppliers could stimulate economic preference for standby generators and increase the run-hours for units in the field. Standby generation can be part of an optimal customer strategy that minimizes power costs through combinations of firm and interruptible service and onsite standby capability. Intermediate Duty The evolving market and power consumption patterns reveal that in between high usage (CHP) and low usage (standby) extremes may lie a potential market of moderate size and intermediate annual usage. Intermediate duty is defined as approximately 1000-4000 hours of annual operation. Traditionally, intermediate-duty power requirements have been met by older central station power plants with lower efficiencies than newer plants. These older plants have higher variable costs and are dispatched after high-efficiency plants are fully deployed. Distributed
5-4

EPRIsolutions Licensed Material Economics & Business Issues

generation resources can compete in the intermediate-duty power market. Likely distributed resource technologies to play in this market include gas reciprocating engines, recuperated small gas turbines, and microturbines. The Solar Turbines Mercury 50 recuperated gas turbine with its high simple-cycle efficiency was developed primarily for this market. The costs for power vary by hour depending on demand and the availability of generating assets. Utilities see these variations in costs, but in the regulated market structure, customers typically do not. Large customers often pay time-of-use (TOU) rates that convert these hourly variations into seasonal and daily categories such as on-peak, off-peak, or shoulder rates. With the advent of wholesale and retail competition, more of these cost variations will be transmitted directly to the customer as price signals. Using distributed generation for intermediate duty could reduce the customers overall power costs as well as reduce the energy service providers need to generate or purchase very high-cost power. The more that the price paid for power is based on actual hourly costs, the greater the economic benefits of developing an intermediate-duty strategy, to both the customer and energy service provider. Depending on rate schedules and peak power costs, optimal intermediate-duty strategies could depend on distributed generation resources for 1000-4000 hr/yr. Competitive Positioning of Small Gas Turbines Gas turbines start to lose their clear advantage over other generation technologies at smaller sizes. The competitive positioning of small gas turbines relative to reciprocating engines is shown in Figures 5-1, 5-2, and 5-3. These figures clearly identify the competition that small gas turbines face from reciprocating engines and the unique positioning of the Solar Turbines Mercury 50 recuperated gas turbine.

5-5

EPRIsolutions Licensed Material Economics & Business Issues

Figure 5-1 Product Positioning Source: GTI

Gas Turbine Products


Simple Cycle Units
1000 Price ($/kW) Efficiency (%, LHV) 35 , Price Efficiency , 40

800 , 600

,, , ,, , , , , , , ,

30

,, ,, ,, , , , ,

400

25

200 0 5 10 15 20 25 Size (kW, thousands)


Actual purchase prices may vary due to market conditions and other factors Source: Gas Turbine World/GTI

20

Figure 5-2 Small Gas Turbine Products

5-6

EPRIsolutions Licensed Material Economics & Business Issues

1000 800 600 400 200

Price ($/kW)
CT Price RE Price

Size (kW, thousands)


Actual purchase prices may vary due to market conditions and other factors. Prices do not include gas compressors (if required). Source: Gas Turbine World/SFA Pacific/GTI

Figure 5-3 Capital Cost Comparison of Small Gas Turbines with Reciprocating Engines

In the industrial market segment (3.5-30 MW), worldwide sales of gas turbines and reciprocating engines are about equal. The size range below 5 MW is dominated by reciprocating engines, both natural gas and diesel fueled. As gas turbines decrease in size, they exhibit higher specific capital costs ($/kW) and lower efficiencies. For example, in the 2-5 MW range, gas turbines and reciprocating engines have comparable capital costs, but the reciprocating engines have substantially better efficiencies, between 37 and 42% (LHV), versus 32% or lower typical of comparably sized gas turbines. In the 2-MW and smaller range, reciprocating engines are priced lower than gas turbines and are considerably more efficient. One distinct advantages of small gas turbines over reciprocating engines is the amount of energy that can be recovered from the turbine exhaust stream. Higher quality recoverable energy, usually in the form of clean, high-temperature heat, allows for a wider range of thermal energy (for example, high-pressure steam) to be generated if needed. Consequently, many small gas turbines are currently deployed in CHP configurations where this recoverable energy can be used and higher total fuel efficiency can be achieved. Another distinct advantage of gas turbines is low emissions. Reciprocating engines lowest NOx emissions levels without exhaust treatment are currently two to six times the lowest levels achieved by gas turbines. Capital Costs The next sections provide representative costs for generator set equipment, switchgear, utility interconnection, emissions control, and installation. The cost components are representative and are intended to illustrate the approximate costs of components and installation. These figures might be useful in developing preliminary budget estimates; however, specific price and
5-7

EPRIsolutions Licensed Material Economics & Business Issues

performance information from equipment providers and specifying engineering firms is needed to develop site-specific plant cost estimates. Generator Set Equipment A summary of genset capital costs is shown in Table 5-1. The prices are for basic generator set packages with standard control and starting systems. These figures do not include switchgear, utility interconnection, back-end emissions control, or a fuel gas compressor if needed. The cost estimates do not include installation or the potential for price discounting.
Table 5-1 Representative Capital Costs for Small Gas Turbine Generator Sets Source: ONSITE Energy Gas Turbine Type
Industrial Industrial Industrial Industrial Aeroderivative Aeroderivative Aeroderivative

Size Range (MW)


0.2 1.0 1.0 2.0 2.1 3.5 3.5 5.5 0.4 1.5 1.5 3.8 3.9 5.5

Capital Cost ($/kW)


450 - 800 600 - 800 380 - 650 320 - 420 600 - 900 440 - 700 330 - 500

Notes

Turbine Cooling Turbine Cooling Single Spool Two Spool Two Spool, Turbine Cooling

The range of prices shown represents the spread in budgetary prices provided over the size range of that particular class of equipment. These ranges underscore the uncertainties in budget level pricing and the need to contact respective providers for actual price quotes. Switchgear and Utility Interconnection A summary of capital cost estimates for paralleling switchgear, transfer switches (in the case of grid-isolated standby units), and basic grid interconnection equipment is shown Table 5-2. Interconnection equipment includes protective relays, load synchronization with the grid, and integration with the genset control system.

5-8

EPRIsolutions Licensed Material Economics & Business Issues Table 5-2 Estimated Representative Switchgear and Interconnection Costs Source: ONSITE Energy Generator Set Output (MW) Automatic Transfer Switch ($/kW) Switchgear ($/kW) Utility Interconnection ($/kW)
* Low Voltage (less than 600 volts)

0.5* 20 60 60

1.0 100 40 40

2.0 50 25 25

5.0 20 12 12

Heat Recovery Equipment The cost of heat recovery equipment in of CHP systems varies significantly from site to site. This is due to the integration of the generation equipment into customer industrial processes or heating systems. Customer-specific engineering is required. Table 5-3 lists equipment and engineering costs for CHP installations in two representative sizes, 1 MW and 5 MW.
Table 5-3 Range of Heat Recovery Cost Estimates for Representative CHP Systems Source: ONSITE Energy 1 MW Gas Turbine Heat Recovery Steam Generators ($) Water Treatment Systems ($) Other Materials & Engineering ($) TOTAL COST IMPACT ($/kW) 240,000 300,000 50,000 60,000 100,000 150,000 390 510 5 MW Gas Turbine 300,000 400,000 140,000 160,000 300,000 400,000 148 192

Emission Control Equipment SCR (selective catalytic reduction) is a commonly employed emission control system for gas turbines where NOx emissions below 10 ppm are mandated by local air quality districts. Installation of such systems can be a significant cost impact, especially in smaller capacity gas turbines. For this reason, the cost of SCR systems is treated separately in this report. SCR costs have dropped considerably in the last two years, according a leading manufacturer, due to more efficient designs and lower design costs. Operating costs have also been reduced through innovations such as using hot flue gas to preheat ammonia injection air to lower the power requirements. Conventional SCR must be placed between sections of the HRSG so that the catalyst is not damaged by excessive exhaust gas temperature. The cost estimate shown below (Table 5-4) does not include the cost to retrofit the HRSG, since this cost is highly project- and design-dependent. These capital and annual costs are based on Cost Analysis of NOx Control Alternatives for Stationary Gas Turbines, November 1999, by ONSITE Energy for U.S. DOE.

5-9

EPRIsolutions Licensed Material Economics & Business Issues Table 5-4 Representative Capital Cost Estimates for SCR on Small Gas Turbine Systems Source: ONSITE Energy

Electric Capacity (kW) 1,000 5,000

SCR Capital Cost ($) n.a. $410,000

SCR Capital Cost ($/kW) n.a. $82

As shown in Table 5-4, SCR capital costs can add more than $80/kW to the unit capital costs of a 5-MW project representing 10-20% of the installed cost, depending on the specifics of the project. The cost impact is even greater for smaller gas turbine projects. Costs are so prohibitively high for SCR to be applied to a 1-MW system that at this time, no projects of this size have been installed or proposed with SCR. In a similar manner, costs to operate and maintain SCR systems can be a significant addition to the annual non-fuel operating budget, as shown in Table 5-5.

Table 5-5 Representative O&M Costs for SCR on Small Gas Turbine Systems Source: ONSITE Energy
Electric Capacity (kW) SCR Operating Cost ($) SCR Maint., Labor & Material Cost ($) SCR Electric Penalty ($) SCR Ammonia, Catalyst Costs ($) SCR Overhead, Insurance, Taxes Costs ($) SCR Total Annual Costs ($) SCR Total Annual Costs*

1,000 5,000 $15,000 $26,000 $11,000 $20,000 $40,000 $112,000

Not economic .0037

* Assuming 6,000 hours ($/kWh). Emission control regulations significantly influence technology development and capital costs for all gas turbines. As shown in this section and referred to in Chapters 3 and 4, the cost impacts of post-exhaust control systems are most severe for small gas turbines. For the past several years, gas turbine manufacturers have focused significant R&D on pollution prevention combustion approaches such as lean premix combustion as a method of complying with regulations cost-effectively. Recent gas turbine permit rulings by environmental regulatory agencies indicate that BACT (best available control technology) and LAER (lowest achievable emissions rate) could require postexhaust pollution control technologies (SCR) to be installed regardless of the emission levels produced by the gas turbine. This practice is starting to be applied to smaller and smaller gas turbine projects. If applied to gas turbines in the size range reviewed in this report, capital costs become prohibitive and can stifle market activity.
5-10

EPRIsolutions Licensed Material Economics & Business Issues

Gas turbine manufacturers recognize this and have openly questioned the value and costs of the next generation of pollution prevention technologies (such as catalytic combustion) if regulators will still require SCR regardless of gas turbine emissions. Given the trend toward higher turbine inlet temperatures (to improve efficiency) and the challenge of controlling NOx at these conditions, manufacturers might possibly discontinue advanced combustion R&D and accept the inevitability of SCR. Total Installed Costs In addition to the equipment costs identified above, complete project installation costs typically include labor, general facilities, engineering and fees, contingencies, and often overlooked soft costs. Soft costs include project development costs, permitting costs, legal costs, financing costs, and land costs. For the small projects considered for this report, these soft costs can be as high as 30-50% of the total installed cost if considerable development and permitting efforts are required. Operation & Maintenance Costs Non-Fuel O&M The O&M costs in Table 5-6 are based on gas turbine manufacturer estimates for service contracts consisting of routine inspections and scheduled overhauls of the turbine generator set. Routine maintenance practices include on-line running maintenance, predictive maintenance, plotting trends, performance testing, fuel consumption, heat rate, vibration analysis, and preventive maintenance procedures.
Table 5-6 Estimated O&M Costs for Representative Small Gas Turbine Systems Source: ONSITE Energy
O&M Costs 1 MW Gas Turbine 5 MW Gas Turbine

Variable (Service Contract) ($/kWh) Variable (Consumables) ($/kWh) Fixed ($/kW-yr) Fixed ($/kWh) Total O&M ($/kWh)

0.0045 0.0001 40 0.0050 0.0096

0.0045 0.0001 10 0.0013 0.0059

Routine inspections are required to ensure that the turbine is free of excessive vibration due to worn bearings, rotors, and damaged blade tips. Inspections generally include onsite hot gas path borescope inspections and nondestructive component testing using dye penetrant and magnetic particle techniques to ensure the integrity of components. The combustion path is inspected for fuel nozzle cleanliness and wear, along with the integrity of other hot gas path components. A gas turbine overhaul is typically a complete inspection and rebuild of components to restore the gas turbine to its original or current (upgraded) performance standards. A typical overhaul
5-11

EPRIsolutions Licensed Material Economics & Business Issues

consists of dimensional inspections, product upgrades and testing of the turbine and compressor, rotor removal, inspection of thrust and journal bearings, blade inspection and clearances, and setting packing seals. Gas turbine maintenance costs can vary significantly depending on the quality and diligence of the preventative maintenance program and operating conditions. Although gas turbines can be cycled, maintenance costs can triple for a gas turbine that is cycled every hour versus a turbine that is operated for intervals of a 1000 hours or more. In addition, operating the turbine over the rated capacity for significant periods of time will dramatically increase the number of hot path inspections and overhauls. Gas turbines that operate for extended periods on liquid fuels will experience higher than average overhaul intervals. Fuel Fuel costs are determined by the net electric efficiencies (heat rates) of the gas turbines. (Refer to Figures 2-12, 5-1, and 5-2 for the range of simple-cycle electric efficiencies for this gas turbine size range.) Most small gas turbine power generation installations operate at full load most of the time. If a unit is expected to run at less than full load during its planned operation, customer load profile data, as well as the part-load efficiency of equipment, should be obtained to estimate fuel costs. Various gas turbines have different part-load efficiencies and experience other performance degradations with ambient conditions. Note that manufacturers often quote efficiency based on LHV (the fuels lower heating value), while fuel consumption is normally calculated based on HHV (higher heating value). Cost of Electricity Using the representative estimates presented above, the cost of electricity was estimated for various small gas turbine systems as a function of capacity factor (Figure 5-4). Four representative gas turbine packages of different outputs are presented (500 kW, 1 MW, 4 MW, and 5 MW). The 4-MW system represents a high-efficiency recuperated-cycle turbine package. The other three configurations represent commercial, state-of-the-art, simple-cycle gas turbines. The calculation assumes a $4.50/MMBtu gas price, capital recovery over a 15-year life, and 10% return on investment. At the higher capacity factors (70% and above) for the 500-kW, 1-MW, and 5-MW systems, a CHP mode of utilization was assumed. The calculations assume a higher capital cost and credit the system with the value of thermal energy generated. In these cases, the value of thermal energy ranged from 2.5 to 2.8/kWh.

5-12

EPRIsolutions Licensed Material Economics & Business Issues

Cost of Electricity for Representative Small Gas Turbines


0.3500

0.3000

0.2500

500 kW 1 MW

COE ($/kWh)

0.2000

4 MW 5 MW

0.1500
500 kW w / HR 1 MW w / HR

0.1000
5 MW w / HR

0.0500

0.0000 10% 20% 30% 40% 50% 60% 70% 80% 90%

Capacity Factor

Figure 5-4 Cost of Electricity for Representative Small Gas Turbine Systems Source: ONSITE Energy

The calculations confirm the relative advantages of each of these systems. The high efficiency 4-MW product offers the most benefit in intermediate-duty applications. The other products are most competitive in high duty cycle CHP applications where their thermal energy offers value. In all representative cases, capital costs would need to improve for their utilization to be enhanced in low capacity factor applications such as standby generation. Unless small gas turbines can offer siting and environmental advantages, they are at a distinct disadvantage relative to other standby options. Market Outlooks CHP Market Positive: The potential for CHP is very large. The barriers of siting, permitting, and interconnection have all been identified and are likely to be addressed as part of industry restructuring. One of the most positive factors is the emergence of well-financed utility affiliates. These new energy service providers will have the ability, means, and, most

5-13

EPRIsolutions Licensed Material Economics & Business Issues

importantly, more motivation than regulated utilities to promote CHP and distributed generation as part of their overall customer service strategies. Negative: The CHP market has been stagnant in recent years due primarily to incumbent utilities ability to negotiate lower rates even in areas with high average retail prices. Utilities have been able to do this because prevailing power rates on the wholesale market up until recently have been relatively low. Customers have taken a wait and see attitude toward restructuring. In many states, restructuring has included competitive transition charges (CTCs). CTCs and exit fees are applied to customers that generate their own power, effectively locking customers into their current rates and usage until the transition period is complete. Finally, siting, permitting, and interconnecting CHP facilities continues to be difficult and expensive. Overall Assessment: The potential of for CHP is very large. Actual activity in this market will be stable or decline for the next few years until the previously identified issues related to restructuring are sorted out. After this transition period, the market is likely to expand, driven by an emerging energy service provider industry. Standby Market Positive: Historically, this is the largest single market for diesel-fueled reciprocating engine generator sets. Small gas turbine systems compete with diesel systems where they offer siting and environmental advantages. Negative: The vast majority of standby generator sets are diesel fueled. Diesel options have a sizable cost advantage for this application. Other diesel advantages that are well-suited for standby applications are quicker start-up to full load and onsite fuel storage. Onsite storage is a requirement for many health-care applications. Overall Assessment: Concerns about reduced power reliability will drive this market. It will continue to be dominated by diesel fuel for the most part, unless siting of diesel units become tougher due to environmental regulations. Intermediate-Duty Market Positive: Growth potential is high due to restructuring. The market will see a greater need to secure peak power. More movement toward time-of-day and time-of-use rates will increase the value of peak generation. Current markets exist in locations where

5-14

EPRIsolutions Licensed Material Economics & Business Issues

customers peakshave due to significant differences between on-peak and off-peak rates in published tariffs. Negative: While potentially large, this is a speculative market. Its growth depends on the evolution of price signals in energy rates that reflect the differential value between on-peak and offpeak electricity. Initially, the market will develop in some geographical areas sooner than others. This will present a service and distribution challenge if concentrated markets spring up throughout the country. If the on-peak/off-peak spread occurs over a short time period, lower annual hours of operation will favor lower capital cost systems, giving diesel engines an advantage. Overall Assessment: Generation for peaking purposes should grow as customers develop an overall energy strategy focused on minimizing total facilities cost and thus maximizing capacity utilization (hours of operation). Integration of gas-fueled generator sets into this strategy will favor high simple-cycle efficiency gas turbine systems due to their low operating costs and environmental advantages.

5-15

EPRIsolutions Licensed Material

A
Brief History of Power Generation and Gas Turbines
Power Generation History While the industrial revolution had its roots in the waterwheels of the Middle Ages, modern industrialization began to spread with the invention of the steam engine. In 1712, Newcomen built and operated the first (reciprocating) steam engine. This was of extremely low efficiency, as it was a (partial) vacuum engine that used atmospheric pressure to move the piston against the vacuum created by condensing steam into water in the cylinder by water injection. Additionally, the water used for condensation cooled the cylinder walls, thereby lowering the engine efficiency as steam was consumed in merely rewarming the cylinder walls. James Watt first invented the external condenser to increase engine efficiency; then in 1775, he invented the above-atmospheric pressure steam engine. With these two design improvements came major increases in efficiency and power density, as well as much profit to the firm of Bolton and Watt, which placed pumping engines in mines under business terms calling for payments to the engine providers based on the shared benefit of reduced fuel costs. In the mid-19th century, the science of thermodynamics developed the study of the conversion of heat to power in engines. In addition to steam engines, interest developed in the possibility of hot air engines. Indeed in the late 19th century, large river boats were built with hot air, externally fired (reciprocating) engines. This led to the Stirling engine, a small version of which was used by farmers for water pumping 100 years ago and was even featured in Sears Roebuck catalogs of the time. Interest developed in both steady flow and reciprocating hot air engines. Experimental reciprocating piston Brayton cycle engines were built, and a few are on exhibit at the Smithsonian Institution in Washington, D.C. In the late 19th century, as steam engines became larger and larger, Parsons invented the steam turbine, which first revolutionized naval ship propulsion, then power generation. For the next 50 years, power generation was dominated by the coal fired steam turbine system. The efficiency of this technology grew by a factor of 10 in the first half of the 20th century, culminating in an efficiency of nearly 40% HHV in the 1960s. Gas Turbine History Early Development After the steam turbine was scientifically understood, the same principles as were used for the design of the expansion turbine were applied to the design of air compressors. Soon thereafter, the steady flow turbomachinery Brayton cycle combustion turbine was built and operated, initially with the compressor consuming more power than the expansion turbine produced. This

A-1

EPRIsolutions Licensed Material

low (actually negative) performance was due to the poor efficiency of the compressor and expander and the low expander inlet temperature. The temperature limit on the expansion process was due to the limited capability of alloy steels that were commercially available at that time. Nickel based alloys, considered too expensive for commercial use a century ago, were later used to get over the temperature barrier. The first experimental gas turbine to operate at positive efficiency was built just after the turn of the prior century, about 100 years ago. Around the end of World War I in 1918, the first centrifugal compressor turbochargers were built for use on military airplanes. These continued to be used for aircraft and high performance automobile propulsion in the 1920s and 1930s and are in such use today. The high efficiency centrifugal (radial flow) compressor and the axial flow compressor (a reverse steam engine expansion type of device) set the stage for the modern gas turbine. The first practical gas turbines were built in the late 1930s and early 1940s. The first commercial machines were Swiss heavy-duty, oil-fired industrial turbines. At the same time, and in secret, Frank Whittle in England and Hans von Ohain in Germany built gas turbines as the source of pressurized hot air for use with an exhaust jet for aircraft propulsion. A startling breakthrough in airplane speed was then achieved in the early 1940s by the jet engine, which has a gas turbine as its core. Today gas turbines power turbojet, turbofan, and turboprop aircraft and helicopter and boat propulsion systems, and performance-enhancing turbochargers are in widespread use on reciprocating engines. These all perform on the same set of thermodynamic and fluid mechanic principles and have become ubiquitous. Most central station power plants recently built or expected to be built in the commercial planning future in the United States and western Europe are gas turbine combined cycle plants. This is because of their combination of low cost, high efficiency, very low pollutant emissions, and high reliability. Small Gas Turbines for Automotive and Other Vehicle Applications In the 1950s and 1960s, several major automobile companies experimented with small gas turbines, generally of the size now referred to as microturbines, for vehicular propulsion, none of which were commercialized. Some of the larger development units, originally intended for truck and bus propulsion, were produced as military aviation start carts and auxiliary generators. Government sponsored R&D on automotive gas turbines began with the energy crises of the 1970s and continued until the mid-1990s. Automotive gas turbines use regenerators rather than recuperators to recover heat from the turbine exhaust for preheating the compressed air flowing to the combustor. Regenerators can be more compact than recuperators and offer the opportunity to fit a high efficiency gas turbine under to hood of an automobile. Recuperators are the type of heat exchanger most often encountered in industrial practice. Recuperators transfer heat through fixed boundaries, as is the case with tube bundles and flow separation plates. Such plates may be corrugated, folded, or finned to increase the surface area exposed to the hot and cold gases. Regenerators are matrices of fine holes in high-temperature materials, usually ceramics, placed in the flow paths of the hot and cold gas streams such that the hot and cold gases pass through the passages alternately in opposite directions and periodically heat and cool the walls. In this manner, the walls act as temporary heat storage materials, with the hot stream first transferring heat to the walls and then, a short time later, having the walls give up this heat to the cold stream flowing in the opposite

A-2

EPRIsolutions Licensed Material

direction through the same passages. Entrance and exit manifolds duct the flows to the proper sections of the regenerator, which rotates somewhat slowly between the inlet and exhaust manifolds. While this may sound somewhat complex, in practice the flow geometry is straightforward. Regenerators are used in industrial heat recovery processes to preheat incoming air for combustion using the furnace exhaust gas as the source of otherwise wasted heat. The challenge in the use of regenerators is the seals that prevents leakage of pressurized air into the exhaust stream. This is of particular concern in gas turbines as the air stream is high pressures of 30-60 psig (3-5 atmospheres) in the case of automotive gas turbines and microturbines and 75195 psig (6-14 atmospheres) in the case of industrial and small aeroderivative gas turbines. As the regenerator rotates between the manifolds, a sliding contact between the seal and the regenerator results. If the seal is loose, compressed air leaks through the seal surface. If the seal is tight, it wears and subsequently leaks. To date, seal designs and materials have not been found even for the modest life (3000 hours or 100,000 miles at about 35 mph) demanded as a minimum for automotive applications. The most promising regenerator materials identified to date are ceramics, some of which have found application in industrial regenerators known as heat wheels. Microturbines Todays microturbines had origins in several fields including: Automotive gas turbine programs of the automobile and gas turbine industries of the 1950s Closed cycle gas turbine products of the Swiss and Germans (Brown-Boveri) before and after World War II Space power program of the U.S. National Aeronautical and Space Administration (NASA) in the 1960s-70s and the DOEs Office of Energy Efficiency automotive gas turbine development programs of the 1980s-90s. The microturbine is a low pressure ratio (3:1 to 5:1), recuperated gas turbine in the 25-75 kW size range. In the future somewhat larger versions, 100-250 kW, may be produced. Above this size, gas turbines are considered small rather than micro. Recuperators enables microturbines to reach efficiencies in the 25-30% range, which might make them sufficiently attractive for use in intermediate duty power generation. To compete, the installed cost of microturbines should result in power costs lower than those of the grid or of competing reciprocating engines. This means installed costs below $1200/kW to $500/kW, depending on local electric utility rates and the hours per year of use planned. In this service, the microturbine is expected to be ultrareliable, with appliance-like reliability, as down time for repairs would seriously compromise the attractiveness of the product and the overall economics. Much difficulty still exists in obtaining automatic acceptance of the electrical interconnection with the grid and for environmental siting permits. With the resolution of these issues and with lower installed costs, microturbines are expected to have major penetration of the market for electricity in the light commercial sector. Some microturbines have air bearings, which could reduce periodic maintenance. Microturbines with oil lubricated bearings do not exhibit any oil consuming process, so the life of the oil is limited only by evaporation, imperfect sealing, and long term chemical degradation. The life of both types of microturbines is expected to be competitive when manufactured with good quality control.

A-3

EPRIsolutions Licensed Material

B
Bibliography
Annual Survey of Engine Sales, Diesel & Gas Turbine Worldwide, October 2000. Barker, Thomas V., Ed., Turbomachinery International Handbook 1998. Norwalk, CN: Turbomachinery Publications, 1998. Bautista, P.J., Rise in Gas-Fired Power Generation Tracks Gains in Turbine Efficiency, Oil & Gas Journal, August 12, 1996. Collaborative Report and Action Agenda, California Alliance for Distributed Resources, January 1998. Distributed Generation Guidebook for Municipal Utilities, GRI-98/0025, January 1998. Dixon, S. L., Fluid Mechanics, Thermodynamics of Turbomachinery. Butterworth-Oxford, 1978. Freedman, S.I., Gas Turbines, in The CRC Handbook of Mechanical Engineering, Kreith, F., Editor-in-Chief. Boca Raton, FL: CRC Press, 1998. Gas Turbine Recuperators, GRI-99/0263, December 1999. Gas Turbine World 1999-2000 Handbook, Pequot Publishing. Hay, G.A., et al., Small Gas Turbines for Distributed Generation Strategies, Power Gen 95 Conference, December 1995. Industrial Applications for Micropower: A Market Assessment, DOE/ORE 2096, Resource Dynamics Corporation, January 2000. Liss, W., Natural Gas Power Systems for the Distributed Power Generation Market, PowerGen International 99, New Orleans, November 1999. Low Emission Combustor for Allison Gas Turbines, GRI-96/0439. Major, W., and Davidson, K., Gas Turbine Power Generation Combined Heat and Power Environmental Analysis and Policy Considerations. Prepared for DOE Office of Industrial Technologies and Gas Research Institute, November 1998.

B-1

EPRIsolutions Licensed Material

Major, W., and Powers, W., Cost Analysis of NOx Control Alternatives for Stationary Gas Turbines, DOE Contract No. DE-FC02-97CHIO877, November 1999. 1999 IGTI Technology Report and Product Directory, International Gas Turbine Institute, January 1999. Opportunities for Micropower and Fuel Cell Gas Turbine Systems in Industrial Applications, DOE/ORE 2095, Arthur D. Little, Inc., January 2000. Proceedings for the International Workshop on Catalytic Combustion, SRI International, GRI94/0501, April 1994. Reciprocating Engines for Stationary Power Generation, GRI-99/0271, December 1999. Rao, A.D., Bautista. P.J., and Francuz, V.J., An Evaluation of Advanced Gas Turbine Cycles, Power Gen 93 Conference, December 1993. Razdan, M.K., Bach, C.S., and Bautista, P.J., Field Experience of a Dry Low Emission Combustion System for Allison 501-K Series of Engines, ASME Asia 97 Congress, October, 1997. Sawyer, J.W., Ed., Sawyer's Gas Turbine Engineering Handbook. Norwalk, CN: Turbomachinery International Publications, 1985. Solt, C., The Ultimate NOx Solution for Gas Turbines, ASME 98-GT-287. Wilson, D.G., The Design of High-Efficiency Turbomachinery and Gas Turbines. Cambridge, MA: The MIT Press, 1989.

B-2

About EPRI EPRI creates science and technology solutions for the global energy and energy services industry. U.S. electric utilities established the Electric Power Research Institute in 1973 as a nonprofit research consortium for the benefit of utility members, their customers, and society. Now known simply as EPRI, the company provides a wide range of innovative products and services to more than 1000 energy-related organizations in 40 countries. EPRIs multidisciplinary team of scientists and engineers draws on a worldwide network of technical and business expertise to help solve todays toughest energy and environmental problems.
EPRI. Electrify the World

SHRINK-WRAP LICENSE AGREEMENT THIS IS A LEGALLY BINDING AGREEMENT BETWEEN YOU (LICENSEE) AND EPRIsolutions, INC., A DELAWARE CORPORATION (EPRIsolutions) PERTAINING TO THE MATERIALS THAT ACCOMPANIES THIS LICENSE AGREEMENT
1. DEFINITIONS For the purpose of this shrink-wrap license agreement ("Agreement") the following definitions shall apply and shall control over any inconsistent definitions of the same term used in any other contract or agreement associated to this Agreement. Materials: Means the document(s) and/or computer software accompanying this Agreement which is hereby designated by EPRIsolutions to be Proprietary Information as defined below. Target: Means an area of research work that created the Materials. Participant: Means an entity that funded the Target, or an entity related to the Participant, or the Participant's parent company, in which the Participant or its parent, owns a fifty percent (50%) or greater interest and which the Participant has designated as being funder of the Target that created the Materials. Proprietary Information: Means the Materials and any and all information or materials (including but not limited to scientific, technical, and business information, materials, concepts and information) disclosed in which the disclosing party has notified the receiving party is confidential, trade secret or proprietary. Internal Use License: Means LICENSEE's use (or use by a consultant acting for LICENSEE) of the Materials for LICENSEE's own business operations. Commercial Use License: Means a license for use of the Materials by LICENSEE for performing Consulting Services for LICENSEE's clients. This does not include the right to sublicense the Materials or any portion thereof, or to copy and distribute the Materials in violation of the copyright protections pertaining to said Materials. LICENSEE: Means an entity that accepts the Materials accompanying this Agreement under the terms and conditions of this Agreement. 2. LICENSE CLASS AND SUBORDINATION OF AGREEMENT The License Class for this Agreement and the Materials as specified above is in reference to the License Classes contained in the terms and conditions of the Target Funding Agreement entered between EPRIsolutions and the Participant, which is herein incorporated by reference if applicable. Furthermore, this Agreement shall be subordinate to the terms of any Target Funding Agreement or any other agreement entered between LICENSEE and EPRIsolutions that applies to the Materials. 3. GRANT OF LICENSE Accordingly having defined the License Class above, EPRIsolutions hereby grants to LICENSEE a nonexclusive, worldwide, nontransferable, perpetual, "for cause" revocable, Internal Use and Commercial Use License to the Materials accompanying this Agreement. LICENSEE hereby agrees that EPRIsolutions reserves its right to terminate this Agreement immediately on a "for cause" basis if LICENSEE fails to comply with any lawful material provision of this Agreement. 4. COPYRIGHT LICENSEE hereby acknowledges that the Materials accompanying this Agreement is owned by EPRIsolutions, is Proprietary Information, and is protected by United States and international copyright laws. Unless stated elsewhere in this Agreement, LICENSEE may not, without the prior written permission of EPRIsolutions, reproduce, this material, in any form, in whole or in part. 5. RESTRICTIONS LICENSEE may not rent, lease, license, disclose or give the Materials to any person or organization, or use the information contained in the Materials for any purpose other than as specified above, unless such use is granted via written permission from EPRIsolutions. LICENSEE agrees to take all reasonable steps to prevent unauthorized disclosure or use of the Materials and any and all Proprietary Information contained therein. Except as specified above, this Agreement does not grant LICENSEE any other intellectual property rights or licenses in respect of the Materials. 6. TERM AND TERMINATION This Agreement is effective until the termination of EPRIsolutions's rights in the Materials, or earlier on a "for cause" basis. LICENSEE may terminate the rights granted herein at any time by destroying the Materials contained herein. Upon any termination, LICENSEE may destroy the Materials contained herein, but all obligations of nondisclosure will remain in effect. 7. LIMITATION OF WARRANTIES AND LIABILITIES EPRIsolutions warrants that it has the right to grant the licenses and rights granted in this Agreement. However, EPRIsolutions assumes no responsibility for abating any infringement of its copyright or other proprietary rights in the Materials. Any action taken by EPRIsolutions with respect to any such infringement shall be at EPRIsolutions 's sole discretion. EPRIsolutions does not warrant the noninfringement by the Materials of any other copyright, patent, trademark, trade secret, or other intellectual property right, domestic or foreign. NEITHER EPRIsolutions, NOR ANY PARTICIPANT THAT FUNDED THE TARGET, NOR ANY PERSON OR ORGANIZATION ACTING ON BEHALF OF ANY OF THEM: (A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, WITH RESPECT TO THE MERCHANTABILITY, FITNESS FOR A PARTICULAR PURPOSE, OR NON-INFRINGEMENT OF ANY MATERIALS OR FREEDOM FROM CONTAMINATION BY COMPUTER VIRUSES, OF THE MATERIALS; OR (B) ASSUMES ANY LIABILITY WHATSOEVER WITH RESPECT TO ANY USE OF THE MATERIALS OR ANY PORTION THEREOF OR WITH RESPECT TO ANY DIRECT, INDIRECT OR CONSEQUENTIAL DAMAGES WHICH MAY RESULT FROM SUCH USE. LICENSEE agrees to comply with all applicable laws and regulations in the performance of this Agreement and the use of the Materials, including United States export control laws. The aforesaid obligations shall survive any satisfaction, expiration, termination or discharge of this Agreement or any obligations hereunder. EXCEPT AS PROVIDED OTHERWISE HEREIN, EPRISOLUTIONS PROVIDES NO INDEMNIFICATION TO LICENSEE. 8. INDEMNIFICATION FOR PARTICIPANTS ONLY For an amount up to the amount of funding the Participant contributed to the Target which created the Materials, EPRIsolutions shall indemnify, save harmless and defend the Participant against all claims of liability for infringement of a third party's United States copyright or unlawful acquisition of a third party's United States trade secrets which arises from the Participant's use of EPRIsolutions 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10414, Palo Alto, California 94303 USA 800.313.3774 650.855.2121 askepri@epri.com www.epri.com

About EPRIsolutions EPRIsolutions, a wholly-owned subsidiary of EPRI, provides R&D, technology applications services, consulting services, field test evaluations, and privately-sponsored initiatives to the power industry. Its areas of focus include power generation, transmission and distribution, end use technologies, market assessment and communications, facility maintenance programs, operator training, and online monitoring systems. EPRIsolutions offers a wide range of technical support to expand and enhance EPRI's overall science and technology program.

2000 EPRIsolutions, Inc. All rights reserved. EPRI is a registered service mark of the Electric Power Research Institute, Inc. Printed on recycled paper in the United States of America 1000768

Das könnte Ihnen auch gefallen