You are on page 1of 10

Handling Nine-Chrome Steel in HRSGs

Steam-Plant Industry Wrestles with Increased Use of P91/T91 and Other Advanced Alloys

By Robert Swanekamp, P.E., Contributing Editor


Nine-percent chrome-moly steels (9Cr-1Mo) have been used successfully in US fossil boilers as far back as the 1980s. Early pioneers included Tennessee Valley Authority, Dayton Power & Light Co, Appalachian Power Co, and Hawaiian Electric Co Inc. In recent years, the alloy (known as P91 in piping applications and T91 for tubing) also has been applied in large heatrecovery steam generators (HRSGs), in order to reduce thermal fatigue and creep damage in main-steam piping and superheaters. However, HRSG users generally have experienced more trouble with the alloy than their counterparts in the fossil-boiler sector have. Combined-cycle plants have been hit with problems in the fabrication, construction, and repair of P91/T91 components. In fact, HRSG users have experienced failures in dissimilar metal welds and transition areas in less than 1,000 operating hours, and failures caused by poor weld geometry or inappropriate heat treatment in less than 5,000 operating hours.

Primer on P91
The history of P91/T91 began in the late 1970s, when high-chrome alloys (9 percent to 12 percent Cr) were being studied by the U.S. government and Combustion Engineering (now Alstom) at the Oak Ridge National Laboratory. Researchers were trying to develop improved steels for the Liquid Metal Fast Breeder Reactor program. While the fast breeder reactor program in the United States never took off, the groundbreaking metallurgical work certainly did. Note: Oak Ridge continues its advanced-alloy research today, developing new structural and cladding material for so-called Generation IV reactors which include thermal and fast water-cooled, gas-cooled, and liquid-metal-cooled Fast Reactors. Other efforts hope to develop nickel-based alloys for ultra-supercritical coal plants (Figure 1).

Click here to enlarge image

Figure 1. Metallurgical advances have enabled the development of more efficient steam plants. Application of ultra-supercritical steam conditions in the next generation of fossil station likely will require the use of expensive, nickel-based alloys. Graph courtesy of Alstom Power Inc

The researchers in the 1970s found that 9Cr-1Mo steels possessed lower thermal expansion, higher thermal conductivity, and improved oxidation resistance, compared to traditional power plant steels such as 2.25Cr-1Mo ferritic steel and 300-series austenitic stainless steels.

These enhanced properties, the power industry quickly learned, would enable steam-plant components to be manufactured with thinner walls, thus they would minimize thermally induced stresses. With the addition of niobium, vanadium, and nitrogen, the standard 9Cr 1Mo (ASTM P9/T9) also exhibited a substantial increase in creep-rupture strength, compared to traditional steels, thus giving birth to what we know today as modified 9Cr -1Mo. Modified 9Cr-1Mo was certified in the 1980s as ASTM A213 Grade T91 for tubing and ASTM A/Sa 335 Grade P91 for headers and piping. Note that while piping and tubing applications are similar, there are subtle differences. In piping, the metal temperature will never exceed steam temperature, because steam is the source of heat. Thermal energy flows from the centerline of the pipe to the pipe"s outside wall. In a fossil boilers superheater and reheater tubes, in contrast, hot furnace gas is the source of heat, and thermal energy flows in the opposite directionfrom the tube"s outside wall toward its centerline. Under these conditions, the tubemetal temperature can be higher than the steam temperature. As a result, 9 percent Cr-steel can be used for piping applications up to steam conditions of 1,100 F. But for tubing applications, the operating steam temperature is limited to 1,050 F.

Kahe Unit 5
By the mid-1990s, P91/T91 was gaining acceptance in conventional fossil plants. Superheater sections made of T91 were retrofit at Tennessee Valley Authoritys Kingston Unit 5, while headers made of P91 were installed at Appalachian Power Co.s (now AEPs) G len Lyn and Clinch River stations and at Dayton Power & Light Co.s Stuart Station. Though these projects faced several challenges, the retrofits overall proved successful. A retrofit completed by Hawaiian Electric Co. Inc. (HECO) typifies the fossil indu strys experience in that decade. HECO"s Kahe station Unit 5 had suffered a series of failures of secondary-superheater tubes. A life-extension study proved the feasibility of replacing the entire secondary-superheater tube bundle and T91 was selected as the optimum material. Kahe Unit 5 is a Babcock and Wilcox Co. El Paso-style pressure-fired radiant reheat boiler with a continuous rated steam output of 965,000 lb/hr at 1,955 psig/1,005 F. The furnace is 27 feet wide, 23 feet deep, and 118 feet from the upper-drum centerline to the centerline of the lower-wall header. Commissioned in 1974, Kahe Unit 5 experienced its first secondary-superheater tube failure in 1988, and five more by early 1993. Five of the six failures occurred in the T22 material of the secondary-superheater inlet tubing. The sixth failure occurred in 304H stainless-steel material. All tube failures were attributed to creep damage from excessive metal temperatures. A contributing factor to one of the T22 failures was misalignment that exposed the tubes to the radiant heat of the furnace. Oil-ash corrosion resulting from the firing of high-sulfur oil earlier in the unit"s life also contributed to the failure of the 304H stainless material. A condition assessment conducted in 1989 showed some tubes in the inlet section had remaining useful life as low as 13,000 hours. HECOs refurbishment specification called for two types of bids: "replacement in -kind" and replacement of the T22 material with T91. HECO engineers understood that T91 would offer the following advantages:

Higher allowable stress for a given temperature Lower coefficient of expansion than stainless alloys Potential for improving plant efficiency by raising turbine-inlet temperature Potential for lowered thermal-fatigue cracking because of thinner walls

Importantly, HECO engineers also understood and paid attention to the potential drawbacks to using T91, which included:

Higher fabrication cost, because of the stress-relieving required after bending and welding, and the straightening and removing of oxide scales required after stress-relieving Concerns with quality assurance of the tubing from U.S. manufacturers with limited experience in 9Cr-1Mo steel Maintaining the design pressure drop through the secondary superheater with the thinner-walled T91 tubes

The successful bidder advocated replacing T22 in Rows 2 through 39 with T91, and using thicker-walled 304H in Row 1. Two of the most important features of the winning bid were the addition of "safe-ends" to the T91 and 304H tubes outside the furnace, and minor rearrangement of the 304H tubing. Together, these features reduced the number of dissimilarmetal welds inside the furnace to one per element total number dropped from 189 to 27. All "safe-ends" were radiographed in the shop during the fabrication process. Because of concerns about quality control and the lack of material availability in the United States, HECO bought its T91 material from Vallourec Industries" Power Generation Division (France), an experienced supplier to European plants where the new alloy had already been applied for several years. Field installation of the T91 secondary-superheater tubing was accomplished in five weeks with 30 people working seven days per week. Existing elements were removed with saws instead of torches to minimize contamination of the headers with foreign debris. The only T91 field welding required was to join the upper and lower sections of secondary superheater. These T91 field-welds were radiographed and stress-relieved according to the supplier"s stringent procedures and ASME Code. All welding documents and radiographs were reviewed and approved by the authorized inspector. Also, a steam-blow was conducted to purge the system of foreign debris prior to conducting a hydrostatic test at the drum working pressure of 2,300 psig.

Quality Assurance
HECOs attention to quality assurance was a key reason for Kahe Unit 5s successful retrofit. Thats also a key difference between the fossil-boiler and the combined-cycle experience, according to William F. Newell, Jr, EuroWeld Ltd.s vice president. Newell made his comments during a presentation at the P91/T91 Workshop, conducted in July 2005 in Philadelphia by the HRSG Users Group (see sidebar). In the fossil -boiler industry of the 1990s, Newell explained, The designers, fabricators and installers followed all the rules. They operated with conservative design margins (wall thickness), at reasonable steam temperatures (1,050 F or lower), and they carefully chose their steel producers and component fabricators. Newell emphasized that all Grade 91 welds regardless of thickness or diameter require a precise post-weld heat treatment (PWHT), and that welds between dissimilar metals should be minimized and properly designed. HECO, along with other P91/T91 users in the fossil-boiler business, applied these principles well. However, by the late 1990s and early 2000s, competitive pressures within the U.S. power industry and the building boom in new capacity pressed many in the combined-cycle sector to neglect some of the principles. Modified 9Cr-1Mo was seen by the booming combined-cycle industry as a silver-bullet remedy for two major troubles plaguing large HRSGs in cycling duty: thermal fatigue of thick-walled components, such as main steam piping and superheater headers, and creep damage in the superheaters (Figure 2). Modified 9Cr-1Mo is effective because the alloys mechanical properties allow pressure-containing components to be made in thinner sections, leading to smaller temperature gradients across the wall and reducing time for the metal to reach

thermal equilibrium, and ultimately resulting in less thermal fatigue. For example, the upgrade of a typical HRSG superheater header from P22 to P91 can reduce wall thickness by 54 percent, and component weight by 65 percent.

Click here to enlarge image

Figure 2. Most early-service life failures in HP superheaters, reheaters, and economizers have occurred at the attachment weld between an acute tube bend and the header, or occasionally at small radius bends used to connect tubes to the header. Photo courtesy of EuroWeld Ltd.

In addition, the alloys high creep-rupture strength and resistance to oxidation would minimize damage to superheater sections, particularly those coupled to the latest F- and G-class gas turbines that experience the highest metal temperatures. Unfortunately, as Newell reported to the HRSG Users Group, many of the owners and builders in the combined-cycle industry downplayed the concerns of the early fossil-boiler adopters, and handled the advanced alloy as if it were a traditional steel. It is anything but.

Focus on Microstructure
Modified 9Cr-1Mo alloy is an advanced material whose mechanical properties depend on the creation of a precise microstructure, and the maintenance of that microstructure throughout the components service life, explains Jeff Henry, director of Alstoms materials technology center. Henry also is chairman of an ASME Task Group assigned to study the material, and develop new standards for handling it. With traditional low-alloy steels like Grades 11 and 22 operating at the low stresses typical of power applications, the microstructure produced during steel-production and component fabrication was of only minor importance. In fact, for these materials it was found that a wide range of microstructures produced by both authorized and prohibited heat treatments would still provide satisfactory service. At the July Workshop, Henry emphasized how the superior properties of P91/T91 hinge on a precise addition of vanadium (V), niobium (Nb) and nitrogen (N), and a carefully controlled normalizing process to produce a complete phase transformation from austenite into martensite. This produces a hard steel with high tensile strength at elevated temperatures and high creep resistance. Next, a controlled tempering process must follow, to allow the V and Nb elements to precipitate as carbides and carbon nitrides at defect sites in the microstructure. This serves to anchor or pin the defect sites, thereby holding the microstructure in place. Henrys advice concerning the importance of the V and Nb additions and their precipitating action was backed up by a recent study in Japan. Conducted by Takashi Onizawa, and others, of the Oarai Engineering Centre, Japan Nuclear Cycle Development (Tokyo), the study was

presented at the ECCC Creep Conference, held in September 2005 in London by the European Collaborative Creep Committee. Until Onizawas recent work, the long -term efficiency and stability of the strengthening mechanisms provided by the fine particles had not been clarified. In the Onizawa study, a series of trial products controlling V and Nb contents were produced, and mechanical tests were conducted both before and after a 6,000-hour aging process. While the eventual goal of the study is to confirm the strengthening mechanisms under the fast breeder reactor operating conditions 1,112 F for 500,000 hours the current study clearly confirmed that higher strength and lower ductility were entirely dependent on the proper additions of V and Nb contents and the attainment of the proper microstructure. Failure to achieve that proper microstructure during original steel production or to maintain it during any subsequent action in the steels life such as the hot bending, forging or welding that regularly occurs during component fabrication, plant construction and component repairs at a power plant will cause a phase change away from 100 percent properly tempered martensite or will disrupt the precipitates. Either of these effects, Henry reported, will destroy the mechanical properties of the alloy.

Click here to enlarge image

Figure 3. The microstructure of P91 components has been damaged at some combined-cycle plants. A common error has been performing localized heating of the component with oxy-fuel torches, which are notoriously difficult to control. Photo courtesy of EuroWeld Ltd.

Unfortunately for HRSG users, the microstructure of P91 components has been damaged at some combined-cycle plants. A common error, as Newell explained during the July workshop, has been performing localized heating of the Grade 91 component with oxy-fuel torches (Figure 3). These are notoriously difficult to control and almost always provide destructive, non-uniform heating. Another common error has been following an incorrect procedure for PWHT temperature too high, temperature too low, or temperature not maintained for the correct duration. Even worse, some contractors are repairing P91 components without performing any PWHT at all.

ASME Code Enhancements


Many combined-cycle plant owners and operators assumed that merely specifying compliance with the ASME Boiler and Pressure Vessel Code in their new-construction and repair contracts would prevent such problems. However, the ASME Code rules are not comprehensive enough to address all of the material-processing demands and to prevent continued failures of P91/T91 components. In fact, Henry believes that the problems in large HRSGs will actually get worse unless more comprehensive Code rules are established and enforced. Thats why he is chairing a Task Group, working under the direction of the chairman of Section II (Materials) of the ASME Boiler & Pressure Vessel Code, to develop detailed rules for processing Grade 91,

as well as other approved steels classified as creep-strength-enhanced ferritic (Grades 23, 92, 122, etc). Henry and his team have identified eight specific issues concerning the use of creep-strengthenhanced ferritic steels, and are studying Code changes that will control their use more effectively. The Task Group has presented its recommendations for several of the eight issues to the Section II Committee of the ASME Boiler & Pressure Vessel Code. It is not clear when the Section II Committee will act on those recommendations with formal changes to the Code. In the meantime, Henrys Task Group continues to study the remaining issues it has identif ied regarding P91/T91 and other high-strength ferritic steels. At the HRSG Users Group workshop, the issues were discussed in detail between Henry and the many P91 authorities participating in the meeting. A few are highlighted below.

Post-Weld Heat Treatment


A second issue being addressed by the Section II Task Group is the effect of weld fillers on PWHT. Certain alloying elements in the filler principally nickel and manganese depress the AC1 and AC3 temperatures, as well as the martensite start (Ms) and martensite finish (Mf) temperatures. During PWHT, this presents the risks of intercritical heat-treating damage and untempered martensite in the weld metal. Henry noted that the AWS standards allow up to 1 percent Ni in weld metal, in contrast to 0.4 percent Ni maximum in base metal specifications. In response to this issue, the Task Group has proposed new PWHT limits on Grade 91 components based on Ni plus Mn content. Specifically:

PWHT temperature range must be 1,350 F to 1,425 F, if chemical composition of the filler metal is not known precisely If the chemical composition of matching filler metal is known precisely, the maximum PWHT temperature can be increased to 1,470 F (if Ni + Mn < 1.00 percent) or 1,450 F (if Ni + Mn is between 1.00 percent and 1.50 percent) For components less than or equal to five inches thick, minimum PWHT time must be one hour per inch with a minimum PWHT time of 30 minutes For components greater than five inches thick, PWHT must be five hours, plus 15 minutes for each inch over five inches For weld thickness less than one-half inch, minimum PWHT temperature is 1,325 F

Intercritical Region/Tempering
One of the most significant problems with Grade 91 is post-production exposure to temperatures in the intercritical region. This is above the temperature where martensite begins to transform back into austenite (referred to as the lower critical transformational temperature or AC1) and below the temperature where phase transformation is complete (called the upper critical transformational temperature, AC3). When Grade 91 is exposed to this intercritical region, the martensite is partially re-austenitized and the carbon-nitride precipitates are coarsened but do not fully dissolve back into solution. The resulting material is a part austenite/part martensite metal that lacks the pinning effect of the precipitates, and therefore has a substantially reduced creep-rupture strength. Exposure to the intercritical region and the resulting reduction in strength leads to the Type IV cracking found in many P91 welds. In a Type IV failure, cracking takes place in the fine-grain section on the base-metal side of the heat-affected zone of a weldment. Abrupt changes in wall thickness or other features that create high stresses in the region of the weld set up the conditions necessary for this cracking. Type IV failures are a matter of significant concern because they occur at a relatively early stage in component life 20,000 to 40,000 hours

at lower operating temperatures than the maximum design temperature of 1,110 F, and they can initiate and grow sub-surface for some distance before breaking through to the surface. There have been about a dozen such failures in P91/T91 components, mostly in the U.K. where the alloy has been in service longer than in the US, but we may be poised to catch up. One U.S. user at the workshop reported several P91 weld failures in steam piping of his companys relatively new fleet of F-class combined cycles. A related problem is over-tempering, which occurs when P91 or T91 components experience prolonged exposure to elevated temperatures below the lower critical transformation temperature. This does not affect the martensite, but it does cause coarsening of the precipitates, with a corresponding loss in creep-rupture strength due to the loss of their pinning effect. Over-tempering is a lesser risk during fabrication, Henry explained, due to the relatively short times of the thermal treatments. But in cases where multiple heat treatment cycles are applied in the fabrication of thick-walled components, weakening could become a significant issue at higher tempering temperatures.

Click here to enlarge image

Figure 4. Incorrect at the steel mill was the cause of this failure in a Grade 91 component. In addition to a loss of creep-rupture strength, risks associated with improper tempering are brittle fracture and stresscorrosion cracking. Photo courtesy of Alstom Power Inc

Under-tempering also can jeopardize the high-temperature properties of Grade 91, since the required precipitation does not go to completion, and the precipitates either are absent or are of insufficient size to stabilize the structure (Figure 4). In addition to a loss of creep-rupture strength, risks associated with under-tempering are brittle fracture and stress-corrosion cracking. To avoid intercritical-region exposure, over-tempering, and under-tempering, Henrys Task Group is recommending several rule changes to the ASME Code. Specific limits being proposed are:

1,900-1,975 F for normalizing 1,350-1,470 F for tempering 1,325-1,470 F for PWHT of components thinner than 0.5 inch 1,350-1,470 F for PWHT of components thicker than 0.5 inch For any component in which a portion of the component is heated above 1,470 F, the component would have to be re-normalized and tempered in its entirety, or as an

alternative, the heated portion could be removed from the component for re-normalizing and tempering and then replaced into the component.

Hardness Testing
Another of the issues that Henrys Task Group is tackling is quality -assurance testing. To determine whether the processing of creep-strength-enhanced ferritic steels has been performed correctly, users need a non-destructive evaluation (NDE) tool that can quickly and inexpensively provide information on the overall condition of the material. Because hardness provides a direct indication of a materials room-temperature tensile strength, which can be used to roughly estimate the elevated-temperature behavior of the material, portable hardness testing has been considered as such a tool. Although commercially available portable indenters have been available since the 1920s, hardness testing equipment really is designed for use in laboratories, not in the field. With portable equipment, there is substantial variability in hardness test results. Whats needed is a rugged attachment for indenting devices that can take accurate hardness readings at a power plant thats on-line and operating at temperature. Until such a tool gets developed, the ASME Task Group has not yet been willing to recommend specific hardness limits on advanced alloys. A leading HRSG manufacturer, NEM b.v. (Netherlands) feels more confident in establishing some guidance in this area. The companys materials and welding engineer, Ing. Patric de Smet, IWE, speaking at the July workshop, asserted that if the Grade 91 material has received proper heat treatment, hardness test results will fall in a relatively tight range not too high and not too low. NEMs guidance on PWHT calls for a temperature range of 1,380-1,420 F, and a duration of two to three hours. If this guidance is followed, de Smet said, the hardness will drop to 200-270 VHN (Vickers Hardness Number) and the ductility and high-temperature strength will be suitable for service.

Determining Creep Damage


A different NDE tool was discussed at Septembers ECCC Creep Conference, by several of de Smets fellow countrymen. Harry J. M. Hulshof and Paul G. M. Welberg pf of KEMA Nederland b.v. (Netherlands) teamed up with Leo E. de Bruijn of E.ON Benelux Generation n.v. (Netherlands) on a paper titled Creep Strain Measurements for Risk Based Monitoring of Steam Pipes and Headers. In the paper, the authors presented information on KEMAs Speckle Image Correlation Analysis (SPICA) system, which enables users to measure the deformation caused by creep in critical areas of their piping and headers such as the heat-affected zone in weldments. The technology can be applied to plants that are on-line and operating at temperature. According to the authors, plant managers can use the strain measurements as an indication of creep rate, from which you can assess how much creep life has been consumed, and by extension how much service life remains in the plant component. Note: Creep is a collection of diffusion processes driven by temperature and mechanical stress, which cause permanent deformation that can be measured as strain. The technique essentially involves making an optical fingerprint of the surface of a component on the basis of textural features. A recorded image of a rough surface is used as a "fingerprint" of the object surface. Two images, one recorded before and one after loading, are compared by image correlation. By analyzing the results, the strain distribution due to the loading is calculated from relative displacements in two directions. KEMAs strain-measurement technology has been applied at several Dutch power plants, the authors reported, to monitor strain during the last seven years. One of them is E.ON Beneluxs coal-fired plant at Maasvlakte (Rotterdam), Units 1 and 2 (540 MW each). Provisions for both SPICA and capacitive sensors were mounted. The locations that were monitored are bends (inner and outer curve), welds in the saddle point of T-joints, Y-joints and circumference welds in straight pipes. The utility company uses the results to plan activities during the next

overhaul and to build a strategy for condition-based maintenance. In the past there have been major overhauls at Maasvlakte every two years. This interval now has been extended to four years.

Creep-Data Assessment
Our understanding of creep is advancing steadily (no pun intended). For example, the European Collaborative Creep Committee (ECCC) was founded in 1991, and today includes nearly 50 organizations from the United Kingdom, Germany, Italy, France, Switzerland, Austria, the Netherlands, Belgium, Sweden, Denmark, Finland, Portugal, The Czech Republic and Slovakia. Companies participating in ECCC cover a wide range of industries and organizations dealing with high-temperature metallurgy: steel makers; boiler and turbine manufacturers; powerplant architect/engineers; power producers; inspection agencies; research institutes; technical universities; and testing laboratories. One of the committees achieved results according to a paper presented by G. Merckling, Istituto Scientifico Breda, Milan, Italy, at the ECCC Creep Conference, conducted September 2005 in London is the development of an accepted evaluation method to understand and to weight the quality of creep-data test results.

Click here to enlarge image

Figure 5. Collaborative efforts in Europe aim to develop one accepted evaluation method to understand and to weight the quality of creep-data test results. Diagram courtesy of Istituto Scientifico Breda.

For years, different creep-strength authorities have applied different evaluation methods, each one resulting in quite different creep-strength and life-assessment predictions using the very same test data. In contrast, the ECCCs evaluation method dubbed the Post Assessment Tests (PATs) yields consistent creep-strength predictions, which according to the committee will be suitable for inclusion into European standards (Figure 5). With only minor variations for

various assessment conditions, Merckling reports, the PATs have proved to be sound and reliable. In addition, mechanisms in the PATs allow for upgrades and additions in the future, so that its application can be expanded to relaxation strength, small and weldment data sets, post-exposure remnant creep strength, and creep strain models. And with the recent development of an Excel-based tool, referred to as the ePAT, the program is aut omated for quicker and more user friendly application.

Other ASME Code Changes


While Henry and his Task Group are working to revise Section II of the ASME Code (Materials) to improve fabrication and repair practices for creep-strength enhanced ferritic steels, two of his Alstom colleagues are working to revise Section I of the Code (Power Boilers). Their goal is to improve the economics for next-generation fossil units. Speaking at the ECCC Creep Conference in September, Alstoms I. J. Perrin and J. D. Fishburn pointed out that the push to advanced cycles with ultra-supercritical steam conditions will require the use of very expensive, nickel-based alloys. New design equations are needed in Section I of the Code, Perrin and Fishburn believe, to better account for secondary stresses during thermal transients (cycling of the plant). Better design rules will enable designers to reduce the wall thickness of tubing, as well as that of headers and piping, without compromising component reliability or safety. This will translate into a significant saving of expensive material, and thereby improve the economics of advanced ultra-supercritical boilers, the authors believe. Perrin and Fishburn estimate that in ultra-supercritical steam generators a 12 percent reduction in the cost of pressure parts can be achieved. Proposed revisions to Section I of the ASME code to permit the use of simplified and more technically defensible design equations were submitted to Subcommittee I and accepted by them on September 2, 2004. Subsequent to that, they were included in the Main Committee ballot. Feedback is awaited.

HRSG Users Gathering in the Rockies


William F. Newell, Jr, will be addressing the HRSG Users Group again, this time at the groups 2006 annual conference. Newell, who is vice president of EuroWeld Ltd, will present Welding Techniques for a Successful Plant Outage. Other presentations at the conference include:

The Risks to HRSGs of Low-Load Operation, by Scott Wambeke, Systems Engineer, HRST Inc New Technologies for HRSG Tube Repair, by David W. Gandy, EPRI; and Ken Brazell, Encompass Machines Inc Safety-Valve Testing & Maintenance, by Robert Pabst, Valve Design & Maintenance Consultant, Movaco Inc

The HRSG Users Group 14th Annual Conference will be held March 13-15, 2006, at the Broadmoor Hotel in Colorado Springs, Colo. For details, visit www.hrsgusers.org/events.php or call 1-718-317-6737. Power Engineering February, 2006 Author(s) : Robert Swanekamp

You might also like