Vous êtes sur la page 1sur 12

OTC 23943-MS Design Guideline Strategies for HPHT Equipment

H. Brian Skeels, Kwok Lun Lee, Anand Venkatesh, FMC Technologies

Copyright 2013, Offshore Technology Conference This paper was prepared for presentation at the Offshore Technology Conference held in Houston, Texas, USA, 69 May 2013. This paper was selected for presentation by an OTC program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been reviewed by the Offshore Technology Conference and are subject to correction by the author(s). The material does not necessarily reflect any position of the Offshore Technology Conference, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Offshore Technology Conference is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of OTC copyright.

Abstract As the need for oil and gas equipment working in hotter and higher pressure environments continues to mount, the effort to come up with an adequate set of design, material, and validation practices continues to be challenging. The higher pressures, and therefore thicker wall sections, have pushed traditional thin-wall pressure vessel design practices to theoretical limits, where their design model assumptions are becoming increasingly inaccurate in predicting design stress or present overly conservative hardware designs that are too thick or complicated to fabricate; not to mention safely pressure test in the factory. Manufacturing processes and material mechanical properties are also being pushed well beyond general practice to meet more stringent ductility requirements and to resist aggressive corrosion environments. Historically, designers simply overdesigned the equipment to provide a robust product with liberal design margins to hopefully out-muscle any unforeseen or complicated set of loading conditions. However, for the newer HPHT oilfield equipment, overdesign is less and less an option. Now the most challenging design failure mode is shifting towards fatigue and fast fracture as the equipment is taxed by very high internal pressures, while increasing temperatures are decreasing material strength. As a result, more rigorous stress analysis and design methods that more closely model fast-fracture burst conditions (rather than leak-before-burst failure) are needed to achieve safe, reliable and cost effective equipment designs. However, the effort to fully embrace the change to more rigorous design methods has been arduous. The Industry is just entering the transition zone between conventional and HPHT environs, and pressure vessel code "demarcation" on pressure ratings overlap in this region. Therefore a lengthy debate has ensued as to where does one draw the line between traditional "quasi-static" (infinite fatigue life) thin wall pressure vessel design practices and begin using a Lam thick-walled model with fatigue and fast-fracture failure (fracture mechanics) design practice. Both have equally valid arguments in this transition region from 10 000 psi at 250F to 20 000 psi at 400F (69 MPa at121C 138 MPa at 205C). Some of the debate on where the line of demarcation is drawn is fueled by the fear that when one crosses that line, then fatigue/fracture mechanics will drive design calculations, material performance and testing, and validation testing exponentially beyond what is expected today. Although this Rubicon will eventually have to be crossed as HPHT conditions tax designs and materials, this paper offers some insight into some simple tests to augment HPHT design practices, building on the insights presented in OTC paper # 23621.

Introduction The path toward designing oilfield High Pressure High Temperature (HPHT) Wellhead and Tree equipment is slowly crystallizing as Industry and Regulators show increased interest and angst in developing and reviewing pending projects and approving investments in long term assets. There are three significant roadblocks in the path to a workable solution: no clear HPHT design methodology, lack of understanding of what tests will validate designs, and very little published material properties data at HPHT conditions. And the scope of all three is exacerbated over the debate whether fatigue and fracture mechanics should be part of the equation. As oilfield equipment is subjected to more severe and longer periods of cyclic forces (internally from irregular wellbore flowing conditions, or externally from metocean conditions), many agree that fatigue life and monitoring for these conditions has to be a part of the design equation. And this fact is mentioned in the ASME Boiler and Pressure Vessel Code recognizing the two failure modes of a pressure vessel: leak-before-burst, or a fastfracture burst. The second failure mode vexed the engineering community until the advent of fracture mechanics and better theories of crack initiation, growth and eventual plastic collapse through the wall of the pressure vessel.

OTC 23943-MS

Many oil field design guidelines such as API 6A, 16A and 17D, negated a need for fatigue analysis because of the long standing practice of keeping maximum design loads and stresses well below 2/3 material yield strength; a practice which also keeps cyclic loading issues well inside the "infinite" region on S-N curves (in most cases). Unfortunately, this "quasi-static" design philosophy grows less accurate as the pressure vessel's wall thicknesses grows or geometries become more complex when addressing HPHT conditions (beyond the traditional cylinders or spheres described in classical engineering physics). As a result, many have turned to the ASME Boiler and Pressure Vessel Code for guidance. The code increases its sophistication of design practices as the art (practical experience and case studies), computational level (elastic and elasticplastic FEA, etc.), and new failure theories (such as fracture mechanics) are recognized; divisions 2 and 3 being alternate design methodologies as internal pressure and temperatures increase. But the designer, using the code, is still free to select which division to follow for his application, working within the overlap from one pressure and temperature realm to the next (all within and using "sound engineering judgment"). API has followed a long standing practice of following a subset of ASME Division 2, but as a result of API's work within 6HP and PER15K, they now recognize many shortcomings as one moves into HPHT conditions. However the published technical report only cites the current state-of-the-art and discusses the issues to be tackled. It became apparent to the task group that the scope could not provide a single general solution or path forward for all oilfield equipment. The authors of this report recognized that each hardware group has its own set of geometries, loading conditions, and environment in which they reside, and so left the details to the API technical subcommittees to resolve. Many rightfully argued that since OCTGs are suitably concentric and located in a stable temperature environment (downhole) then the current API practice is applicable for both standard and HPHT. Others argue that Wellhead and Tree equipment needs to use more rigorous design practices as HPHT equipment geometry becomes more complex and/or the environment extremes (between shut-in and production conditions) increase for HPHT service. Yet everyone agrees that leaving it to "sound engineering judgment" would result in many variations of opinion between operators and manufacturers, and this would likely result in a design and validation "do over" for every application. So the Industry continues the search for an agreeable HPHT guideline. Current efforts are addressing two options: either extending current API design practices to higher pressure ratings, or establish a demarcation when different design practices are needed. Two facts are slowing progress toward a design philosophy solution for Wellhead and Tree equipment. The first is the fact that ASME Boiler and Pressure Vessel Code "recommends" a switch from Division 2 to 3 as the pressure rating increases beyond 10 000 psi (69 MPa), but still allows the designer the choice to to use either one. Several current API specifications set the published upper limit between 15 000 or 20 000 psi (103.5 138 MPa), further muddling the problem as to where the line is drawn. The second is the fear that if one "selects" Division 3, then by default, they have to embrace fatigue/fracture mechanics in their design, forcing them to use the limited set of ASME materials listed in the code. Many of the listed materials are not readily used or understood in oil field environments. Once the designer crosses from Division 2 to Division 3, the scope and duration of design calculations, qualifying materials with mechanical performance at higher temperatures and more corrosive environments, and the number of validation tests and individual test times of HPHT hardware would increase exponentially. However, this too is not a hard and fast rule; again ASME leaves the choice between Div 2 vs Div 3 as a recommendation to the designer, with no hard and fast rules. Several papers which discussed the advantages and limitations of established Pressure Vessel design methods suitable for HPHT have been published and presented in previous conferences. These methods are not readily intuitive or straightforward. So this paper also addresses the challenges faced by HPHT design solutions and offers a test as to when a more rigorous method should be followed over a simpler more conservative approach. A second test is also proposed; assessing whether cyclic loading analysis is warranted, independent of the design methodology. The proposed test is based on two fracture mechanics principles used to create a "representative" S-N Fatigue curve. The test allows the designer to check his static design against a maximum/minimum stress criteria for the material the designer has selected with only a rudimentary understanding of fracture mechanics.

Pressure Vessel Design API 6A (which is followed as the design calculation method for API 16A and 17D) describes two methods of verifying the design of bodies, bonnets and end connections: Linear Elastic Stress Analysis and the Distortion Energy Theory. Linear Elastic Stress Analysis involves manual calculations or computer analysis to determine stress levels, followed by stress categorization and comparison to allowable stress limits to determine if a particular design is sufficient. However, ASME VIII Division 2 itself spells out the limitations of this method. It goes into great detail to explain that this method should not be used for heavy-wall pressure vessels as it can yield non-conservative results. It also states that this method is not accurate for complex geometry or complex loading. ASME VIII Division 3 limits the use of this method to thin wall construction (R/t > 4). Distortion Energy Theory refers to a method of equating stress to the amount of energy stored in an elastically deformed material. This verification method is solely applied at the hydrostatic (shell) test pressure condition. The theory assumes that if the part doesn't yield at test pressure, then it will operate properly at the rated working pressure condition. The drawback

OTC 23943-MS

with this method is that it is intended for simple geometries. Structural discontinuities and stress concentrations are beyond the scope and certain designs may result in localized areas of yielding at structural discontinues. The designer then has to use sound engineering judgment to call: how much yielding is allowed? HPHT applications compound the problem because of material strength de-rating. The "test pressure" is performed at room temperature, where the material strength is greater, while at operating pressure, the material strength is progressively weaker as the operating temperature increases. So, for equipment which will be operated at HPHT temperatures, we do not have the same margin of safety from the factory hydrostatic proof test which is conducted at room temperature, unless the test pressures are raised in ratio to the loss of strength at operating temperature. This means as the working temperature increases the Design Margin provided by this method decreases. Another limitation to distortion energy theory is that it requires excessive (exponentially increasing) wall sections, yet constrained for high pressure applications by limiting the test pressure not exceeding the material yield strength divided by 3. Division 2 vs. Division 3 and Load Resistance Factor Design As previously mentioned, whenever design practice is envisioned in the oil field, one often assumes Linear Elastic Stress Analysis when they refer to ASME VIII Division 2. However, Division 2 cites many design and analysis methods, including elastic-plastic and elastic-pure plastic methods; not just linear elastic. ASME VIII Division 3 contains most of the same techniques but with more stringent requirements for design pressures generally above 10 000 psi. Division 3 cites lower Design Margins and Test Pressures, seen as beneficial to keeping pressure vessel wall thicknesses reasonable. However, these lower values come with tighter controls on material strength and with far more material ductility (higher toughness) than Division 2 requires (covered by its more conservative Design Margins). So the argument made earlier that either Division 2 or Division 3 design methods may be used is a valid one. However, the caveat is "what are the actual values of these Design Margins?" Those specifically listed in the codes are related to pressure vessels alone and not so much combined load with other equipment. One of the more accurate analysis methods within Division 2 and 3 is Load Resistance Factor Design (LRFD). This technique uses finite element analysis (FEA) along with prescribed Load Factors to verify that a pressure vessel design has an adequate margin against collapse or rupture. ASME has prescribed tables of Load Factors for use with LRFD, one set for Division 2 and another for Division 3; established for common pressure vessel Load Combinations. The Load Combinations focus on pressure containment. But, can it be applied to pressure vessels with highly stressed load bearing parts, such as Tubing Hanger and Casing Hanger load shoulders? Many oilfield applications pressure containment are contingent on a load bearing interface that are integral with varying degrees of pressure end load. Therefore, the oil industry needs to develop its own set of Load Combinations and Load Factors, for HPHT and combined load bearing/pressure containment situations.

Existing Failure Assessment Diagram API 6A and 17D, which are most applicable to wellhead equipment, do not require the designer to address fatigue, only consider it. However, when designing HPHT equipment to either ASME VIII Division 2 or 3, cyclic loading and fastfracture failure need to be tested analytically. If an ASME pressure vessel can be shown to fail in a safe (leak-before-burst) mode, then S-N Fatigue analysis is acceptable to see if cyclic loads will adversely affect design. And if the load's stresses are kept below a certain limit, infinite cycle life can be inferred (and neglected as assumed in API 6A design calculations). If not, fracture mechanics is required (which scares nearly everyone in the oil industry). Division 2 realizes there is a point where stresses elevate to a level where leak-before-burst gives way to fast-fracture burst. But this threshold is very dependent on geometry, wall thickness, material toughness, material yield strength, along with flaw shape, size and where the high stress is located within the geometry. So, to test for this threshold, a fitness-for-service procedure typically used for evaluating existing pressure vessel hardware is cited as a means to evaluate a hypothetical design situation. This test, commonly referred to as the Failure Assessment Diagram (FAD) is found in several fitness-for-service codes: DnV RP C203, BS 7910 and API 579-1/ASME FFS-1. The FAD curve and process are diagrammed in Figure 1.

OTC 23943-MS

Figure 1 Definition of FAD Curve from API 579, figures 9.20, D.1, and equation 9.30

Figure 2 Overview of FAD Analysis Process from API 579, figure 2.2

FAD was developed to explain the brittle or sudden failure of materials. The intent of the test is to determine the Assessment Point given the geometry of flaw and where it's located, the type and thickness of the pressure vessel, and the vessel's material properties. The x-axis is defined as the Load Ratio (Lr), defined by a Reference Stress to the material Yield Stress: Lr = ref/ys The Reference Stress is a function of the maximum to minimum values of the cyclic plus mean stresses acting on the pressure vessel. The Reference Stress is a way of accounting for plastic behavior at the crack tip. It is based on the part geometry and crack dimensions and can be determined from formulas and tables of empirical data in API 579 for simple geometry, and by running 3D FEA for more complex geometry. The y-axis is defined as the Stress Intensity Ratio (Kr), defined by the StressIntensity Factor (K1), divided by the material's Fracture Toughness value (K1c): Kr = KI/KIC The Stress-Intensity Factor (K1) is a function of the flaw geometry, its location within the pressure vessel and is proportional to the stress at the crack's tip. K1c is a function of the material's ductility or (fracture toughness). If the assessment point ends up in the Acceptable Region, then there isn't enough energy from the surrounding stress field to

OTC 23943-MS

grow the crack and things remain stable. If it ends up in the Unacceptable Region, the crack is considered unstable and large enough that the crack will continue to grow leading to a brittle structural failure.

Figure 3 FAD Analysis Process from API 579 Stable crack vs. crack growth

Proposed Failure Assessment Diagram As cited, the FAD is an assessment tool for looking at an "existing part" and determining its remaining useful life. It looks at one assessment point at a time. If left here, the FAD provides little intuition as to predicting cyclic (dynamic) vs. a static or quasi-static state for a particular design and selected material without a litany of experimental cyclic loading tests to create a series of cracks to be plotted. To bring in cycle count into the FAD process, the use of another fracture mechanics principle is needed, the Paris Law for crack growth:

Figure 4 Paris Law for Estimating Crack Growth

Where a is crack depth, N the number of load cycles applied, K represents the fluctuation in Stress-Intensity Factor due to the cyclic load, C and m are crack growth parameters of the material (the linear Paris line in Region 2), and Kth represents the initial or Threshold Stress-Intensity Factor Range. Now using the same Stress-Intensity values for the material used in the FAD analysis, one can iterate through increasing cycles to generate a locus of Assessment Points starting with an assumed starting crack size and growing the crack according to the Paris Law equation it until the crack becomes large enough that the last assessment point intersects with the FAD's critical flaw curve (shown in Figure 3). The total number of cycles is then tabulated for the particular cyclic load criteria. The tabulated results can be graphed to create a representative S-N curve (Figure 5).

= ()

OTC 23943-MS

Figure 5 Representative S-N Curves for an Open-Ended Cylinder with Various Flaw Orientations

The representative curves represent the cycle count where the crack grows to critical flaw size based on the cylinder wall thickness, the material toughness, and the ratio of maximum to minimum stress applied to the pressure vessel. To the left of the curve infers the FAD's Acceptable Region, and to the right the Unacceptable Region. Now the graph becomes more intuitive to the designer, providing input on what kind of material offers the best performance for the size of the pressure vessel and flaw detection size detectable at the factory. The maximum to minimum stress ratio also provides a means to test various cyclic stresses or stress spectrums (through superposition) to see if any one set of cyclic loads creates a problem. In most instances the Acceptable Region remains fairly narrow, between 1.0 and 1.3 Stress Ratio. As cyclic stress becomes more prominent, the curve pinches off quickly, predicting cyclic unacceptability below 500 000 cycles. Sensitivity studies were performed to see how the representative curves move with respect to various parameters, illustrated in Figure 6.

OTC 23943-MS

Figure 6 Sensitivity Study of Pressure Vessel Cylinder Size, Initial Flaw Size, and Temperature Effects on Representative S-N Curves

The trends indicate that pressure vessel size, especially wall thickness will have the largest effect on cycle count. This can be inferred from the fact that the FAD predicts a much larger crack size has to break through a thicker walled object than a thinner one. Initial flaw size variation has a relatively minor effect on cyclic performance, bunching the curves close together. Temperature also seems to have a lesser impact because all of the stress values used to define Reference Stress and StressIntensity Factors are based on the same material yield strength, all going up and down proportionally with the temperature. However, there is some movement due to changes in the Paris Law variables, C and m. An increase in these values with increased temperature indicates an increase of the crack growth rate, and thereby moving the curve slightly to the left as temperature increases. This proposed representative S-N curve also has its drawbacks. First, the generated curves are only as good as the representative Paris Law information (C, m, K and Kth) to predict the correlation between cycle count and crack growth. Thereto the closeness of the geometry between the representative open-ended cylinder is compared to the component's actual configuration may affect the predicted performance. Second, the ratio between the cylinder's wall thickness and the crack depth (a) is a critical part of the FAD's estimation of Stress-Intensity (K1). If a component's cross section has an unusually thin cross section in a particular location, the pressure vessel cylinder used to model it may be too thick to be representative of crack growth and failure at that location. Third, the FAD process from API 579-1 is a linear elastic fracture mechanics model; and the designer is working in the realm of elastic-plastic behavior. So from the start, the FAD curve may have some level of inaccuracy as to where the critical flaw size occurs. Therefore, as an assessment point edges closer to the demarcation line, one should be skeptical with the resulting cyclic load capacities or cycle count which can be considered "safe". Furthermore, assessment points that fall to the right of the demarcation curve should not necessarily infer that the design has failed or has a limited fatigue life. It simply means that the design now squarely falls into the realm of detailed fracture mechanics or experimental cycle fatigue testing to properly assess design life. But it may indicate early on that the material selected, the cylinder size of the pressure vessel, detectable flaw size or operating temperature need to The proposed curves are simply a way of determining a quasi-static application from a dynamic one; and make the FAD test more of a front-end engineering exercise, than a back end inspection assessment exercise. The proposed representative S-N curve also depends on a lengthy list of assumptions to cut down the number of variables and provide a level of uniformity, independent of the actual component's design. The intent is to allow the representative S-N curves to be nearly generic as commonly published S-N curves, to be useful as a general design tool regardless of design methodology or specific knowledge of fracture mechanics. The assumptions are: 1. An open ended cylinder is assumed as the pressure vessel shape. The geometric size of the cylinder represents the inner most pressure containing member of a connector or tree body (outside diameter (D), inside diameter (d), and wall thickness (t)). The diameters and cross sectional wall need to be representative of the component's geometry which passes global stress requirements as dictated by Division 2 or 3 requirements in terms of mean and peak stresses (defined below). Reference Stress (ref) and Stress-Intensity factors for the cylinder may be found in Annex D of API 579-1 or "fit-for-service" software. Cyclic stresses are inferred from internal pressure generating hoop membrane stress in the cylinder wall. The same principal can be extended to infer similar stresses from external tension or bending. The upper value loading is generated by the stress at rated working pressure (max) and the minimum stress (min = (max - cyclic)) equals the difference of the peak stress and cyclic stress. In general, the primary stress as the peak stress is the input used to

2.

OTC 23943-MS

3. 4. 5.

6.

compute the initial FAD point. The redistribution of stress due to stress concentration anamolies (such as thread forms, machining run-out reliefs, or localized cross section differences) can be accounted by linear or polynominal stress distribution input in a static FEA run when assessing the applicable stress ratio (max / min). The crack is assumed to be a surface crack with a semi-elliptical shape. This is most widely used geometry when theoretically determining the Stress-Intensity Factor (K1) in fracture mechanics analysis exercises. Crack depth is a, and crack width is 2c. The relationship between a and 2c is an ASME recommended: 2c = 3a. Crack orientation may be: internal axial or circumferential, internal circumferentical, external axial or external circumferential. Internal axial appears to be the most conservative generated curve and makes the most intuitive sense when using the assumption of "membrane stress inferred by internal pressure". Material properties include: Yield Strength (ys), Ultimate Tensile Strength (ts), C and m for the da/dN Paris Law curve, Critical StressIntensity Factor (K1c), and Young's Modulus (E). Additional data sets need to be generated for every representative working temperature. Material properties and pressure vessel geometries are for parent material applications s. Analysis at weld locations require a special set of equations which are assumed to be beyond the intent of the representative S-N curve. Reference Stress and Stress-Intensity calculations for weld geometries, stress relaxation, and materials are discussed in detail in Annex E of API 579-1.

Figure 7 Representation of Maximum/Minimum Stress and Surface Crack Orientation on the Pressure Vessel Cylinder for Representative S-N Curves

Going Forward Since the completion of PER15K, API technical subcommittees have formed task groups to assess the HPHT issues posed in PER15K, to determine how many of them apply to their specific hardware group, and to develop a recommendation for the suite of specifications within the subcommittee's purview. API Subcommittee 17 has formed the 17TR8 task group to establish recommendations as they pertain to subsea hardware. The task group has been focusing on design methodology, material performance properties, and validation testing requirements to round out what is needed to build subsea HPHT hardware. API Subcommittee 6 (surface wellheads and trees), API subcommittee 11/19 (downhole hardware), and API subcommittee 16 (BOPs and drill through equipment) are also participating in the task group as they see similar issues and design practices developing within 17TR8 which may meet much of their technical specification needs. To date, the task group has agreed on a design decision flow chart that is helping to frame technical requirements and practices (Figure 8) going forward from this point. At the top of the decision chart is decision diamond with a "Rated Working Pressure > 15ksi?" question, then a pathway of arrows to Division 2 and Division 3 task boxes and an optional cross path line. All of this is to represent the current recognition that above 15ksi rated equipment, it is recommended to go the Division 3 route of elastic-plastic FEA design methodology. However the "optional" line speaks to the designer's choice to go the Division 2 route if engineering judgment dictates that the geometry of the component and design pressure, external loads, etc., are straightforward enough to be adequately modeled by elastic or other recognized Division 2 methods. If one continues down the "Division 2" path, then much of design methodology, PSL designations, and traditional Design Margins may be used. However, there are two caveats to continuing on this path for HPHT designs. First there are two decision diamonds: one to determine the thickness to OD ratio (Lam theory) and one of ratcheting (progressive strain movement in the pressure vessel wall as pressure is increased to working pressure then dropped back to ambient and up again).

OTC 23943-MS

Figure 8 Design Decision Flowchart for API 17TR8

10

OTC 23943-MS

As pointed out by ASME, if either one of these two decisions may force the designer back down the Division 3 path. As pointed out in OTC paper # 23621, the thin walled pressure vessel equations eventually become inaccurate at highly elevated internal pressures, skewing design results. So these decision diamonds direct the designer accordingly in this transitional region of working stress between 15 000 and 25 000 psi working pressure. The second caveat is that of elevated temperature. Above 350F (181C), pronounced reduction in material Yield and Tensile Strength may occur depending on the operating temperature in the field. Yet at factory test conditions, the material acts as if it is much stronger. To counter this, a Yield Strength Ratio () is multiplied to the factory shell test pressure. It is envisioned that Yield Strength Ratio will properly proof test the component even at factory temperatures. It should also alert the designer that selecting certain materials less conducive to HPHT environs will have too high a Yield Strength Ratio to for the application and subsequently steer the designer toward more favorable materials. The Division 3 route is no less arduous because of the more rigorous design methodology used. And it too comes with its own set of caveats. As previously mentioned, the more rigorous method comes with somewhat relaxed Design Margins. But this is because more stringent control of material properties, chemistry, melt practices, and ductility (toughness). To this the decision flowchart adds PSL 5 and 5G. Although specifics are still being worked, the designer requires more mechanical property data to make it common practice. True stress-true strain curves, Paris Law values, fracture toughness (K1c), Poisson Ratio, etc. will need to be performed and tabulated for most oil industry materials (many not currently listed in the ASME code) various levels of operating temperature 350(181), 400(205), 450(232) and so on. In addition, multiple batches of tests will need to be performed to gain stochastic confidence in the performance data, along with more conventional and less costly Quality Control values for each production run (surface hardness, Charpy impact values, etc.) performed on each production melt of the material to ensure efficacy. Also mentioned earlier, new Design Margins for pressure containing/load bearing etc. conditions need to be worked out with Industry agreement that are not addressed in current Division 3 Combination Load Design Margin tables. The Division 3 path also adds to the test pressure multiplier for temperature difference. At the bottom of each path is a final decision diamond; that of determining if cyclic loads are present and performing a FAD test. This is where the proposed Representative S-N curve comes into play. The proposed FAD process looks at the cyclic loading case to see if the component is "quasi-static" or "requires fracture mechanics" analysis, regardless of the path taken. The paths taken afterwards are dependent on the materials selected, and the loading conditions, largely irrespective of the component's design. Should fracture mechanics not be required, then design calculation process nears completion with the design life based on calendar life and corrosion allowances. Should the operating temperature fall below 350F (181C), then validation (qualification/performance tests on prototype equipment) may use conventional API protocols as defined by either PR1 or PR2 requirements. Should the operating temperature go above, then a new qualification threshold Performance Requirement 3 (PR3) is invoked (with higher performance standards to be worked out). On the other hand (and regardless of which design path taken), if the proposed FAD test points toward fracture mechanics, the design life is now dictated by cycle life and fit-for-purpose monitoring of cracks, etc. as defined in Division 3 and API 579-1 (or BS 7910 or DnV RP C203). The task group is also working on the concept of a Performance Requirement 4 (PR4) validation protocol for prototype cycle testing under load or analysis based on the "criticality level" the component is intended to be used and for how long it is planned to be in service.

Conclusions HPHT design methodology is starting to take shape and gain Industry acceptance. The choice between Division 2 and Division 3 design paths are better understood but still require clarity in establishing the necessary design margins set to properly assess combined loading situations. Similar clarity is being achieved on the materials front; knowing what mechanical properties are needed and at what level of stochastic accuracy. This is also reflected in adopting the new concept of PSL 5 and 5G (currently under discussion within the API HPHT work groups) to acknowledge these more stringent mechanical property requirements. However the arduous task to obtain and qualify a full suite of HPHT materials with their properties at five or more operating temperatures remains. Factory (shell proof) pressure test's test margins are also being adapted to take into account for material temperature de-rating in the field at operating conditions, which may not be possible to test for at the factory. A Yield Strength ratio is added as a multiplier to the test margin so that components may be "put under stress" more closely to what the component will be subjected to in the field. And designers may be less inclined to select materials exhibiting a higher ratio value due to higher temperature conditions. The paper also proposes a FAD test which offers a "representative S-N curve"; based on fracture mechanics, but presented in a more intuitive format. Its generic cylinder-surface flaw configuration in combination with widely accepted FAD and Paris Law fracture mechanics assessment are combined to create a generic representation that decouples the test from the perception that a is required, regarless of whether its warrented or not. The representative curves also lend themselves to providing a demarcation line between equipment working in a "quasi-static" environment from those working in a cyclic

OTC 23943-MS

11

loaded dynamic environment. However, these curves are a representation, and the designer is encouraged to carefully review and understand the steps and assumptions in generating the curves for a particular material and cylinder size. The demarcation between standard API design, HPHT design, and fracture mechanic analyzed regions lends itself to two new Performance Requirement criteria, PR3 and PR4, being discussed within the API HPHT work groups. PR3, cites more involved performance testing, and PR4, deals with cyclic loading until a damage limit or structural failure are demonstrated to validate the design. As a final note, it should be pointed that this design decision flowchart covers pressure containing, pressure controlling and load bearing equipment. However, nothing is mentioned regarding seals and fasteners, which are just as important to the HPHT design effort as the pressure vessel itself. The 17TR8 task group is only now forming a study group to address the particular design and performance needs of seals and fasteners and the materials to manufacture them. Seals and fasteners, by their nature, are highly loaded components at both rest and operating conditions (either to affect a sealing point or structurally hold pressure containing or pressure controlling components together. As such they will require all of the scrutiny mentioned in this paper, but need to address stress relaxation and high temperature ageing (for elastomers and thermoplastics) as well. Hopefully subsequent papers and reports will address them.

Acknowledgements The authors would also like to thank the management and technical staff of FMC Technologies for permission to publish this paper and provide a constructive critique of its contents. In addition, a continued thanks to John Bednar, Paul Bunch, Terry Cook, Dan Peters, Earl Shanks, Bob Sokoll, Bob Sims, John Vicic, and Ken Young for their moral support and technical expertise along the way as the Industry migrates along the HPHT trail from 6HP to PER15K to 17TR8.

Nomenclature a API ASME BOP BPVC BS C c K crack (flaw) depth American Petroleum Institute American Society of Mechanical Engineers Blow Out Preventer Boiler and Pressure Vessel Code British Standard Fracture Mechanics material constant half of crack (flaw) width (2c = crack width) Change in Stress-Intensity Factor during Operational Cycle Threshold Stress-Intensity Factor Range Kth D Pressure Vessel Outside Diameter (2Ro) d Pressure Vessel Inside Diameter (2Ri) da Change in crack size a dN Change in number of operational cycles N DnV Det Norske Veritas E Young's Modulus of a material EPFM Elastic-Plastic Fracture Mechanics FAD Failure Assessment Diagram FEA Finite Element Analysis HPHT High Pressure High Temperature ISO International Standards Organization Stress-Intensity Factor KI Fracture Toughness for a material KIC Stress Intensity Factor Ratio - KI/KIC Kr ksi 1000 Pounds per Square Inch (6.89 MPa) Load Ratio - ref / ys Lr LRFD Load Resistance Factor Design Crack Threshold (Initial) Stress-Intensity Kth Fracture Mechanics material constant Non-Destructive Examination Oil Country Tubular Goods PER15K Protocol for Equipment Rated Greater than 15 000 psi PR API (6A-17D) Performance Requirement PSL API (6A-17D) Product Specification Level Inner Radius of the Pressure Vessel Ri Outer Radius of the Pressure Vessel Ro RP Recommended Practice RWP Rated Working Pressure S-N Allowable Stress Amplitude 'S' - Number of Design Cycles 'N' t Wall Thickness of the Pressure Vessel TR Technical Report cyclic Cyclic Membrane Stress of the Pressure Vessel Membrane Stress of the Pressure Vessel m Yield Strength of a material ys Reference Stress ref Membrane Stress at Rated Working RWP Pressure Maximum Membrane Stress of the max Pressure Vessel Minimum Membrane Stress (max - cyclic) max Tensile Strength of a material ts ys at Factory Temperature / ys at Operating Temperature Ratio m NDE OCTG

12

OTC 23943

References
American Petroleum Institute, "Specification for Wellhead and Christmas Tree Equipment", Specification 6A, 19th Ed., 2005 2. American Petroleum Institute, "Design and Operation of Subsea Production Systems - Subsea Wellhead and Tree Equipment", Specification 17D, 2nd Ed., 2011 3. American Society of Mechanical Engineers, "Boiler & Pressure Vessel Code, VIII Division 2", 2011 4. American Society of Mechanical Engineers, "Boiler & Pressure Vessel Code, VIII Division 3", 2011 5. American Petroleum Institute, "Fitness-For-Service", Standard 579-1 / ASME FFS-1, 2nd Ed., 2007 6. Bascom, J.M., Rolfe, S.T., Fracture and Fatigue Control in Structures, 3rd Ed., ASTM, 1999 7. Cordes, R.D., San Pedro, R.I., "Fracture Mechanics-Based Fatigue Predictions for HPHT Equipment", Offshore Technology Conference, Paper # 19502, May 2008 8. Kiciak, A., Glinka, G., Burns, D.J., "Calculation of Stress Intensity Factors and Crack Opening Displacements for Cracks Subjected to Complex Stress Fields", Transactions of the ASME, Vol. 125, pp. 260-266, August 2003 9. Koucurek, C. Pathak, P., Melancon, C., Sohn, S., " Merging ASME and API Design Methods for Subsea Equipment up to 25000 PSI Working Pressure", Offshore Technology Conference, Paper # 23063, May 2012 10. McKie, N., Skeels, H.B., Williams, M.R., Marotta, E., Peters, D., "15K Plus The Limits of Establishing Pressure Vessel Design Limits", Offshore Technology Conference, Paper #23621, May 2012 11. Quest Reliability, "Signal Fit-For-Service Users Manual", Ver. 3.0 12. Skeels, H.B., The Impact of HPHT on the Oil Industry, Deep Offshore Technology Conference, Perth Australia, December 2008 1.

Vous aimerez peut-être aussi