Proceedings of the ASME 2014 33nd International Conference on Ocean, Offshore and Arctic Engineering OMAE 2014 June 8-13, 2014, San Francisco, USA

Paper OMAE-2014-23234

FATIGUE OF RISERS: CALIBRATING RELIABILITY ESTIMATES FROM FULL-SCALE DATA

Michael A. Tognarelli BP America Production Co. Houston, TX, USA [email protected]

Steven R. Winterstein Probability-Based Engineering Menlo Park, CA, USA [email protected]

ABSTRACT A number of measurement campaigns have sought to quantify fatigue damage on drilling and production risers in the field. The data from these campaigns have recently been used to test the accuracy of software predictions of riser motions and hence fatigue. It is often found that state-of-the-art software shows a conservative bias in predicting fatigue damage—conservative factors of 10 or more. At the same time, short-term damage estimates (e.g., over 10-minute intervals) show significant scatter: COV (coefficient of variation) values of 1.0–2.0 are typical. We suggest methods here to better incorporate the information from these measurements into fatigue reliability predictions. We note that (1) software error statistics show markedly different behavior in different regimes; e.g., for different levels of predicted damage; and (2) our fatigue reliability concern lies with the variability not in individual 10-minute damage rates, but rather in the long-run damage rate over the entire structural life. These features suggest a non-parametric approach, in which software error statistics are binned and separately analyzed for different predicted damage levels. The sample means in each bin, together with their variabilities, then form the basis for our reliability assessment. Because our interest lies with the average behavior in each bin, variability reduces notably: COVs of 1.0–2.0 (on individual 10-minute rates) are reduced by an order of magnitude or more. This variability is commonly dominated by the uncertainty in fatigue capacity (S–N curve, Miner’s rule, etc.) Thus, our measurement campaigns are of most interest in predicting only the mean bias in damage predictions; conventional methods are

then available to propagate these into a standard fatigue reliability analysis. Results are shown for a number of applications, and implications considered for design (i.e., load factors).

INTRODUCTION Risers on offshore structures are commonly subject to fatigue. Their typical design basis uses analytical models of riser motions/stresses, together with Miner’s rule, to predict fatigue life. We then require that this predicted life exceed the actual service life by a prescribed safety factor—commonly a factor of 10. Note that this factor of 10 does not assume that the software producing the prediction is conservative (in damage terms) by a factor of 10. It typically assumes the software is unbiased, and the factor of 10 is intended to cover the rather large uncertainty in fatigue capacity (number of cycles to fail under given loading). Recently, a number of measurement campaigns have been used to test the accuracy of software in predicting riser motions. This is of special interest in the technically challenging area of VIV (vortex-induced vibrations). In these cases, conventional software such as SHEAR7 [1] often yields strongly conservative estimates of fatigue damage. At the same time, the biases in short-term damage estimates (e.g., over 10-minute intervals) show significant scatter: COV (coefficient of variation) values of 1.0–2.0 are typical. The question, which has been pursued over a number of years in the technical literature [2–9], is how to incorporate the information of these measurements to better assess riser fatigue. This question, of course, is of interest for WIF (wave-induced fatigue) as well. 1

c 2014 by ASME Copyright

This paper seeks to improve the methodology to address both VIV and WIF problems. Notable features of these problems include

GOM Riser 10 1 Measured Damage Rate

1. Software error statistics show markedly different behavior in different regimes; e.g., for different levels of predicted damage. 2. Our fatigue reliability concern lies not with the variability in individual 10-minute damage rates, but rather in the longrun damage rate over the entire structrual life. Both of these features suggest a non-parametric approach, in which software error statistics are binned and separately analyzed for different predicted damage levels. The sample means in each bin, together with their variability, then form the basis for our fatigue reliability prediction. Because our interest lies with the average behavior (in each bin), variability reduces notably: COVs of 1.0–2.0 (on individual 10-minute rates) are reduced by an order of magnitude or more. This variability is commonly dominated by the uncertainty in fatigue capacity (S–N curve, Miner’s rule, etc.) Thus, our measurement campaigns are of most interest in predicting the mean bias in damage predictions; conventional methods are then available to propagate these into a standard fatigue reliability analysis. Results are shown for various applications, and implications considered for design (i.e., load factors).

0.1 0.01 0.001 0.0001 1e-05 1e-06 1e-07

10-Minute Data Perfect Prediction 1e-08 1e-08 1e-07 1e-06 1e-05 0.0001 0.001 0.01 0.1 Predicted Damage Rate

1

10

FIGURE 1. Measured damage rate versus predicted damage rate [yr−1 ] from SHEAR7 (version 4.6) for a Gulf of Mexico deepwater drilling riser.

X = Measured Damage / Predicted Damage

GOM Riser

APPLICATIONS TO RISER VIV VIV of a Gulf of Mexico Drilling Riser We first consider a MODU drilling riser subjected to highly sheared currents, typical in the Gulf of Mexico. We use an extensive dataset, comprised of n=363 10-minute VIV episodes. During each episode, measured riser motions are used to infer the ˙ based on (1) a standing wave response fatigue damage rate, D, model and (2) Miner’s rule. For each episode, a corresponding predicted damage rate, D˙ pred , is found from the software code SHEAR7. The maxima along the length, for each case, are then compared. This dataset has been studied extensively; for example, Refs [3, 9] have compared measurements with version 4.4 of SHEAR7. Here we instead use version 4.6 of this software, with the developer-recommended parameters for bare risers. Figure 1 compares the measured and predicted damage rates. It is clear that the predictions span an extremely wide range— more than 7 orders of magnitude, from D˙ pred =.98·10−7 to 2.17 [per year]. It is also clear that the predictions often fail to closely follow the measurements. In particular, large predicted values tend to be conservative, while low predictions are often nonconservative. Indeed, the data show two distinct “clouds,” of non-conservative and conservative behavior respectively, separated at around the D˙ pred =10−4 level. (There are considerably more data in the rightmost, conservative cloud.)

1000

10-min data Bin average Bin std dev Perfect Prediction

100 10 1 0.1 0.01 0.001 0.0001 1e-05

1e-06 1e-08 1e-07 1e-06 1e-05 0.0001 0.001 0.01 Predicted Damage Rate

0.1

1

10

FIGURE 2. Modelling error versus predicted damage rate [yr−1 ] from SHEAR7 (version 4.6) for a Gulf of Mexico deepwater drilling riser.

These observations motivate our statistical study of the modelling error

Xi =

D˙ i ˙ D pred,i

(1)

in which D˙ i is the damage rate inferred from the measurements, and D˙ pred,i is the damage rate predicted by the software. In particular, we separate the data into different bins of D˙ pred,i , and 2

c 2014 by ASME Copyright

estimate the mean and standard deviation of Xi in each bin: mXi = E[Xi |D˙ pred,i ] q σXi = Var[Xi |D˙ pred,i ]

Here ni and n j are the number of data in bins i and j, σXi and σX j are the conditional standard deviations (Eqn. 3), and ρi j is the correlation coefficient between Xi and X j .

(2)

Modelling Issue 2: In previous work (e.g., [3, 9]), the variance of X was taken as the variability shown in the raw data; i.e., √ the 10-minute damage rate values. The ni n j term in Eqn. 6 reflects our interest here in the variability of the long-run damage rate, not of individual 10-minute damage rates.

(3)

Figure 2 shows the result1 . Note that conservative predictions are here associated with X < 1. The figure thus shows a clear trend toward increasing conservatism (decreasing mean X) with increasing D˙ pred . The standard deviation σX tracks rather closely with its mean, suggesting a coefficient of variation, σX /mX , on the order of 1.0. This is consistent with previous findings. (Note that one bin contains only a single 10-minute data point, hence a standard deviation cannot be estimated.)

In the simplest case, the failure time T f is taken as 1/D˙ LT , the time when Miner’s damage D(T f )=1. In probabilistic analysis, it is common to generalize this by predicting failure when D(T f ) exceeds a (random) threshold ∆. It follows that Tf =

Modelling Issue 1: In previous work (e.g., [3, 9]), the modelling errors Xi were effectively taken as independent of D˙ pred,i . This ignores the strong conservative trend shown in Fig. 2 and similar datasets.

E[T f ] ≈

i

(4)

E[∆] = e2σln N = 102σlog N ; V∆ = 0.7

(8)

(9)

(5) This result for E[∆] is straightforward: the design S–N curve is shifted conservatively by 2 standard deviations (2σlog N ) on log scale. (For the Class E S–N curve used here, σlog N =.2509 so that E[∆]=3.18 [10].) The assumption that V∆ =0.7 is more tenuous. This variability must include not only the scatter in the S–N curve, but all other uncertainties (e.g., Miner’s rule that generalizes from constant to variable amplitude loads). The value V∆ =0.7 is roughly consistent with values suggested in the literature (e.g., [11, 12]), and hence is adopted here. For highly uncertain fatigue loading processes, this value may need to be increased2 .

i

Here D˙ pred,i is the damage rate predicted in current conditions i, and pi is the long-run fraction of time this current condition acts. The result depends only on their product, di = pi D˙ pred,i , which can be viewed as a “damage density.” The sum ∑i di yields the long-run damage estimated by the software. The mean correction factors mXi (Eqn. 2) are introduced to account for software biases. Having binned data based on D˙ pred,i , mXi values are estimated by X i , the sample mean observed in that bin. These are the mean values shown in Fig. 2. The statistical uncertainty in our damage rate estimate (Eqn. 5) can then be expressed by the variance σXi σX j Var[D˙ LT ] = ∑ ∑ di d j ρi j √ ni n j i j

q E[∆] ; VT f ≈ V∆2 +VD2˙ LT D˙ LT

In the second of these results, V refers to the coefficient of variation (ratio of standard deviation to mean) of the subscripted variable. (This result becomes exact if each V is replaced by the standard deviation of the logarithm of the variable.) p To use these results, D˙ LT and VD˙ LT = Var[D˙ LT ]/D˙ LT are evaluated with Eqns. 5–6. Corresponding statistics of ∆ are assumed here to be

This long-run damage rate, D˙ LT , can in turn be decomposed into a weighted mixture of damage rates that occur in different current conditions: D˙ LT = ∑ pi D˙ pred,i mXi = ∑ di mXi ; di = pi D˙ pred,i

(7)

Corresponding moments of T f can be estimated as

Long-Term Reliability Analysis We assume that D(T ), the value of Miner’s damage after a sufficiently large number of cycles, is well-approximated by its long-run average: D(T ) = D˙ LT T

∆ D˙ LT

Numerical Results We now apply these results, using long-term probabilities pi of different current profiles from a deepwater Gulf of Mexico site. The long-term damage rates, with and without software

(6)

1 The

mean and standard deviation values are only plotted at the (logarithmic) midpoint of each bin—and a straight line connected for visual purposes between these points. The bin widths are constant on log scale.

2 The

converse benefit, if V∆ is increased beyond 0.7, is that the effect of software uncertainty (reflected through VD˙ LT ) becomes relatively less important.

3

c 2014 by ASME Copyright

Case:

E[∆]

D˙ LT (yrs−1 )

E[T f ] (yrs)

(Eqn. 5)

(Eqns. 8–9)

VD˙ LT

VT f

(Eqns. 5–6)

(Eqns. 8–9)

0.0

0.051

0.702

SHEAR7: Ver 4.6, default parameters

0.5

0.103

0.708

S–N: DNV Curve E

1.0

0.137

0.713

0.0

0.108

0.708

SHEAR7: Ver 4.6, default parameters

0.5

0.165

0.719

S–N: DNV Curve E

1.0

0.206

0.730

0.0

0.106

0.708

Flexcom plus SHEAR7(v4.6 default)

0.5

0.150

0.716

S–N: DNV High-Strength with C.P.

1.0

0.184

0.724

VIV, Gulf of Mexico MODU drilling riser

VIV, North Sea MODU drilling riser

WIF plus VIV, Jackup production riser

3.18

3.18

2.11

TABLE 1.

V∆

0.7

0.7

0.7

1.02

×10−5

ρ

310000

3.74 ×10−2

85

1.94 ×10−3

1100

Results across various cases.

correction, are found to be

GOM Drilling Riser, E[Tf] = 310,000 yrs 1.8e-07

i

D˙ LT = ∑ di mXi = 1.0 × 10−5 i

1 26 yrs

1 = 1.0 × 105 yrs

1.6e-07

(10)

1.4e-07 1.2e-07

(11)

P [ Tf < t ]

D˙ LT,pred = ∑ di = 3.9 × 10−2 =

There is thus a significant effect of software bias—the corrected damage rate is reduced by more than 3 orders of magnitude:

1e-07 8e-08 6e-08 4e-08 rho = 0.0 rho = 0.5 rho = 1.0

2e-08

1.0 × 10−5 Mean software bias = B = = = 2.5 × 10−4 3.9 × 10−2 D˙ LT,pred (12) To understand this, note that most of the significantly damaging currents have D˙ pred values in the 0.1–3.0 [yr−1 ] range. In this range, Fig. 2 indeed shows mean biases on the order of 10−3 – 10−4 . Table 1 shows corresponding estimates of VD˙ LT , the coefficient of variation of D˙ LT due to software errors. Results are shown for 3 different (hypothetical) values of ρ, the correlation coefficient between errors in different D pred bins. Notably, these values (between .05–.14) are considerably smaller than the “background” variability of V∆ =0.7. In view of the SRSS operation in Eqn. 8, the effect of VD˙ LT will therefore be rather negligible. This is shown in Table 1: VT f varies only from .702–.713. D˙ LT

0 0

1

2 3 Service life, t [yrs]

4

5

FIGURE 3. Failure probability versus service life t for GOM drilling riser. Weibull model of T f assumed, with various error correlation values of ρ.

the various assumed values of the correlation coefficient ρ. The Weibull model is often used to model failure times of deteriorating systems, as its hazard function increases monotonically (as a power-law function) for VT f < 1. It also arises as a minimum strength distribution, hence is relatively heavy in its lower (left) tail. As this is the tail of interest (low failure times due to rare weak members), this would seem a fairly reasonable choice. Figure 4 compares Weibull, Gamma, and lognormal models fit to the same moments. (The correlation ρ=1 case is assumed.) As may be expected, the Gamma model predicts somewhat lower

The Choice of Probabilistic Model Figure 3 fits a Weibull model of T f , based on its moments in Table 1. As shown in the table, the COV values are based on 4

c 2014 by ASME Copyright

GOM Riser, E[Tf] = 310,000 yrs

North Sea Riser 100

0.0001

10 Measured Damage Rate

0.01

P [ Tf < t ]

1e-06 1e-08 1e-10 1e-12

1e-16 1

10

100 Service life, t

1000

Weibull PDF Gamma PDF Lognormal PDF

0.7 PDF

0.6 0.5 0.4 0.3 0.2 0.1 0 0.001

0.01 0.1 Normalized Failure Time, Tf / E[Tf]

0.001

0.01 0.1 1 Predicted Damage Rate

10

100

North Sea Riser X = Measured Damage / Predicted Damage

0.8

10-Minute Data Perfect Prediction

FIGURE 6. Measured damage rate versus predicted rate [yr−1 ] from SHEAR7 (version 4.6) for a North Sea riser.

PDF of Tf = Failure Time; COV = 0.7 Assumed 1

0.01

0.0001 0.0001

FIGURE 4. Failure probability versus service life t. Various probabilistic models of T f assumed, with error correlation ρ=1 assumed.

0.9

0.1

0.001

Weibull Gamma Lognormal

1e-14

1

1

10

10-min data Bin average Bin std dev Perfect Prediction

1

0.1

0.01

0.001 0.0001

0.001

0.01 0.1 1 Predicted Damage Rate

10

100

FIGURE 7. Modelling error versus predicted damage rate [yr−1 ] from SHEAR7 (version 4.6) for a North Sea riser.

FIGURE 5. Probability density function (PDF) fTf (t) versus normalized time t/E[T f ]. Various probabilistic models of T f assumed, all with VTf =0.7.

(right) tail, and hence is quite narrow in its lower tail. This point is fairly clear in the PDF comparison in Fig. 5. In view of the popularity of the lognormal model in fatigue applications, this point should perhaps be emphasized.

failure probabilities than the Weibull (with its heavy left tail). More dramatic is the lognormal result: it predicts a failure probability of 10−16 only when service lives approach 1000 years—at which time the Weibull and Gamma models yield failure probabilities 13 and 12 orders of magnitude larger, respectively. This suggests that it may be imprudent to use a lognormal model in these applications, where capacity uncertainty dominates and the lower tail (weakest strength, smallest failure time) is of critical interest. The lognormal, due to its exponential transformation, assigns relatively large probabilities in its upper

VIV of a North Sea Drilling Riser We consider now a riser subjected to VIV in a North Sea environment. Data have been measured for a drilling riser, without strakes or any other VIV suppression device. Measured damage rates are inferred from measured riser accelerations. Predicted damage rates are found from version 4.6 of SHEAR7, again us5

c 2014 by ASME Copyright

ing default parameters and the DNV Class E S–N curve. Figure 6 compares measured and predicted damage rates (per year) for 234 10-minute data segments. Results are rather better behaved than in Fig. 1: the data show stronger correlation, and no distinct subpopulations. Nonetheless, there is considerable scatter, and predictions show a clear conservative bias. Figure 7 shows the result of binning data based on the predicted damage level D pred . Results are qualitatively similar to those in Fig. 2: the bin average and standard deviation are numerically similar, and both fall off (showing increasing conservatism in SHEAR7) as D pred grows. We again apply the reliability analysis steps in Eqns. 5–9, now with long-run pi appropriate for a set of North Sea current profiles. The long-term damage rates, again with and without software correction, are D˙ LT,pred = ∑ di = 4.9 × 10−1 =

1 2.0 yrs

(13)

D˙ LT = ∑ di mXi = 3.7 × 10−2 =

1 27 yrs

(14)

i

i

North Sea Drilling Riser, E[Tf] = 85 yrs 0.02 0.018 0.016

P [ Tf < t ]

0.014

0.008

0.004 rho = 0.0 rho = 0.5 rho = 1.0

0.002 0 0

1

2 3 Service life, t [yrs]

4

5

FIGURE 8. Failure probability versus service life t for North Sea drilling riser. Weibull model of T f assumed, with various error correlation values of ρ.

2. “Measured” fatigue damage is inferred from accelerometer measurements on several risers, based on a transfer function approach developed in BPs Upstream Engineering Centre (UEC). 3. The location of interest is a particular fatigue detail on a component. 4. In both the predicted and measured cases, damage is conservatively assumed to accumulate at a single location on the component circumference. 5. Fatigue damages are calculated using the DNV HS (highstrength) S-N curve with Cathodic Protection. From Section 2.4.10 of [13], the design S–N curve is log N=17.446 − 4.70 log S, while the mean S–N curve is log N=17.770 − 4.70 log S. It follows that E[∆]=10(17.770−17.446) or 2.11. 6. Predictions have been made based on a fixed upper boundary condition, which corresponds better with measured data when compared to a pinned condition.

3.7 × 10−2 = 0.076 4.9 × 10−1 D˙ LT,pred (15) Note that there are markedly different current environments in the two cases. The Gulf of Mexico example is in very deep water marked by steeply sheared currents, while the North Sea case is in shallower water with slab-like currents. These may therefore impose different predictive challenges to a code like SHEAR7, and hence lead to different biases. Finally, Fig. 8 shows the resulting variation of failure probability with service life in this North Sea case. D˙ LT

0.01

0.006

Note that the software conservatism here is only an order of magnitude, compared with 3 or more orders of magnitude in the Gulf of Mexico case:

Mean software bias = B =

0.012

=

Numerical Results Figure 9 compares predicted and measured fatigue damage (per 30 minutes), and Fig. 10 shows the resulting modelling error X. Trends are similar to other cases: increasing conservatism (lower X values) as the predicted damage D pred increases. As may be expected, the conservatism (in this wave-induced fatigue case) is not as large as in the VIV Gulf of Mexico case. The long-term damage rate, D˙ LT , is again estimated from Eqn. 5. The only conceptual difference is that for wave-induced fatigue, the pi denote probabilities that various (HS , TP )i pairs occur (rather than current profiles, as in the VIV case). For our long-term predictions, we consider wave-induced fatigue only.

APPLICATION TO WAVE-INDUCED FATIGUE Finally, we consider the fatigue of production risers on a jackup, at a water depth of roughly 100m. The motions of a number of risers have been measured. These have been used to compare fatigue damage predictions with the damage inferred from the measurements across different conditions. Some features of these data: 1. Predicted fatigue damage is the sum of wave fatigue damage calculated in Flexcom and VIV fatigue damage calculated in SHEAR7v4.6. The wave-induced damage typically dominates in this case. 6

c 2014 by ASME Copyright

Jackup Riser, Wave-Induced Fatigue and VIV

Jackup Production Riser 0.01 pf (T) = Prob [ Failure in time T]

Measured Damage Rate

1e-05

1e-06

1e-07

1e-08

1e-09 Data Perfect Prediction

1e-10 1e-10

1e-09

1e-08 1e-07 Predicted Damage Rate

1e-06

0.001

0.0001

1e-05

Weibull, pf(T) Weibull, pf(T)-pF(T-1) Gamma, pf(T) Gamma, pf(T)-pf(T-1)

1e-06 1e-05

5

FIGURE 9. Measured damage rate versus predicted fatigue damage rate (per 30-minute interval) for a jackup riser. Predictions include both wave-induced fatigue and VIV.

10

15 Time T (yrs)

20

25

30

FIGURE 11. Failure probability versus service life T for the jackup riser. Weibull and Gamma models of T f assumed, with ρ=1.

ably less than in the VIV cases: X = Measured Damage / Predicted Damage

Jackup Riser, Wave-Induced Fatigue and VIV 100

data Bin average Bin std dev Perfect Prediction

10

Mean software bias = B =

D˙ LT D˙ LT,pred

=

1.1 = 0.61 1.8

(18)

Finally, Fig. 11 shows failure probability estimates for various values of service life T . Results are shown both for total failure probability over the life, and worst annual failure probability (in the last year of service).

1

0.1

0.01

0.001 1e-10

1e-09

1e-08 1e-07 Predicted Damage Rate

1e-06

ESTIMATION OF LOAD FACTORS The foregoing results are all problem-specific—the absolute numbers are tied directly to the situation at hand. In this final section, we seek to produce more generalized results. In particular, from a design perspective, we seek to give insight into plausible values of the safety factor γ=Tdes /Top , the ratio between the design life Tdes and the operating life Top of the fatigue-sensitive component. To anchor the ensuing discussion, γ=10 is a typical value used in this context. In probabilistic design, we will typically seek to impose a target value on the failure probability

1e-05

FIGURE 10. Modelling error versus predicted fatigue damage rate (per 30-minute interval) for a jackup riser. Predictions include both wave-induced fatigue and VIV.

The long-term damage rates, again with and without software correction, are D˙ LT,pred = ∑ di = 1.8 × 10−7 per 30-min = i

D˙ LT = ∑ di mXi = 1.1 × 10 i

−7

pF (Top ) = P[D˙ LT Top > ∆]

1 (16) 320 yrs

(19)

It is clear from the foregoing results that it is critical to estimate statistics of the software bias B:

1 per 30-min = (17) 510 yrs

B=

As noted above, the the software conservatism here is consider7

D˙ LT D˙ LT,pred

(20)

c 2014 by ASME Copyright

To incorporate Eqn. 20 into Eqn. 19, we note that since Tdes =1/D˙ LT,pred :

Target failure probability, pf , over operating life

Top B D˙ LT Top = BD˙ LT,pred Top = B = Tdes γ

Weibull Model, COV = 0.8

(21)

Substituting Eqn. 21 into Eqn. 19,

pF (Top ) = P[

B ∆ 1 > ∆] = P[ < ] γ B γ

(22)

The latter form of this result is most convenient for our purposes, as it collects all uncertain terms into a “capacity” variable ∆/B that should not exceed 1/γ. A normalized version of this variable is defined as

U=

∆/∆ B/B

1

Bias= 1.0 Bias= 0.1 Bias= 0.01 Bias= 0.001

0.1 0.01 0.001 0.0001 1e-05 1e-06 1e-07 1

10 Safety factor gamma

100

FIGURE 12. Safety factor γ versus failure probability p f over operating life. Weibull distribution assumed with V =0.8.

(23) CDF of Tf = Failure Time; Weibull Model Assumed 1

in which ∆ and B are the mean values of the respective variables. U will thus have (approximately) unit mean. Recasting Eqn. 22 in terms of U yields B ] γ∆

0.01 P [ Tf < Top ]

pF (Top ) = P[U <

0.1

(24)

0.001 0.0001 1e-05

The quantity B/γ∆ serves as an (inverse) reliability index— smaller values yield higher reliability. Thus, higher reliability is achieved by (1) reducing the (unconservative) software bias B, (2) increasing the (conservative) bias ∆ in the S–N curve, or (3) increasing the safety factor γ. Figure 12 shows typical results from Eqn. 24. A Weibull distribution has been assumed for U, with V =0.8 (V here is the same as VT f in the previous results.) ∆=3 has been assumed, an intermediate value between those for the F2 and E S–N curves. Bias values B=1.0, 0.1, 0.01, and 0.001 have been assumed. It is clear that the software bias B plays a critical role here. For our Gulf of Mexico case we find B values less than .001 (Eqn. 12). Using a Weibull model for T f in this case, even with no safety factor (γ=1) we have p f (Top ) on the order of 10−6 , well within the highest DNV safety class. (DNV permits p f =10−5 as the worst annual failure probability, which in turn permits a larger p f over the entire operating life.) In contrast, if the bias increases to B=0.1, one cannot achieve p f (Top )=10−5 even if γ=100. Another way to consolidate our foregoing results is to consider a plot of the Weibull CDF (cumulative distribution function) of T f versus the normalized variable Top /E[T f ] (analogous

COV = 0.7 COV = 0.8 VIV GOM VIV NS WIF

1e-06 1e-07 1e-05

0.0001

0.001 0.01 0.1 Normalized time, Top / E[Tf]

1

10

FIGURE 13. Cumulative distribution function of Weibull failure time T f versus normalized time Top /E[T f ]. Also shown are Top /E[T f ] levels for the three cases considered here.

to the PDF plot in Fig. 5). The explicit form of this CDF is shown in Appendix I (Eqn. 27). Figure 13 shows the resulting CDF for COV values of 0.7 and 0.8, which bracket the cases shown here. Note that this same normalized curve applies for all our cases, when entered at the appropriate value of Top /E[T f ]. The figure shows these values for our 3 cases, with Top =5 yrs for the drilling risers (VIV GOM and VIV NS) and 20 yrs for the production riser (WIF of jackup). E[T f ] values are found in Table 1. The VIV GOM case is shown to be strikingly more reliable than the other two cases, by about 4 orders of magnitude. Again, this is due to the extreme conservative bias in this case (Eqn. 12). 8

c 2014 by ASME Copyright

REFERENCES [1] Vandiver, J., Wajnikonis, C., Resvanis, T., and Marcollo, H., 2011. User guide for SHEAR7 Version 4.6. Technical report, MIT. [2] Stahl, B., 2000. Fatigue safety factors for deep water risers. Technical report for BP Deepwater Underwater Technology Group, Bernie Stahl Consulting, Houston, TX. [3] Tognarelli, M. A., Gabbai, R., and Campbell, M., 2009. “An approach to include observed VIV likelihood in drilling riser fatigue analyses”. In Proceedings, 28th International Conference on Ocean, Offshore and Arctic Engineering. Paper number OMAE2009-79443. [4] Campbell, M., Tognarelli, M. A., and Deka, D., 2010. “Drilling riser VIV—fact or fiction?”. In Proceedings, IADC/SPE SPE-129125-PP. [5] Shi, C., Manuel, L., and Tognarelli, M., 2011. “Empirical procedures for long-term prediction of fatigue damage for an instrumented marine riser”. In Proceedings, 30th Intl. Conf. on Offshore Mech. and Arctic Eng. Paper number OMAE2011-50235. [6] Tognarelli, M., Fontaine, E., Beynet, P., Santosa, M., and Marcollo, H., 2013. “Reliability-based factors of safety for vortex induced fatigue using field measurements”. Journal of Offshore Mechanics and Arctic Engineering, 135. [7] Fontaine, E., Marcollo, H., Vandiver, K., Triantafyllou, M., Larsen, C., Tognarelli, M., Constantinides, Y., and Oakley, O., 2011. “Reliability based factors of safety for VIV fatigue using NDP riser high mode VIV tests”. In Proceedings, 30th Intl. Conf. on Offshore Mech. and Arctic Eng. Paper number OMAE2011-49820. [8] Fontaine, E., Rosen, J., Marcollo, H., Vandiver, J., Triantafyllou, M., Resvanis, T., Larsen, C., Tognarelli, M., Oakley, O., Constantinides, Y., and Johnstone, D., 2013. “Using model test data to assess VIV factor of safety for SCR and TTR in GoM”. In Proceedings, 32th Intl. Conf. on Offshore Mech. and Arctic Eng. Paper number OMAE2013-10984. [9] Tognarelli, M. A., Panicker, N., Campbell, M., and Winterstein, S. R., 2013. “VIV safety factors for drilling risers: Propagating model uncertainty in a long-term reliability analysis”. In Proceedings, 32nd Intl. Conf. on Offshore Mech. and Arctic Eng. Paper number OMAE2013-10738. [10] Det Norske Veritas, 1984. Fatigue strength analysis for mobile offshore units. Classification note 30.2, DNV. [11] ASCE, 1982. “Committee on fatigue and fracture reliability of the committee on structural safety and reliability of the structural division: Fatigue reliability 1–4”. Journal of Structural Division, ASCE, 108(1), January, pp. 3–88. [12] Madsen, H. O., 1982. Deterministic and probabilistic models for damage accumulation due to time varying loading. Technical report DIALOG5-82, Danish Engineering Academy.

Finally, if we wish to use Fig. 13 in a design context to evaluate p f (Top ), it is useful to note that

Top B = E[T f ] γ∆

(25)

This result is derived briefly in Appendix II. It implies that the horizontal axis of Fig. 13 can be equivalently viewed as B/γ∆. Consider, for example, a typical wave-induced fatigue (WIF) case designed with γ=10, B=1 (unbiased software), and ∆=3 (between E and F2 S–N curves). In this case, the critical level B/γ∆ is 1/30 or about 0.03. This is close to the values found for both our WIF and VIV NS cases (Fig. 13). Substantially higher reliability can be achieved only if the software has significant conservative bias (as in the VIV GOM case), or if the safety factor γ is increased.

SUMMARY AND CONCLUSIONS We have presented a method to estimate fatigue reliability, using measured data to calibrate statistical uncertainty in fatigue software predictions. Applications to both VIV and waveinduced fatigue have been shown. The method shown here is a considerable extension of our previous work (e.g., [9]). First, a non-parametric model has been introduced, to reflect different software errors X for different levels of predicted damage D˙ pred . This is found useful to capture the general trend of increasing conservatism (lower X values) for higher D˙ pred (e.g., Fig. 2, Fig. 7, and Fig. 10). As a √ second modelling extension, we have included a ni n j term in Eqn. 6. This reflects our interest here in the variability of the long-run damage rate, not of individual damage rates in 10–30 -minute segments. Finally, we have compared the effect of different probability models—e.g., Figs. 4–5—and concluded that the lognormal is likely over-optimistic in these capacity-driven fatigue applications. The software bias B plays a critical role here. The software conservatism is greater for VIV applications than wave-induced fatigue—and this conservatism is especially large in the Gulf of Mexico example (see Eqn. 12 vs Eqn. 15 vs Eqn. 18). Using a Weibull model for T f in this case, even with no safety factor (γ=1) we have p f (Top ) on the order of 10−6 , well within the highest DNV safety class. In contrast, if the bias increases to B=0.1, one cannot achieve p f (Top )=10−5 even if γ=100 (Fig. 12). Of course, it should be noted that these methods depend directly on the quality of the data collected. Therefore, great care should be taken to assure the accuracy of the hardware and the calculation methodology used to infer fatigue damage. 9

c 2014 by ASME Copyright

1.6

of variation of W :

VW = Coefficient of Variation

1.4

1/2  1/2 (2p)! E[W 2 ] −1 = −1 VW = E[W ]2 p!2 

1.2

(29)

1

This is an implicit result for p, which in general requires numerical solution. Fortunately, a simple approximation is available:

0.8 0.6

VW ≈ p ; 0 ≤ p ≤ 1.5

0.4 0.2

Exact Vw = p

0 0

0.2

0.4 0.6 0.8 1 Weibull power-law parameter, p

1.2

(30)

Figure 14 shows that agreement is excellent for 0 ≤ p ≤ 1.5, which covers most practical cases of interest. Thus, in practice one first uses Eqn. 30 to estimate p, and then Eqn. 28 to find c.

1.4

FIGURE 14. Coefficient of variation, VW , of Weibull variable versus its power-law exponent, p.

APPENDIX II: DERIVATION OF EQUATION 25 First note that we can use Eqn. 21 to solve for D˙ LT :

[13] Det Norske Veritas, 2011. Fatigue design of offshore steel structures: Recommended practice. Technical report DNVRP-C203, DNV.

B D˙ LT = γTop

(31)

Substituting this result into Eqn. 7 yields APPENDIX I: WEIBULL PROBABILITY MODEL Since it forms the basis of much of our work here, we briefly describe some properties of the Weibull model. It is convenient to relate a Weibull random variable, W , to a standard (unit-mean) exponential random variable E: W = cE p

Tf =

∆γTop ∆ = B D˙ LT

(32)

Using the same approximations applied previously, the mean value of T f can be estimated as

(26) E[T f ] =

Here c and p are distribution parameters. Since P[E < x] is 1 − exp(−x), the cumulative distribution function of W is

∆γTop B

(33)

Algebraic manipulation of Eqn. 33 yields Eqn. 25. FW (w) = P[W < w] = 1 − exp[−(w/c)1/p ]

(27)

This result, with suitable choice of parameters c and p, has generated the failure probability curves in Fig. 3, Fig. 8, and Fig. 11. It remains to relate the parameters c and p to moments of W . For the standard exponential variable E, the expected value of E p is simply p!, or Γ(p + 1) for non-integral p. Therefore,

E[W ] = cp! → c =

E[W ] E[W ] = p! Γ(p + 1)

(28)

The unitless parameter p is in turn related to VW , the coefficient 10

c 2014 by ASME Copyright

Paper OMAE-2014-23234

Jun 8, 2014 - Measured damage rate versus predicted damage rate .... WIF plus VIV, Jackup production riser. 2.11 ... Flexcom plus SHEAR7(v4.6 default). 0.5.

498KB Sizes 3 Downloads 138 Views

Recommend Documents

paper
fingerprint, hand geometry, iris, keystroke, signature, voice and odour [1]. In recent ... As most of the cryptographic keys are lengthy and in random order, they are ...

paper
fingerprint, hand geometry, iris, keystroke, signature, voice and odour [1]. In recent years, biometric methods ... authentication [1]. Biometric cryptosystems are ...

paper
In this paper we discuss statistical methods for the investigation of cross- cultural differences ... The ACI data have been analyzed before to find the dimensional ...

Paper
the case where the cost of R&D for one firm is independent of its rivals' R&D .... type of market structure: in 2001, while only 0.2% of the IT service companies in.

Paper
swers has a dedicated category named poll and survey. In our approach, we collect poll and ..... question-answering server. In Proceedings of TREC. Hearst, M.

paper
Page 1 ... own architecture, which can be deployed across different designs by the use of a ... has its own tile and routing architectures, and supporting CAD.

paper
Jun 15, 2009 - In the early 1800s, property owners on New ... In this paper, we specifically consider the extent to which Business Improvement .... California cities finds that one-fifth of cities have BIDs; this number rises to one-half for cities o

Paper
Apr 3, 2007 - originally posted, accountancy age, david pescovitz, tom sanders, global ... howard schultz, shaun nichols, san francisco, de la, hong kong, ...

Paper
Abstract. This paper discusses an optimization of Dynamic Fuzzy Neural Net- .... criteria to generate neurons, learning principle, and pruning technology. Genetic.

Paper
Model-driven architecture (MDA) [1] is a discipline in software engineering .... for bi-directional tree transformations: a linguistic approach to the view update.

Paper
Aug 24, 2009 - Email: [email protected]. University of North ... Email: [email protected]. 1 ...... 0) = 1 a d Prob V = 0. ∣. ∣. ∣(Z.

paper
(1)Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr, Pasadena, CA, 91109, United. States - E-mail: [email protected].

Paper
Graduate School of Information Science and Technology. University of Tokyo, Hongo .... Languages (POPL), Long Beach, California. (2005) 233–246. 7. Jouault ...

Paper Title (use style: paper title) - Sites
Android application which is having higher graphics or rendering requirements. Graphics intensive applications such as games, internet browser and video ...

Paper Title (use style: paper title) - GitHub
points in a clustered data set which are least similar to other data points. ... data mining, clustering analysis in data flow environments .... large than the value of k.

Paper Title
Our current focus is on the molecular biology domain. In this paper .... (5) X (: a list of experimental results), indicating that Y .... protein names in biomedical text.

Non-paper
Oct 5, 2015 - integrating adaptation into relevant social, economic and environmental ... (b) Prioritizing action with respect to the people, places, ecosystems and sectors that are most vulnerable to ... 10. There shall be a high-level session on ad

paper - GitHub
LEVEL to~. UNIVRSU~Y OF MARYLAD. COMPUTER SCIENCE CETER. COLLEGE PARK, MARYLAD. Approved frpb~. budnUk~led10. 3 116 ...

[Paper Number]
duced-cost approaches to an ever growing variety of space missions [1]. .... each solar panel; bus and battery voltage; load currents; and main switch current.

Paper Title (use style: paper title)
College of Computer Science. Kookmin ... of the distinct words for clustering online news comments. In ... This work was supported by the Basic Science Research Program through .... is performed on class-wise reviews as depicted in Fig. 1(b).

Paper Title (use style: paper title)
School of Electrical Engineering, KAIST .... [Online]. Available: http://yann.lecun.com/exdb/mnist/. [5] Design Compiler User Guide, Synopsys, Mountain View, CA, ...

Paper Title (use style: paper title)
on the substrate, substrate pre-deposition process, and Pd deposition .... concentration is below the ignition threshold, which is often important for such a sensor.