Вы находитесь на странице: 1из 8

Nuclear Engineering and Design 240 (2010) 13551362

Contents lists available at ScienceDirect

Nuclear Engineering and Design


journal homepage: www.elsevier.com/locate/nucengdes

Thermal frequency response studies of a hollow cylinder subject to loads of different amplitude and shape
E. Paffumi a, , K.-F. Nilsson a , N.G. Taylor b
a b

Institute for Energy, European Commission Joint Research Centre, P.O. Box 2, 1755 ZG Petten, The Netherlands Institute for Energy, European Commission Joint Research Centre, Ispra (Va), Italy

a r t i c l e

i n f o

a b s t r a c t
Thermal fatigue is an important degradation mechanism, which must be considered in life management of nuclear plant piping systems. The analysis is very complex due to a number of complicating factors, with the determination of the load as the primary one. There is clearly a need for simplied engineering approaches, such as simulating the spectrum load by a constant frequency thermal load with the nominal temperature difference between mixing uids. The fatigue life is determined by the frequency that gives the shortest life using fatigue curves for initiation and Paris law for crack propagation. This paper analyses three aspects that affect the conservatism of such an approach: the thermal load shape (sinusoidal or square shaped), the structural boundary conditions at the edges of the modelled pipe (traction free or clamped) and the defect shape (circumferential or elliptic). Furthermore it is shown how the methodology can be used to determine screening criteria, i.e. a lower limit of the temperature difference below which there will be no component failure due to thermal fatigue. 2010 Elsevier B.V. All rights reserved.

Article history: Received 30 November 2009 Received in revised form 29 January 2010 Accepted 5 February 2010

1. Introduction During operation of light water reactors (LWR), temperature uctuations occur in many areas such as core outlet zone, lower part of hot pool, free surface of pool, secondary circuit and water/steam interface in steam generators. In certain conditions, these temperature uctuations can lead to thermo-mechanical damage and component failure (IAEA, 2002; Chapuliot et al., 2005). Thermal fatigue is also an important issue for liquid metal fast rectors (LMFR) where the temperatures are higher than for LWR and the properties of the coolant may infer additional problems (IAEA, 2002; Hu et al., 2004). Calculation of the thermal fatigue damage and the associated fatigue life are difcult and prone to large uncertainties due to the complexity of the phenomena involved. The temperature uctuations can be local or global and induce spectrum loads at the wall (Hu et al., 2004; Paffumi et al., 2008a). These resulting thermal stresses and strain variations result in surface degradation followed by formation of surface cracks and crack growth that eventually may lead to component failure. In addition to the amplitude of the thermal loads, the degradation depends also strongly on their frequency. At very high frequencies the thermal stresses are conned to the surface and no deep cracks develop. The thermal stresses are low for very low frequencies as the temperature gradients

Corresponding author. Tel.: +31 224 565082; fax: +31 224 565641. E-mail address: elena.paffumi@jrc.nl (E. Paffumi). 0029-5493/$ see front matter 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.nucengdes.2010.02.036

through the wall are small (Kasahara et al., 2002; Boley and Weiner, 1960). Hence there is an intermediate frequency range for which most fatigue damage is expected. Numerical simulations of thermal stripping and high-cycle thermal fatigue of tee junctions of LWR piping systems imply that the fastest crack initiation and propagation occur for frequencies in the range 0.11 Hz (Hu et al., 2004; Paffumi et al., 2008a; Dahlberg et al., 2007; Buckthorpe et al., 1988; Lee et al., 1999). Much effort continues to be devoted to experimental studies and development of models with different levels of complexity (Paffumi et al., 2008a, 2005a,b, 2004; Ancelet et al., 2007; Haddar et al., 2005). Given the importance of thermal fatigue for the safe and economical operation of nuclear power plants as well as other process industries, there is a general need to improve engineering tools and predictions techniques that capture the main characteristics of thermal fatigue and associated experimental studies that support model development and increase the understanding of the relevant phenomena (Hu et al., 2004; Paffumi et al., 2008a,b; Green, 1985; Jones and Lewis, 1994; Jones, 1997; Paffumi, 2004). The loading is the most important unknown factor for a reliable assessment and it is very difcult to compute or measure the load spectrum for a given situation. The sinusoidal approach (SIN-method), where the thermal loads are assumed to vary sinusoidally (Radu et al., 2007, 2008, 2009) with the frequency that gives the shortest life for a given nominal temperature difference between the mixing uids, is one of the simplest methods. In Europe the SIN-method has been the basis for the development of a European Procedure for thermal fatigue in nuclear components (Faidy, 2007). As a rst step

1356

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362 Table 1 316L temperature dependence material data. Temperature, C Youngs modulus, MPa Poisson ratio Density, kg/m3 Thermal conductivity, W/mK Thermal expansion, 1/K Specic heat, J/kg K 20 192e3 0.3 8000 14.5 15.16e6 480 300 170e3 0.3 7870 18 18.92e6 550 500 153e3 0.3 7780 20 20.36e6 580 700 137e3 0.3 7680 23 21.28e6 600

Nomenclature a E N, n K K Keff T Tm T crack depth Youngs modulus thermal expansion coefcient number of cycles stress intensity factor stress intensity factor range effective stress intensity factor range temperature mean temperature temperature range variation stress yield stress o x distance across the wall thickness from the inner pipe surface t wall thickness of the cylinder ri inner radius of the cylinder ro outer radius of the cylinder f frequency Poisson ratio thermal conductivity LWR light water reactor LMFR liquid metal fast reactor NESC network for evaluation of structural components NULIFE nuclear plant life prediction-network of excellence TF thermal fatigue FEA nite element analysis SIN-method sinusoidal method

2. Model description The pipe has an inner radius is ri 120 mm and outer radius is = ro = 129 mm, which is a representative pipe geometry (Chapuliot et al., 2005). The material is 316L stainless steel and the physical material properties at different temperatures are given in Table 1. The analysis is performed in two steps. First the through-wall temperature distribution is computed and the resulting stresses and strains are calculated in a subsequent stressstrain analysis. For an elastic case with an uncracked body and axisymmetric loading and no temperature dependence of the material properties, the time dependent temperature and stress distributions for a given load can be computed analytically, e.g. (Radu et al., 2007, 2008). In more complicated cases, for instance involving temperature dependence of material properties, plasticity effects, cracked body analysis or complex thermal loads or boundary conditions, a nite element analysis is required. In this paper we have adopted the nite element method with the commercial code ABAQUS as it gives more versatility for assessment of different parameters and modelling conditions. The analysis is performed using axisymmetric 8-node elements for both the thermal and stress analysis. Fig. 1 shows the nite element mesh. Symmetry conditions are imposed at the lower edge whereas the upper edge is either traction free (free end) or with zero axial displacements (xed end) (Fig. 1a and b). The nite element mesh has a gradual renement from the pipes outside toward the inner surface to capture the large thermal and stress variations at the inside. The inside of the pipe is subject to a

the SIN-method was applied using a fatigue curve approach and surface stresses to assess crack initiation in a project under the auspices of the Network for Evaluation of Structural Components (NESC) (Dahlberg et al., 2007; Faidy, 2007). Crack propagation with the SIN-method is more complicated than crack initiation since the stress distribution and the stress intensity factors need to be computed as function of crack depth. In the European Network NULIFE, a procedure based on Paris law and the SIN-method was proposed (Paffumi and Radu, 2009). The SIN-method is generally assumed to provide conservative estimates for both the crack initiation and crack propagation life. The analysis uses the nominal difference between the mixing uids and a conservative value is assigned for the heat transfer coefcient. The mechanical stresses result from the through-wall thermal gradients and the geometrical constraints are therefore displacement controlled and an elastic analysis therefore provides a conservative estimate. But the most important conservatism is that the most damaging frequency is used for the loading. There are, however, also a number of assumptions that are not necessarily conservative. The thermal variation at the pipe wall may be more abrupt than the sinusoidal function and a square shaped function, which could be more representative, is expected to give higher stresses. The stressstrain analysis is performed for a pipe segment which should be representative for the loaded section of the piping system but the boundary condition that is adopted for the pipe segment will affect the computed stresses and associated fracture parameters. In this paper we investigate how the assumed shape of the thermal load and the boundary conditions inuence the crack propagation life using the SIN-method. The thermal load shape is assessed by comparing a sinusoidal and a squared thermal load variation whereas boundary condition effect is assessed by comparing pipe segment with clamped ends (upper bound for the constraint) and with traction free ends (lower bound for the constraint).

Fig. 1. FE model for thermal and stress analysis: mesh and the boundary conditions. Specimen completely free to expand (a) or xed in the axial direction (b) with symmetry condition at the symmetry line. Thermal load applied at the inner surface as shown.

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362

1357

periodic thermal load that is either sinusoidal or square shaped:

sin(2 ft) Tsin (t)=Tm + T /2 N/f , T (H(ttk 1/f )H(ttk )) Tsq (t)=Tm + T /2+
tk =1/f

(1)

where H represents the Heaviside function, f is the frequency, N the number of cycles within the time considered and Tm the mean temperature. The temperature transferred from a uid to the structure depends on the heat transfer coefcient. The impact of the heat transfer coefcient, h, is measured by the Biot number, ht/ , where t is the wall thickness and is the thermal conductivity. The Biot number may vary in the range 550 (Dahlberg et al., 2007). The inuence of the heat transfer increases with the frequency. For 0.1 Hz the effect is very small whereas for 10 Hz the stresses may vary by a factor two for the Biot number range considered (Dahlberg et al., 2007). The effect on the fatigue life would be even higher since crack initiation and propagation depend on the stresses through a power law relationship. In this analysis the wall temperature follows exactly the uid temperature, which corresponds to no heat transfer loss between the uid and the wall and to an innite heat transfer coefcient. This is a conservative assumption but does not affect the purpose of our analysis at relatively low frequencies. The outer surface has zero heat ux. The loading frequency varied from 0.06 to 6 Hz and the temperatures T applied for each given frequency of load were: 100, 150, 200, 250 and 300 C. The temperature range is applied to the surface so that the maximum temperature for each case always corresponds to 300 C. The stress free temperature is set to the mean temperature Tm . The stress intensity factors used for the crack propagation analyses below were calculated from the through-wall stress distributions using Handbook solutions (Marie et al., 2007; Chapuliot, 2000) from the general formula: K= a x x b0 i0 +b1 i1 +b2 i2 Q t t
2

temperature dependent properties given in Table 1 and linear interpolation. Calculations assuming no temperature dependence were also performed with the material properties at the mean temperature value for 20 and 300 C. The resulting stresses were lower than for temperature dependent material properties cases but the effect was small (typically 10%). Due to the stress free condition at Tm , the stresses are symmetric with respect to the zero stress. As expected the maximum stress range occurs at the free surface and it is almost unaffected by the frequency for the sinusoidal load cases, whereas it decreases slightly with the frequency for the squared load cases. The stress ranges are somewhat higher for the squared shape thermal load in particular for the lower frequencies. This means that, in a fatigue curve approach, for the same T a crack can initiate with fewer cycles with a square thermal loads (Radu et al., 2009; Faidy, 2007). It is also quite clear that the stress gradient close to the free surface increases with the frequency. The stresses become very low after a short distance from the inner wall for frequencies above 1 Hz, whereas the stress range at deeper depths increases with lower frequency. The stress ranges at deeper depths are generally higher for the xed end case than for the free end case. For the sinusoidal load and low frequency the stress range is however higher at the outer surface with the free end. The time needed for the thermal load to cross the thickness of the cylinder by conduction is around 1.75 s (Paffumi, 2004), which corresponds to a frequency of about 0.6 Hz. The high frequency 6 Hz has a period of about 0.16 s, which is quite small compared to the time it takes to conduct the heat through the thickness and in that case there will only be large stress gradients and stresses close to the free surface. The low frequency, 0.06 Hz, corresponds to a period of 16.6 s, which is much longer than 1.75 s. If the frequency is sufciently low the stress range will be constant through the thickness, because of thermal homogenisation (Kasahara et al., 2002; Boley and Weiner, 1960), and in fact it would be zero for the free edge. 3.2. Stress intensity factor

+b3 i3

x t

+b4 i4

x t

(2) Crack growth in this paper is assumed to be governed by Paris law. The crack growth rate is related to the stress intensity factor range ( K) for the postulated crack geometry. The K for the elastic cases were determined for the different amplitudes of load and frequencies using the handbook solutions (Chapuliot, 2000) (Eq. (2)) directly and the time dependent stress distributions (Paffumi et al., 2008a). The results in this section are all derived for a circumferential crack. As the mean temperature is stress free, the computed stress intensity factors will be both positive and negative. When applying Paris law only the range when the crack is open is used (the effective stress intensity range Keff ). In absence of plastic residual stresses this means that Keff = Kmax (Itatani et al., 2001). The effective stress intensity factor range is plotted for ve different frequencies in Fig. 3ad, for the two load shape functions and pipe end conditions for the temperature range T = 150 C. The shape of the applied thermal load function and the boundary condition have a much larger impact on the stress intensity factor range than on the stress range. The stress intensity factor range increase with lower frequencies. For crack lengths below 0.5 mm this frequency effect is small whereas for deeper cracks the stress intensity factor range depends strongly on the frequency but also on the boundary conditions and the shape of the thermal load function. For both the free and xed edge cases, the squared thermal load function gives stress intensity factor ranges that are consistently higher than for the sinusoidal. It can also be noted that for high frequencies there is a crack depth for which the effective stress intensity factor has a peak value. As the frequency decreases the crack depth with the peak value increases and the peak becomes less pronounced. For the xed end conditions the stress intensity range increases

where x is a variable indicting the distance across the wall thickness, t; b0 , b1 , b2 , b3 and b4 are coefcients for the polynomial: = b0 + b1 (x/t) + b2 (x/t)2 + b3 (x/t)3 + b4 (x/t)4 , which tted to the stress distribution through the wall thickness for 0 x a, with a crack depth; i0 , i1 , i2 , i3 and i4 are inuence coefcients which depend on the crack/thickness ratio, a/t, on the shape of the crack, a/c, and on the location along the crack front, and are given by handbook solutions (Marie et al., 2007; Chapuliot, 2000); Q is a shape correction factor, dened for a semi-elliptical crack as (Chapuliot, 2000): Q = 1 + 1.464(a/c)1.65 , a/c < 1. For each T cycle, the analysis was run for 10 full cycles. 3. Analysis The thermal loads result in a bi-axial stress eld at the inner surface for the free edge case. For the constrained edge case the axial stress at the inner surface will be somewhat higher than the hoop stress. Before performing the simplied fatigue life analysis it is instructive to have a closer look at the resulting stresses and stress intensity factors. 3.1. Stress distributions Fig. 2ad shows the computed stress range for four frequencies, for the sinusoidal and squared thermal load functions and for the free and constrained edge boundary conditions for temperature range T = 150 C. These analyses were all performed with the

1358

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362

Fig. 2. Range of axial stress across the wall thickness for T equal to 150 C and different loading frequencies: (a) sinusoidal load and top free boundary conditions, (b) sinusoidal load and top xed boundary conditions, (c) square load and top free boundary conditions, and (d) square load and top xed boundary conditions.

Fig. 3. Stress intensity factor range as function of crack depth for T equal to 150 C and different loading frequencies: (a) sinusoidal load and top free boundary conditions, (b) sinusoidal load and top xed boundary conditions, (c) square load and top free boundary conditions, and (d) square load and top xed boundary conditions.

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362

1359

Fig. 4. Fatigue crack growth lifetime in hours in function of frequency for different temperature ranges and sinusoidal load up to 1 Hz for the free boundary conditions (a) and for a temperature range equal to T = 150 C and different shape of the load up to 2 Hz, for both free and xed boundary conditions (b).

monotonically with crack depth for the lower frequencies. The analyses in Fig. 3 were performed with temperature dependent material properties. Analysis with no temperature dependence and properties at the mean value of the temperature range 20300 C resulted in about 10% lower values for the K range. 3.3. Lifetime assessment The crack growth rate is assumed to follow Paris law: da n = C0 ( KI ) . dN (3)

For austenitic steels the ASME standard (ASME, 2004b) recommends n = 3.3 and C0 = CS for crack growth rates in units of mm/cycle and KI in units of MPa m. The parameter C accounts but the temperature dependence for temperature effects by C = 10 is rather small in this temperature range ( equal to 8.4520 at 300 C and 8.6075 at 100 C and C differs by a factor 1.4). S is used to account for R-ratio effects; S = 1.0 when R 0, which is the case in the present assessment. The number of cycles Np required for a crack to advance from ai to af follows directly from integration of Eq. (3):
af

Fig. 4a scale almost perfectly with Tn and we also note that the critical frequency is temperature independent, indicating that the material property dependence is very weak for the elastic analysis. Fig. 4b allows us to compare the crack propagation life for the two end conditions of the pipe and the two thermal load functions. The crack propagation life is shorter for the square load than for the sinusoidal load and the xed end condition gives a shorter life than the free end condition. The critical frequency is about 0.2 Hz for the sinusoidal and 0.3 Hz for the square load. It should also be noted that the minimum in the crack propagation curves are relatively shallow and the crack propagation life does not differ more than a factor 2 in the frequency range between 0.1 and 1 Hz for the different cases. The critical frequency depends on the material parameters, such as the heat conductivity, but the difference is small between different steels. In fact 0.2 Hz was also found for 304L (Radu et al., 2009). Crack propagation life curves versus the temperature amplitude can be computed directly from the propagation life curve at a given frequency for different load cases and boundary conditions. In particular if we neglect the temperature dependence of the material parameters, which in any case is small as mentioned above, then the crack propagation life can be computed from the result for one

Np ( T, f ) =
ai

da C0 ( KI ( T, f ; a))

n.

(4)

Eq. (4) is integrated from an initial crack depth, ai = 0.9 mm, which corresponds to 10% of the wall thickness, to a nal crack depth af = 7.2 mm, which corresponds to 80% of the wall thickness. The corresponding crack growth in time (hours) is given by Dcg ( T, f ) = Np ( T, f ) . 3600 f (5)

The frequency at which the shortest life is obtained is referred to as the critical crack growth frequency, fcr (Radu et al., 2009) and is found by minimizing Eq. (5). Fig. 4a depicts the fatigue crack growth lifetime in hours as function of the frequency for different temperature ranges for the sinusoidal load and the free edge condition using the temperature dependence of the material properties. Fig. 4b shows the fatigue crack growth lifetime in hours versus the frequency for T = 150 C and for the two boundary conditions and two load cases. If there is no temperature dependence of the material properties then the stress intensity range is directly proportional to the temperature range ( K T ) and consequently the crack growth rate da/dN is proportional to Tn , the propagation life, Dcg , scales with Tn and the critical frequency is independent of T. The different curves in

Fig. 5. Fatigue life curve versus T for the two shapes of the loading function, sinusoidal and square and for the corresponding critical frequency 0.2 and 0.3 Hz. Specimen free to expand in the axial direction.

1360

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362

Fig. 6. Fatigue life curve versus

T with and without considering the threshold

Kth equal to 5 MPa m. Specimen free to expand in the axial direction.

temperature range,

T*, by a simple scaling procedure: T T


n

Dcg ( T, f ) = Dcg ( T , f )

(6)

Fig. 5 shows the resulting fatigue life curves for the squared and sinusoidal load cases and the two end conditions at the critical frequencies 0.2 and 0.3 Hz, respectively. It can be noted that the propagation life is shorter for the square load function and the clamped boundary condition reduces the lifetime further, but there is not a dramatic difference in the crack propagation life for the two boundary conditions. Most of the life is spent in the propagation of short cracks. As seen in Fig. 3, the difference in Keff between the two end conditions increases with crack depth. Thus the nal stage of crack propagation prior to component failure would be much shorter for the xed end, which should be considered in a safety assessment. Fig. 6a shows the resulting crack propagation curves for the sinusoidal load with free end condition for different frequencies in the range 0.066 Hz. In constant amplitude tests there is often a K ( Kth ) below which the crack propagation is small enough (typically 108 mm/cycle) to be assumed zero. Such threshold value depends on the R-ratio but 5 MPa m is a typical value for R = 0 (ASME, 2004a,b). If the thermal load is sufciently low then K will fall below this value and the crack would arrest. This would correspond to a threshold value for the temperature difference, Tth . It was shown in Fig. 3 that K for a given T decreases with increasing frequency and that there is a certain depth at which the peak K is attained. From Fig. 3 it is possible to estimate the T at which the peak values of K equal the threshold value by the scaling K( T ) = K( T ) ( T / T ). Fig. 6b shows the same crack propagation curves as in Fig. 6a, but indicating the corresponding threshold T. A relation between Tth and crack depth can be directly obtained from the K curves. Fig. 7 depicts this relationship for the sinusoidal load case for the free and constrained edge for 0.06, 0.3, 1 and 6 Hz and shows the maximum allowable temperature range for the growth of a defect versus crack depth. The T needed to reach the Kth increases with frequency and for xed end conditions, smaller Ts are needed to fulll this condition. In the NESC Thermal Fatigue Procedure (Dahlberg et al., 2007) a screening criterion for the temperature difference between mixing uids was adopted below which there would be no thermal fatigue failure. This value was set to 80 C and was based on industrial experience. The results in Fig. 7 suggest a value around 50 C for a critical frequency of 0.3 Hz. The local temperature variation of the uid at the pipe wall is always lower that the nominal difference between the mixing uids. In Dahlberg et al. (2007) it was proposed that 80% of the nominal temperature would still be on the conservative side. Moreover the actual temperature variation

in the pipe wall is still lower due to the heat transfer coefcient, than in the analysis where a perfect heat transfer is assumed. Thus a computed value of 50 C with the conservative assumptions in the analysis is well in line with the suggested threshold of 80 C. Of course a threshold value for the temperature range assumes that there is a threshold value for stress intensity factor range. Physically a threshold value is related to crack tip closure, which is affected by residual crack tip stresses and oxidation of crack surfaces. For constant amplitude there are ample data to support a threshold value for crack arrest, although the actual value depends on for instance the R-ratio. For variable amplitude loading the picture is less clear and there are data that suggest that there is no threshold value (Ohji et al., 1988). A threshold value may also depend on the environment where for instance the threshold may increase at elevated temperatures due to formation of oxidation products on the crack surfaces (Nishikawa et al., 1987). In the analyses above the cracks were circumferential. This is a conservative estimate. Service experience as summarised in (Gosselin et al., 2007) showed that typically a large number of fatigue cracks initiate at multiple sites and then link together to form a single fatigue crack that is much longer than the standard 6:1 aspect ratio aws used by ASME Section XI for damage tolerance calculations. In ASME Section XI Appendix L aw tolerance assessments, a value of 2c/a = 30 ratio is proposed for thermal fatigue to

Fig. 7. Threshold temperature range values as a function of crack depth for different striping frequencies and Kth equal to 5 MPa m. Specimen free and xed in the axial direction.

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362

1361

Fig. 8. Stress intensity factor range as function of crack depth for two different crack aspect ratios: c = 16a and fully circumferential cracks. Thermal sinusoidal load with maximum temperature range of 150 C and different frequencies. Specimen free to expand in the axial direction (a) and xed (b).

4. Conclusions The assessment of fatigue crack growth due to turbulent mixing presents signicant challenges, in particular to determine the thermal loading spectrum. The sinusoidal method is a simplied approach for addressing this problem, in which the entire spectrum is replaced by a sine-wave variation of the temperature at the inner pipe surface. Such estimates are generally intended to be conservative but not unduly conservative. In general any constant shape function could be considered in this engineering approach. In this work sinusoidal and square thermal loading functions of different frequencies and amplitudes have been considered together with the two boundary conditions, cylinder free to expand in the axial direction or completely restrained. The stress variation at the surface is directly related to the crack initiation estimate, while the stress intensity factor distributions are correlated to the crack growth life and applying different loading functions. A perfect heat transfer from uid to the pipe has been assumed. This is a conservative assumption but does not affect the results of our analysis at relatively low frequencies.

Fig. 9. Temperature range values as a function of the life in hours for the critical frequency 0.2 Hz for the two shapes of the crack, fully circumferential and c = 16a, and the two boundary conditions; sample free to expand or fully constrained in the axial direction.

account for coalescence of small cracks from multiple initiation and representative for thermal fatigue in mixing conditions. Stress intensity factor range versus crack depth are shown in Fig. 8 for a circumferential and a semi-elliptical crack with aspect ratio c = 16a, where 2c is the length and a the depth of the crack. The stress intensity factors in Fig. 8 have been calculated for a sinusoidal load at the surface with maximum temperature range of 150 C, different frequencies and the two boundary conditions. The fully circumferential crack gives a higher value for the stress intensity range but the difference is not very large (maximum a factor 1.2 at deepest cracks and higher K). The same is valid for the square loading function not reported here. Fig. 9 compares the fatigue crack growth lifetime in hours for different Ts for the two crack aspect ratio, fully circumferential and c = 16a, for both free and xed end using the sinusoidal approach. The fully circumferential shape is slightly more conservative than the semi-elliptical shape with the ratio c = 16a. The aspect ratio generally increases with crack depth and the value 2c/a = 30 is typical for a deep crack and moreover the crack growth rate increases with the aspect ratio (Paffumi et al., 2008a). Thus taking a xed aspect ratio of 30 gives a conservative life prediction. If a very low aspect ratio, representative for a shallow crack, were adopted, then the analysis could be non-conservative (Paffumi et al., 2008a,b).

The loading functions shapes and boundary conditions inuence more the crack growth behaviour than the crack initiation and the square load gives a shorter life than the sinusoidal load (typically a factor 2). In case of a rapid transient variation of the temperature the square load may be a more appropriate representation. The boundary conditions have also an inuence on the crack growth life. Over-conservative estimates may be obtained for the completely restrained component. The elastic analyses were performed considering the temperature dependences of the material properties. Their effect on the life estimates is negligible and hence can be neglected in an engineering approach. A threshold of 5 MPa m in Paris law suggests a possible screen C for the temperature difference below which ing criteria of 50 there is no crack growth, in line with the 80 C suggested by industrial experience high cycles thermal fatigue in turbulent mixing (Dahlberg et al., 2007), considering the conservative assumptions of these present analyses. A circumferential crack gives a conservative, but not unduly conservative, estimate of the crack propagation life with respect to the ASTM recommendation of a semi-elliptical crack with aspect ratio equal to 30.

1362

E. Paffumi et al. / Nuclear Engineering and Design 240 (2010) 13551362 Jones, I.S., 1997. The frequency response model of thermal striping for cylindrical geometries. Fatigue & Fracture of Engineering Materials & Structures 20 (6), 871882. Jones, I.S., Lewis, M.W.J., 1994. A frequency response method for calculating stress intensity factors due to thermal striping loads. Fatigue & Fracture of Engineering Materials & Structures 17 (6), 709720. Kasahara, N., Takasho, H., Yacumpai, A., 2002. Structural response function approach for evaluation of thermal stripping phenomena. Nuclear Engineering and Design 212, 281292. Lee, H.-Y., Kim, J.-B., Yoo, B., 1999. Tee-junction of LMFR secondary circuit involving thermal, thermomechanical and fracture mechanics assessment on a stripping phenomenon, IAEA-TECDOC-1318. Validation of fast reactor thermomechanical and thermohydraulic codes. Final report of a coordinated research project 19961999. Marie, S., Chapuliot, S., Kayser, Y., Lacire, M.H., Drubay, B., Barthelet, B., Le Delliou, P., Rougier, V., Naudin, C., Gilles, P., Triay, M., 2007. French RSE-M and RCCMR code appendices for aw analysis: presentation of the fracture parameters calculation. Part III: cracked pipes. International Journal of Pressure Vessels and Piping 84, 614658. Nishikawa, I., et al., 1987. The role of crack closure in the fatigue threshold at elevated temperature. Transaction of Japan Society of Mechanical Engineers 53490A, 993 (in Japanese). Ohji, K., et al., 1988. Fatigue crack growth and its threshold in the creep regime. Journal of the Society of Materials Science 37 (419), 951 (in Japanese). Paffumi, E., 2004. Simulation and Modelling of Thermal Fatigue Damage in Austenitic Piping Components. Ph.D. Thesis. University of Swansea, UK. Paffumi, E., Radu, V., 2009. Status on the knowledge on cracks evolution under loadings from a thermal spectrum. Crack propagation and possible arrest/penetration, NULIFE (09) 10, JRC53157, Scientic and Technical Report, April, 2009. Paffumi, E., Nilsson, K.F., Taylor, N., Hurst, R., Bache, M.R., 2004. Measurement of shallow crack growth during thermal fatigue of 316L tubular test piece. In: Proceedings of the 3rd International Conference on Fatigue of Reactor Components, OECD/EPRI Seville, 36 October. Paffumi, E., Nilsson, K.F., Taylor, N., Hurst, R., Bache, M.R., 2005a. Cracks initiation, propagation and arrest in 316L model pipe components under thermal fatigue. Journal of ASTM International 2 (5 (May)). Paffumi, E., Nilsson, K.F., Taylor, N.G., 2005b. Analysis of the thermal fatigue cracking of 316L model pipe components under cyclic down-shocks. In: McCabe, J. (Ed.), Design and Analysis of Piping Components, Thermal Stresses and CFD in Vessel Design. Proceedings of PVP2005. Denver, Colorado, USA, July, 1721, 2005. Paffumi, E., Nilsson, K.-F., Taylor, N.G., 2008a. Simulation of thermal fatigue damage in a 3156L model pipe component. International Journal of Pressure Vessel and Piping 85, 798813. Paffumi, E., Nilsson, K.-F., Taylor, N.G., 2008b. Thermal fatigue cyclic-down shocks on 316L model pipe components. In: Proceedings of PVP2008, 2008 ASME Pressure Vessels and Piping Division Conference, Chicago, IL, USA, July, 2731. Radu, V., Paffumi, E., Taylor, N., 2007. New analytical stress formulae for arbitrary time dependent thermal loads in pipes. European Commission Report EUR 22802 DG JRC, Petten, NL, June, 2007. Radu, V., Paffumi, E., Taylor, N., 2008. Development of new analytical solutions for elastic thermal stress components in a hollow cylinder under sinusoidal transient thermal loading. International Journal of Pressure Vessels and Piping 85, 885893. Radu, V., Paffumi, E., Taylor, N., Nilsson, K.F., 2009. A study on fatigue crack growth in high cycle domain assuming sinusoidal thermal loading. International Journal of Pressure Vessels and Piping 86 (12), 818829.

The critical frequency is typically 0.20.4 Hz and the crack propagation life is only weakly affected by critical frequencies in the range 0.11 Hz. The crack propagation life is also weakly affected by the material temperature dependence. The two observations suggest that the fatigue life as function of the temperature range, T, can be derived from one analysis with a xed frequency, e.g. 0.3 Hz and one temperature range, T*. The complete curve is then obtained by simple scaling the crack propagation life n Dcg ( T, f ) = Dcg ( T , f ) ( T / T ) . References
Ancelet, O., Chapuliot, S., Henaff, G., Marie, S., 2007. Development of a test for the analyses of the harmfulness of 3D thermal fatigue loading in tubes. International Journal of Fatigue 29, 549564. ASME, 2004a, Boiler and Pressure Vessel Code, Section III ComponentsRules for Construction of Nuclear Power Plant, American Society of Mechanical Engineers, New York, 2004. ASME, 2004b, Boiler and Pressure Vessel Code XI, Rules for Inservice Inspection of Nuclear Power Plant Components, American Society of Mechanical Engineers, New York, 2004. Boley, B.A., Weiner, J., 1960. Theory of Thermal Stresses. John Wiley & Sons. Buckthorpe, D., Gelineau, O., Lewis, M.W.J., Ponter, A., Final report on CEC study on thermal stripping benchmarkthermo mechanical and fracture calculation, Project C5077/TR/001, NNC Limited 1988. Chapuliot, S., 2000. Formulaire de KI Pour les Tubes Comportant un Defaut de Surface Semi-elliptique. In: Longitudinal ou Circonferentiel, interne ou externe, Direction de linformation scientique et technique, CEA/SACLAY-R-5900, France, ISSN 0429 3460. Chapuliot, S., Gourdin, C., Payen, T., Magnaud, J.P., Monavon, A., 2005. Hydrothermal-mechanical analysis of thermal fatigue in a mixing tee. Nuclear Engineering and Design 235, 575596. Dahlberg, M., Nilsson, F., Taylor, N., Faidy, C., Wilke, U., Chapuliot, S., Kalkhof, D., Bretherton, I., Church, M., Solin, J., Catalano, J., 2007. Development of a European procedure for assessment of high cycle thermal fatigue in light water reactors: nal report of the NESC-Thermal fatigue project, NESC Network Report NESC06-04, Published as European Commission EUR 22763 EN June 2007. Faidy, C., 2007, European procedure for thermal fatigue analysis of mixing Tees. Prepared by EDF-SEPTEN France, with the contribution of: Sweden, PSI Switzerland, JRC EC, CEA-Saclay France, UK, EON Germany, TRACTEBEL Belgique, Finland, Netherlands. Gosselin, S.R., Simonen, F.A., Heasler, P.G., Doctor, S.R., 2007. Fatigue Crack Flaw Tolerance in Nuclear Power Plant Piping; A basis for Improvements to ASME Code Section XI Appendix L, NUREG/CR-6934, PNNL-16192, May. Green, D., 1985. Crack growth under constant amplitude rapid thermal striping, thermal fatigue and thermal striping. In: ASME Pressure Vessel and Piping Conference, New Orleans. Haddar, N., Fissolo, A., Maillot, V., 2005. Thermal fatigue crack networks: a computational study. International Journal of Solids and Structures 42, 771788. Hu, L.-W., Lee, J., Saha, P., Kazimi, M.S., 2004. Numerical simulation study of high thermal fatigue caused by thermal stripping. In: Third International Conference on Fatigue of Reactor Components, Seville, Spain, October, 36, 2004, NEA/CSNI/R, p. 21. International Atomic Energy Agency, 2002. Validation of Fast Reactor Thermomechanical and Thermohydraulic Codes, IAEA-TECDOC-1318. IAEA, Vienna. Itatani, M., Asano, M., Kikuchi, Suzuchi, S., Lida, K., 2001. Fatigue crack growth curve for austenitic stainless steels in BWR environment. ASME JPVT 123, 166, doi:10.1115/1.1358841.

Вам также может понравиться