You are on page 1of 12

Nuclear Engineering and Design 239 (2009) 2754–2765

Contents lists available at ScienceDirect

Nuclear Engineering and Design


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

Thermal striping limits for components of sodium cooled fast spectrum reactors
P. Chellapandi ∗ , S.C. Chetal, Baldev Raj
Indira Gandhi Centre for Atomic Research, Kalpakkam-603 102, Tamil Nadu, India

a r t i c l e i n f o a b s t r a c t

Article history: With the objective of establishing thermal striping limits for future sodium cooled fast spectrum reac-
Received 11 March 2009 tors (SFR), a fracture mechanics-based method employing ‘␴-d approach’ recommended in RCC-MR:
Received in revised form 7 August 2009 Appendix A16 has been followed. Towards this, an idealized geometry, thermal fluctuations in the form
Accepted 14 August 2009
of constant power spectral density and pessimistic material data were considered and temperature and
thermal stresses are computed taking in to account frequency-dependent thermal attenuation on the
structural wall. The effect of attenuation is found to be significant. The limits are derived at various
potential locations in SFRs, which are also subjected to creep-fatigue damage due to major cycles caused
by startup, shutdown, power failures and pump trips, etc. The maximum range of temperature fluctua-
tions can be as high as 70 K where there is practically no accumulated creep-fatigue damage and 45 K is
acceptable where the creep-fatigue is significant (0.9). These limits are found to be consistent with the
broad limits extrapolated from the failure experiences of international SFRs and sodium facilities. Pool
hydraulic computations carried out to identify and quantify the thermal striping zones confirmed that
the proposed limits can be respected with good margins for SFRs.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction In the pool type concept, both hot sodium pool which is at about
820 K and cold pool at about 670 K co-exist, which imposes high
Liquid sodium is the preferred coolant in almost all liquid metal T (150 K-maximum) in sodium in the narrow transition regions
cooled fast spectrum reactors (SFR). High boiling point implies that of hot and cold pools, during normal operating as well as transient
sodium remains in liquid state up to the temperature of 1175 K conditions. This is termed as ‘thermal stratification’. With the ASS
at atmospheric pressure. Excellent heat transfer characteristics of as structural material which has low thermal conductivity and high
sodium are advantageous. Higher boiling point permits high oper- coefficient of thermal expansion and the sodium with its inherently
ating temperature for the reactor, still ensuring sufficiently high high heat transfer coefficient, the adjoining structural wall surface
margin to avoid boiling of the coolant under all the design basis is subjected to high T, created in sodium without any significant
events. Sodium remains in liquid state during operating conditions film drop and time delay. This causes high thermal stress range
without calling for any pressurization and hence design pressure () in the structural wall. Further concern of thermal stratification
for components is nearly atmospheric, in turn requiring lower wall is steady oscillations, relatively at lower frequencies (<1 Hz), which
thickness for structures. The excellent heat transfer properties pro- is one of the sources of high cycle thermal fatigue damage for the
vide high natural heat removal capability, particularly in the pool metal wall. Apart from this mechanism, high thermal fatigue cycles
type concept. While pool type concept has many distinct advan- are caused by a special phenomenon called ‘thermal striping’. Ther-
tages from the point of view of safety, such as high thermal capacity mal striping is a complex thermal hydraulics phenomenon, which
to accommodate thermal transients without significant tempera- generates random fast temperature fluctuations, originating from
ture rise and natural convection capability, there are certain critical the incomplete mixing of hot and cold jets of fluid, sodium in the
structural mechanics issues, especially with the austenitic stain- present context, in the vicinity of adjoining structural wall surface
less steels (ASS), commonly used structural materials in view of (Fig. 1). Thermal striping occurs at a few locations in the hot and
their good compatibility with sodium and high strength at ele- cold sodium pools in the reactor assembly, predominantly on the
vated temperatures. These problems are addressed in this paper core cover plate of control plug and at mixing ‘Tee’ junctions in
with reference to a typical 500 MWe capacity SFR. the secondary sodium pipelines. It is worth mentioning that apart
from thermal striping, oscillations of thermal stratification layers
and sodium free level do cause temperature fluctuations in certain
locations in the sodium pools. Fig. 2 shows a few possible areas of
∗ Corresponding author. Tel.: +91 44 27480106; fax: +91 44 27480104. level fluctuation, thermal stratifications and thermal striping. The
E-mail address: pcp@igcar.gov.in (P. Chellapandi). concerns of thermal fluctuations are addressed comprehensively

0029-5493/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.nucengdes.2009.08.014
P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765 2755

Nomenclature

 thermal stress range


 el elastic stress range
 elpl elastoplastic stress range
m mean primary stress
k governing stress for creep damage
d sigma-d stress
Deff equivalent creep-fatigue damage
nL applied load cycles
 Poisson’s ratio
∈c creep strain range
∈elpl elastoplastic strain range
∈el elastic strain range
∈t total strain range
Tmax peak through-wall temperature gradient
Tp permissible range of temperature gradient
Ts temperature fluctuations on the metal wall surface
Kf stress intensity factor range
Ks symmetrisation coefficient
Np permissible number of cycles
Jr creep strength reduction factor
Jf fatigue strength reduction factor
tr time to rupture
Sr minimum stress to rupture
V fatigue damage
W creep damage Fig. 2. Thermal striping, thermal stratification and level fluctuations in FBR.
d distance from crack tip
D/h slenderness ratio of shell
There are a few reported failures, in the form of extensive
nH number of high frequency thermal cycles
cracking due to thermal striping in the operating reactors at sec-
T specified plant life
ondary sodium pump vessel in Phenix (Gelineau and Sperandio,
LF load factor
1994), Tee junction of an auxiliary pipe of the secondary circuit
a crack depth
in SPX (Gelineau and Sperandio, 1994), control rod guide tube
f thermal striping frequency
in PFR (Bettes et al., 1994), and primary cold trap in BN 600
ı uniform wall thickness
(Sobolev and Kuzavkov, 1994). In view of these reported failures,
ω frequency of oscillations
the structural mechanics specialists have been investigating ther-
˛ temperature attenuation factor on the surface
mal striping phenomenon both experimentally and theoretically.
Investigations focus both on thermal hydraulics and structural
mechanics aspects. Thermal hydraulics studies cover basic under-
by Ohshime et al. (1994) and Gelineau and Sperandio (1994). The standing of the phenomenon (Czeslaw and Trass, 1991; Tokuhiro
range of frequency of oscillations under thermal striping is found and Kimura, 1999), establishing experimental simulation princi-
to be 0.1–10 Hz as seen from spectra (Fig. 3), obtained from a rig at ples and assessments based on tests with water and air (Tenchine
AEA technology (Jones and Lewis, 1994). and Moro, private communication), numerical simulation tech-

Fig. 1. Thermal striping phenomenon in the vicinity of core cover plate in FBR. Fig. 3. Power spectral density of a typical thermal striping (PFR) .
2756 P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765

niques such as Large Eddy Simulation (LES) (Muramatsu, 1994a, ble to most of the current and evolving SFR designs. Accordingly,
1998a, 1998b; Hu and Kazimi, 2006; Menant and Villand, 1994) the structural material considered is austenitic stainless steel type
and attenuation characteristics of thermal striping on the metal SS 316 LN, which is the general choice of current and future SFR
wall (Muramatsu, 1994b; Wakamatsu et al., 1995). By attenuation, designs. From thermal striping considerations, alternate material
it is meant that the temperature fluctuations experienced on the is considered for above core structure parts to allow higher ther-
metal wall are expected to be lower depending upon the frequency mal striping limits. However, in this case, thermal striping would
of oscillation and heat transfer coefficient. The outcome of thermal not be of concern. Further, the hot pool is at isothermal temperature
hydraulics investigations is the random temperature fluctuations at 823 K during normal operating condition. This temperature has
on the metal wall, which involve many system parameters and been selected for the new designs such as EFR (France), JSFR (Japan)
hence need to be assessed case by case. However, a few generic and BN-series of reactors (Russia). In view of high heat transfer coef-
conclusions can be derived with reference to frequency contents, ficient of sodium, the structural wall surface where the damage is
attenuation characteristics, etc. These informations are useful to maximum, sees the temperature of hot pool sodium (823 K). Since,
derive thermal striping limits. Structural mechanics investigations the maximum temperature should be considered for the creep-
are reported extensively in literature (Clayton and Irvine, 1987; fatigue damage assessment as per design codes such as ASME Sec III
Miller, 1980; Tanaka and Toyoda, 1996; Gruter and Huget, 1982; NH (ASME, 2007) and RCC-MR (RCC-MR, 2007), the temperature of
Durbay and Acker, 1994; Jones, 1997; Kasahara, 1999; Kasahara, 823 K is considered for the damage assessment. For computing the
2001; Muralidharan et al., 2001). The main objective of the inves- through-wall temperature distribution, heat transfer coefficient of
tigation is to derive high cycle design fatigue curves and thereby sodium in the vicinity of wall surface is required. Since the sodium
establishing thermal striping limits, based on numerical analysis has excellent heat transfer properties, its heat transfer coefficient
as well as from the results derived from dedicated facilities such as is high, generally varying from 10,000 to 40,000 W/m2 -K. For the
SOMITE & SUPERSOMITE in UK (Picker, 1996) and FAENA & SPALSH computation of structural wall temperatures and stresses, higher
in CEA France (Lejeail, 1994; Poette, 1998). Reflecting the available heat transfer coefficient value yields conservative results and hence
data and numerical results, design rules have been reported for the 40,000 W/m2 -K is used. As regards the wall thickness, the maxi-
structures subjected to thermal striping by Buckthorpe (1997) and mum thickness of hot pool components is 30 mm, which is assumed
Gelineau et al. (1999). In spite of such extensive data and publica- in the analysis. It will be shown in the paper that beyond this thick-
tions available, robust design rules to specify the thermal striping ness, it does not have any significant effect. Apart from the above
limits reflecting the latest creep-fatigue design criteria and further mentioned material, temperature and wall thickness, another input
confirmation of such limits with thermal hydraulics analysis are data required is the accumulated creep and fatigue damage at the
not reported comprehensively any where to the best of authors’ specified locations, which is a parameter in the present study.
knowledge. This is the scope of the present paper. Accordingly, allowable thermal striping limits are recommended
The paper presents a general methodology based on fracture as the function of accumulated creep-fatigue damage.
mechanics concepts for establishing allowable T between hot and For deriving thermal striping limits, it is assumed that the struc-
cold sodium streams in the vicinity of metal wall surface and the tural wall has accumulated an effective creep-fatigue damage,
recommendations of thermal striping limits at a few potential loca- termed as ‘Deff ’, due to the major load cycles caused by nor-
tions in the reactor assembly. Gelineau et al. (1999) Descatha et mal and thermal transient conditions. Further, it is also assumed
al. (1980), Saanouni and Bathias (1982), Roche (1987), Moulin et that the structural wall has an initial part-through crack size of
al. (1987, 1995), and Autrusson et al. (1988) address the fracture 0.1 mm which might not have been detected by the existing NDE
mechanics approaches relevant to structural integrity assessment techniques, that can grow due to the imposed major load cycles,
under thermal striping. depending upon the value of Deff . As per the applicable design codes
The paper has four sections. Section 1 is the introduction to the for the structural design of SFR components (ASME, 2007; RCC-
subject matter of this paper. Section 2 presents the overall method- MR, 2007), the maximum permissible value for Deff is equal to 1.
ology adopted for (1) idealization of geometry, (2) derivation of Thus, the structure can have a value for Deff , varying from 0 to 1. As
equivalent crack depth corresponding to Deff , (3) idealization of per the approach, proposed by Picker (1996), when the Deff value
temperature fluctuations through power spectral density (PSD), (4) varies from 0 to 1, the crack grows from 0.1 mm to 0.5 mm. This
main steps involved in the derivation of thermal striping limits and implies that we are permitting/restricting a maximum crack size of
(5) computation of temperature, stress and stress intensity distri- 0.5 mm in the design stage itself, which can be detected confidently
butions. Section 3 deals with the application of methodology to during periodic in-service inspection stage, employing appropri-
derive thermal striping limits for 500 MWe prototype fast breeder ate NDE techniques. The plate can have any crack size (a) within
reactor (PFBR), as a function of accumulated creep-fatigue dam- the range of 0.1–0.5 mm, depending upon the accumulated dam-
age and assessment of these limits with the operating experiences. age (Deff ) at that location. With the presence of thermal fluctuations
Section 4 presents the summary of thermal hydraulics analysis car- imposed by thermal striping, it is restricted that the crack should
ried out for PFBR to demonstrate that the recommended limits not further grow due to thermal striping. This condition is ensured
are respected. The highlights of creep-fatigue damage and fracture by satisfying the fatigue damage design criteria recommended by
assessment procedures of RCC-MR (edition 2007) (RCC-MR, 2007) A16 through ‘␴-d approach’ (RCC-MR A16, 2007). The extract of
and A16 (RCC-MR A16, 2007), which form the basis for establish- the design rules are presented in the Appendix of this paper. The
ing thermal striping limits are presented in Appendix. In the main step by step procedure of establishing thermal striping limits as
part, the expression for the equivalent damage (Deff ) to represent per the approach described above is illustrated in the following
the creep-fatigue damage interaction is only presented. sections.

2.1. Idealization of geometry


2. Methodology to establish thermal striping limits
SFR geometries are generally cylindrical shells with L-seam and
Thermal striping limits are established by specifying acceptable C-seam welds. These are idealized as plates. This assumption also is
T in the sodium jets in the vicinity of structural wall surface, valid for the cylindrical shell with high diameter by thickness ratio
delinking the thermal hydraulics of the sodium system. Associated (D/h). For a smaller D/h ratio, the plate assumption yields conser-
parameters are chosen to make the recommended limits applica- vative results (Jones, 1997). An idealized temporal random surface
P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765 2757

Fig. 5. Equivalent crack depth for the effective creep-fatigue damage.

Integration of above equation with the conditions: a0 for t = 0


Fig. 4. Creep-fatigue interaction diagram for SS 316 LN. and af for t = tf , yields:
 t
1/0.19
a = a0.19 + (a0.19
f
− a0.19
0 ) (4)
temperature history is derived from the operating experiences, tf
which is presented in Section 2.3.
Since t/tf = W,
The number of high frequency thermal cycles (nH ) is determined
from the specified plant life (T), load factor (LF) and frequency (f): 1/0.19
a = [a0.19 + (a0.19 − a0.19 )W ] (5)
nH = (T × LF × f) × 365 × 24 × 3600 = 3.15 × 107 (T × LF × f). For typi- 0 f 0

cal values of current SFRs with T = 45 years, LF = 75% and f = 1 Hz,


Under the assumption of only fatigue damage, the crack growth
nH works out be ∼1.0 × 109 . However, T = 60 years and LF = 85% are
is due to fatigue crack growth, governed by the following growth
targeted for future SFRs.
law recommended in A16.
n
2.2. Effective creep-fatigue damage (Deff ) da/dn = C2 (K) (6)

Applying the recommended value for n = 3.33 for 1S material


As per RCC-MR creep-fatigue interaction diagram, which is bi- √
and K ∝ a,
linear in nature, the creep damage (W) and fatigue damage (V)
should lie within the safe operating domain (Fig. 4). A single param- da/dn = C2 a1.65 (7)
eter called ‘effective creep-fatigue damage (Deff )’ is derived from
the individual values of W and V such that if a particular com- Integration of above equation with the conditions: a0 for n = 0
bination of W and V just lies on the bi-linear curve, then Deff and af for n = nf , yields:
is equal to 1. Accordingly the expression for Deff is defined as  n
−1/0.665
follows: a = a−0.665 + (a−0.665 − a−0.665 ) (8)
0 f 0 nf
Deff = (3V + 7W )/3 for W ≤ V
(1) Since n/nf = V,
= (3W + 7V )/3 for W ≥ V
−1/0.665
It may be verified that, for the combinations (W = 0 and V = 1), a = [a−0.665
0
+ (a−0.665
f
− a−0.665
0
)V ] (9)
(W = 0.3 and V = 0.3) and (W = 1 and V = 0), the value for Deff is equal
to 1. As per UK fatigue design procedure recommended by Picker
(1996), the crack depth a resulting due to the applied creep-fatigue
damage Ve , is expressed as follows:
2.3. Equivalent crack length for accumulated creep-fatigue
damage a = a0 e1.61Deff (mm) (10)

Under the assumption of only creep damage, the crack growth It may be checked that a = 0.1 mm for Deff = 0 and a = 0.5 mm for
from 0.1 mm to 0.5 mm is due to creep crack growth. A16 recom- Deff = 1.
mends following growth law: Using the values of a0 = 0.1 mm, af = 0.5 mm, nf = 1 × 109 and
tf = 3 × 105 h, the curves obeying Eqs. (5)–(9) and (10) are plotted
da/dt = C ∗ [J]k (2) in Fig. 5. Assessment of creep-fatigue damages for various reactor
assembly components of PFBR indicates that the creep damage is
Applying the recommended value for k = 0.81 for 1S material at dominant. The trend curve recommended by Picker (1996), reflects
820 K and J ∝ a, this situation and hence the Eq. (10) is used for deriving the equiv-
alent crack size, ‘a’ for the given Deff value depending upon the
da/dt = C1 a0.81 (3) location.
2758 P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765

Fig. 6. Water test facility to simulate thermal striping (IGCAR).

2.4. Idealized power spectral density for thermal striping ing a part-through-wall crack of size a, and subjected to random
temperature history surface temperature history with the peak value Tmax , whose
PSD has a constant value up to 10 Hz and zero beyond 10 Hz.
The temperature fluctuations either in the fluid or in the metal The plate has accumulated fatigue and creep damage values of
wall surface are random in nature. Hence, temperature histories are V and W respectively due to normal and design basis thermal
defined as power spectral density (PSD), from which RMS values transients. With these idealizations, the temperature, stress com-
can be derived by integration. Thermal striping phenomenon has ponents and stress intensity values are derived analytically with
been simulated in water tests to get an idea of frequency contents the objective of determining permissible temperature range (Tp )
through a dedicated test setup, developed at IGCAR (Fig. 6). The
setup simulates the situations of thermal hydraulics in the vicinity
of core cover plate, placed on the control plug just above core. Tem-
peratures and flow rates of hot and cold water are varied over wide
ranges. A maximum T of 90 K is possible in the setup. A few typi-
cal power spectral density (PSD) diagrams generated from test data
are shown in Fig. 7. These are broadly similar to the idealized PSD
shape (‘Profile A’) as shown in Fig. 8, which was considered for the
analysis by Jones and Lewis (1994). Further, the Fig. 8 also shows a
constant spectrum (white noise) up to a frequency of 10 Hz (‘Profile
B’), considered for the present analysis in view of its computational
simplicity. It has been confirmed from the analysis by Jones and
Lewis, who used the ‘Profile A’ and ‘Profile B’ and have shown that
results are similar for small crack size up to about 0.5 mm (Jones
and Lewis, 1994).

2.5. Main steps to derive thermal striping limits

Based on the assumptions described in Sections 2.1–2.3, the


problem is idealized as a plate of uniform wall thickness h, hav- Fig. 7. Typical PSD generated from experimental data (IGCAR).
P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765 2759

2f. f is the frequency expressed in cycles/s (Hz). The back face is


adiabatic. The temperature distribution across the plate thickness
h is expressed as

T (x, t) = Tf A(ω, x) (13)

where temperature response function: A(ω,x) = (P(ω,x) +


iQ(ω,x))/(R + iS).
The functions P and Q are obtained so that the Eq. (13) sat-
isfies the governing differential Eq. (14) and adiabatic boundary
condition defined by Eq. (15):
∂2 T 1 ∂T
= (14)
∂x2 k ∂t
where k is thermal diffusivity of structural material.
Fig. 8. Idealized PSD considered for establishing thermal striping limits. Adiabatic wall condition:
∂T
=0 at x=h (15)
as a function of Deff . The following steps are followed to derive such ∂x
relation. The following analytical expressions for P and Q satisfy the above
Eqs. (14) and (15)
• Knowing V and W, Deff is computed using Eq. (1).
• For the known Deff , equivalent crack depth a is computed using P(ω, x) = cos  (h − x) cosh (h − x)
(16)
Q (ω, x) = sin  (h − x) sinh (h − x)
Eq. (10).
• The temperature distribution within the metal wall of the plate, 
which is subjected to random temperature history defined in Sec- w
=
tion 2.3 is determined using frequency response function method 2k
(Jones and Lewis, 1994). The analytical expressions for R and S are obtained by satisfying
• The thermal stress range ( f ) that is associated with the tem- the following convective heat transfer boundary condition (21) on
perature fluctuation (Ts ) and subsequently the range of stress the surface facing the fluid temperature variations.
intensity factor (Kf ) corresponding to  f are obtained as a
∂T
function of thermal striping frequency f. Ts is a parameter here. −K = H (Tf − Ts ) (17)
• Krms and subsequently Kmax are computed using idealized ∂x
PSD.  dP dQ
 
• Knowing Kmax , the maximum stress range is computed using −KTfm eiωt (0) + i (0) /(R + iS)
dx dx
Crager’s formula recommended in A16 (RCC-MR A16, 2007) at  
the distance of 50 ␮m ahead of crack tip. P(0) + iQ (0)
= Heiωt Tfm − Tfm (18)
• ∈t is computed as per RCC-MR procedure presented in Section R + iS
2.1.
• As per ‘␴-d procedure’, ∈t /1.5 is compared with the allowable Simplification of Eq. (18) yields:
value (∈allowable ). K
 dP dQ
 
P(0) + iQ (0)
• ∈allowable for the ASS type 316 LN (1S as per RCC-MR designa- − (0) + i (0) /(R + iS) = 1 − (19)
H dx dx R + iS
tion) at the mean temperature of 820 K (typical situations of an  K dP
 
K dQ

SFR), is equal to 0.104% as read from the Fig. 5 corresponding to P(0) − (0) + i Q (0) − (0) = R + iS (20)
the base metal. H dx H dx
• The T value for which this strain limit is respected is termed as From Eq. (20), the following expressions for R and S are obtained.
Tp . ⎫
• By varying Deff , corresponding Tp is computed by repeating the R = P(0) −
K dP
(0) ⎪

above steps. H dx
(21)

(0) ⎭
K dQ
S = Q (0) −
2.6. Transient temperature distribution H dx
From Eq. (16):
The solution for the transient temperature distribution in the
plate, surface of which is subjected to sinusoidal temperature vari- P(0) = cos h cosh h and Q (0) = sin h sinh h (22)
ation, is given by Carslaw and Jaeger (1959) and also by Jones and
Differentiation of Eq. (16) and putting x = 0,
Lewis (1994). In the present case, the convective boundary hav- ⎫
(0) = − [cos h sinh h − sin h cosh h] ⎪
ing heat transfer coefficient, H, is assumed at the structural wall dP

surface, for which the solution is given below: dx
(23)
Temperature fluctuations in the fluid (Tf ) as well as on the ⎪
(0) = − [sin h cosh h + cos h sinh h] ⎭
dQ
surface (Ts ) are assumed to vary in phase as per the following Eq. dx
(11):
Substituting Eq. (23) into Eq. (21),
Tf = Tfm eiωt (11)
K
R = cos h cosh h +  [cos h sinh h − sin h cosh h]
Ts = Tsm eiωt (12) H
(24)
K
where Tfm and Tsm are the maximum temperature difference S = sin h sinh h +  [sin h cosh h + cos h sinh h]
(range) in the fluid and at the metal surface respectively. ω is the H
frequency of oscillations expressed in radians/s, which is equal to If H → ∞, R = cos h cosh h and S = sin h sinh h (25)
2760 P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765

Fig. 9. Attenuation of thermal striping temperature on the metal wall surface.


Fig. 10. Decay of stress across the structural wall thickness.
The expressions are matching with solutions given by Jones and
Lewis (1994) for case H → ∞.
once cracks are initiated by high cycle fatigue cycles (accumulated
The attenuation of peak temperature on the metal wall surface
quickly within a few years), other major load cycles causing creep-
is given by
fatigue damage can cause propagation and ultimate rupture.

Tsm (P 2 + Q 2 )
=˛= (26) 2.8. Determination of KI
Tfm (R2 + S 2 )

For a typical practical case: plate thickness (h) = 30 × 10−3 m, For the mode-I field, the crack tip stress intensity factor cor-
heat transfer coefficient of sodium (H) = 40,000 W/m2 -K and ther- responding to a part-through crack a is determined using the
mal diffusivity of austenitic stainless steel (k) = 4.76 × 10−6 m2 /s at Bueckner weight function as (Bueckner, 1973):
820 K, the attenuation factor (˛) is plotted as function of frequency a
(f in Hz) using the Eq. (26) in Fig. 9. K1 (x, t) = (x, t)M(x) dx (30)
0
2.7. Transient thermal stress distribution
where the weight function M(x) for an edge cracked plate is given
by
It is conservatively assumed that the plate is fully constrained.

The axial stress range induced at any distance x from surface is M(x) = 2a{(1 − x/a)−1/2 + m1 (1 − x/a)1/2 + m2 (1 − x/a)3/2 } (31)
expressed as
where m1 and m2 are the polynomials independent of x:
(x, t) = [E˛ Tf /(1 − )]A(ω, x) (27)
2 6
m1 = 0.6147 + 17.1844(a/h) + 8.7822(a/h)
where E, ˛ and  are the Young’s modulus, coefficient of thermal 2 6
m2 = 0.2502 + 3.2889(a/h) + 70.0444(a/h)
expansion and Poission’s ratio respectively for the structural mate-
rial. The peak-to-peak stress variation at any location measured Using Eq. (27) for (x,t) and Eq. (11) Tf for 0 ≤ a ≤ h/2
from the front face is expressed as a function of ω in the following a
equation: K1 (a, t) = −[E˛ Tfm /(1 − v)]eiωt A(x, ω) M(x) dx
0 (32)
(x, ω) = [E˛ Tf /(1 − )] (P 2 + Q 2 )/ (R2 + S 2 ) (28)
K1 (a, ω) = −[E˛ Tf /(1 − v)] (I12 + I22 )
From Eq. (6), the stress range per unit Tf is expressed as
a a
where I1 = 0 P(x)M(x)dx and I2 = 0 Q (x)M(x)dx.
(x, ω) = [E˛/(1 − )] (P 2 + Q 2 )/ (R2 + S 2 ) (29)
The stress intensity factor (SIF) per unit Tf is expressed as
The stress decay has been plotted as a function of frequency 
of striping in Fig. 10, using Eq. (26) across the plate thickness K1 (a, ω) = −[E˛/(1 − v)] (I12 + I22 ) (33)
(ı) = 30 × 10−3 m. Other parameters used are: heat transfer coef-
ficient (h) → ∝ W/m2 -K, thermal diffusivity (k) = 4.76 × 10−6 m2 /s, In Fig. 11, KI is shown as a function of crack size a (0.1–5 mm)
Young’s modulus (E) = 1.63 × 105 MPa,  = 0.3 and ˛ = 20.0 × 10−6 /K for 3 frequencies, 0.0625 Hz, 1.0 Hz and 6.25 Hz for the plate of
and Tfm of 20 K. From Fig. 10, it is seen that, in the case of constant thickness of 10 mm. The SIF increases monotonically for frequen-
temperature on the wall surface (frequency = 0), the stress is con- cies less than 1 Hz and for frequencies higher than 1 Hz, it increases
stant across the thickness. When the frequency increases from 0 to only up to a certain crack size. After reaching a maximum value, SIF
1 Hz, the stress decays rapidly across the thickness, which reaches decreases. This sort of drooping behaviour indicates that there is
nearly saturation at the frequency of about 1 Hz. Hence at higher a possibility of crack arrest under high frequency thermal fluctua-
frequencies, the stresses are concentrated in the vicinity of surface tions. More pronounced drooping characteristics can be seen in the
without significant penetration, implying that thermal striping can case of cylinders which are presented by Jones (1997). This ensures
only initiate cracks and does not have potential to cause growth. that the assumption of plate geometry yields conservative thermal
This physical understanding can be derived from Fig. 10. However, striping limits, which are preferred in the design stage.
P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765 2761

Table 1
Thermal striping experiences.

Reactor/facility Geometry T (K) No. cycles

Operating reactors
Phenix Expansion tank 170 5.50 × 106
Phenix Mixing tee 90 3.24 × 108
SPX Mixing tee 220 1.44 × 104
BN 600 Cold trap 160 2.20 × 107

International test facilities


FAENA Cylindrical tube 1 378 1.60 × 104
Fig. 11. Through-wall variation of stress intensity factors. Cylindrical tube 2 192 1.05 × 106

SUPERSOMITE Cylindrical tube 1 210 3.60 × 106


Cylindrical tube 2 180 3.60 × 107
3. Establishing thermal striping limits for SFR

The methodology dealt in previous section is used for deriving 3.1. Operating experiences
permissible temperature fluctuations (peak-to-peak temperature
range) for PFBR. The following parameters are considered. As mentioned in Section 1, two sodium leaks were detected in
Material SS 316 LN the secondary sodium circuit of 250 MWe Phenix reactor during the
Plate thickness h = 5–30 mm
inspection campaign after operation for about 90,000 h. The mate-
Cutoff frequency f = 10 Hz
Plant life (60 years with 85% LF) T = 4.47 × 105 h rial of construction is AISI 304. The first one was found at the C-seam
Effective damage Deff = 0–1.0 weld on main pipeline near the mixing Tee where the tempera-
Maximum metal surface temperature 820 K ture between the hot sodium (703 K in the branch line) and cold
The material properties at 820 K: sodium (613 K in the main line) is 90 K. The second one is again
Young’s modulus E = 1.49 × 105 MPa
on the weld line in the expansion tank in the vicinity of sodium
Density
= 7739 kg/m3
Specific heat Cp = 582 J/kg-K discharge area. The temperature of the hot leg is 823 K and the
Thermal conductivity  = 21.54 W/m-K temperature of sodium in the tank is 623 K. Taking into account the
Thermal expansion coefficient ˛ = 20.4 × 10−6 /K proximity of mixing zone from the wall, the T between the hot
and cold sodium is 170 K at the failure location. The low frequency
Weld material properties are used for the damages assessment. displayed is 0.017 Hz and even less. In SPX reactor, sodium leak was
Permissible thermal striping limits (Tp ) are derived as a func- observed on a circumferential weld downstream of a Tee junction
tion for two cases: (1) there is no thermal attenuation on the surface in an auxiliary pipe of the secondary sodium circuit. The crack was
wall temperature, which has been implemented by setting h = ∝ and due to thermal striping T of 220 K, caused by partial opening of
(2) there is thermal attenuation depending upon the frequency for dry-out valve. The duration of the unexpected flow is estimated as
h = 40,000 W/m2 -K. The results are shown in Fig. 12. It is noted from 4 h. In BN 600 reactor, after operation for about 10 years with 72%
this figure that the Tp reduces significantly with the increase of load factor, primary sodium leaks took place in cold trap. Thermal
creep-fatigue damage: 39–26 K under the assumption of no atten- striping was identified as a possible reason. The associated T is
uation and 75–46 K with attenuation. This also indicates that the estimated as 160 K (possible if cold sodium leak exists).
effect of attenuation due to high frequency components of temper- In the international test facilities, viz. SUPERSOMITE (AEA, Ris-
ature spectrum is very significant. This aspect has also been brought ley) and FAENA (CEA-Cadarache), thermal striping cracks were
out in Kasahara (1999). simulated on AISI 316 steels. In SUPERSOMITE tests on cylinder
It is worth mentioning that the creep-fatigue damage estimation with circumferential TIG welds, typically, T of 210 K and 180 K
approach and material data for establishing the limits have been
completely adopted from RCC-MR, which in turn accounts for the
various uncertainties including material data systematically. Thus
the present approach is considered as robust.

Fig. 12. Permissible thermal striping limits for the FBR structures, made of SS 316
LN. Fig. 13. Thermal striping limits and failure experiences.
2762 P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765

Fig. 14. Temperature fluctuations in thermal striping zones in PFBR sodium pools.

caused crack initiation and some limited crack propagation after cal Tee in hot secondary sodium pipeline) where the creep-fatigue
about 1000 h and 10,000 h respectively. The FAENA tests indicated damage can be as high as 0.9.
that the T of 192 K (equivalent strain range = 0.382%) and 378 K Pool thermal hydraulics analysis of reactor assembly of PFBR,
(equivalent strain range = 0.75%) caused cracks of about 50 ␮m size has been carried out based on direct numerical simulation (DNS)
after 1 × 106 and 1.6 × 104 cycles respectively. These data are sum- formulation (Velusamy et al., 2005). Fig. 14 shows the summary
marised in the Table 1 with the aim of assessing the reccommended of thermal hydraulics analysis, which indicates that the temper-
thermal striping limits. Wherever the frequency data is not avail- ature fluctuations in the reactor assembly components, due to
able, 1 Hz is assumed and for the Phenix reactor, a load factor of thermal striping are well below the respective permissible values
50% is assumed to determine the number of cycles (N). (Tp ). In order to have realistic assessment of damage, the com-
From the data presented in Table 1, T and N are related by bined approach, i.e. computational fluid dynamics (CFD) and stress
the relation: T = CN˛ . For determining C and ˛, 2 sets of data are and strain analysis, is preferred solution. With such approach, it
required. Accordingly, 3 pairs viz. Phenix, FAENA and SUPERSOMITE is possible to demonstrate the availability of higher safety mar-
test data are used to extrapolate for the 1 × 109 cycles (60 years gin above the inherent factor of safety specified in design codes.
design life with 85% load factor and 1 Hz frequency). The corre- This approach, of course involves many geometrical and process
sponding extrapolated values are: 68 K, 63 K and 144 K based on parameters that are system dependent. So in the present case, the
Phenix, FAENA, SUPERSOMITE data respectively. It is worth men- CFD analysis is performed, after establishing limits to demonstrate
tioning that, at all the failure locations in the operating reactors, that the PFBR plant meets the limits.
creep-fatigue damages due to major cycles were present. Hence, the
values are on lower side (63 K). For SUPERSOMITE, failure has been 4. Further works in progress
detected after certain crack propagation and hence it yields a higher
value (144 K). Extrapolation using SPX and BN-600 data could not The assessment of conservatism in the present analysis with
be made because the required minimum two sets of data are not actual 3D component analysis including the effects of the major
available. SPX data is not explainable due to limited information load cycles, understanding the thermal striping failure mechanisms
reported. Despite the above, this exercise provides good confidence in the materials with a possible variations in the properties, valida-
on the thermal striping limits arrived in Section 3. This is under- tion of the structural mechanics analysis carried out for the damage
standable from Fig. 13 where the extrapolated T based on Phenix, assessment and thermal hydraulics analysis carried out for deter-
FAENA, SUPERSOMITE data are superimposed. With assumption of mining the temperatures in the potential places of thermal striping
LCF damage fraction in PHENIX and FAUNA, the permissible limits are the activities being pursued at the centre.
recommended reflect the operating experiences. In SUPERSOMITE
tests, there is no LFC damage and hence, the T at which failure 5. Conclusions
occurs is higher (144 K). Hence, they can be used for the future
FBRs, which are designed for long design life (60 years) and with The effect of frequency-dependent thermal attenuation on the
high capacity factor (85%). structural wall surface is quantified and found to be significant. The
As per the recommendation, at the location such as core cover limits are derived at various potential locations in SFRs which are
plate where the accumulated creep damage is insignificant and also subjected to creep-fatigue damage due to major cycles caused
fatigue damage is small (<0.1), Tp can be 70 K. At the control plug by startup, shutdown, power failures and pump trips, etc. The max-
lower portion where the Deff is <0.3, Tp can be 62 K and at the main imum range of temperature fluctuations can be as high as 70 K
vessel near IHX outlet, where Deff is <0.5, Tp can be 55 K. Further, where the accumulated creep-fatigue damage is insignificant (<0.1)
minimum Tp of 45 K can be acceptable at any location (e.g. typi- and 45 K is acceptable where the creep-fatigue is very significant
P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765 2763

(0.9). These limits are found to be consistent with the broad limits
extrapolated from the failure experiences of international SFRs and
sodium facilities. Pool hydraulic computations carried out to iden-
tify and quantify the thermal striping zones have confirmed that
the proposed limits can be respected with good margins (>25 K) for
future SFRs with the targeted design life of 60 years and load factor
of 85%.

Acknowledgements

The colleagues in Structural Mechanics Laboratory at IGCAR who


helped to carry out experimental work, Dr. K. Velusamy who pro-
vided thermal hydraulics results and Miss M.D. Ambuja for typing
the manuscript, are sincerely acknowledged for their contributions.

Appendix A. Creep-fatigue damage estimation procedure of


RCC-MR
Fig. A2. Creep rupture curves for base metal and weld for SS 316 LN at 823 K.
A.1. Fatigue damage of base metal

where Ks is the symmetrisation coefficient and  m is the mean pri-


Elastically computed equivalent stress range ( el ) and number
mary stress sustained during the hold period which is computed
of applied load cycles (nL ) are the input data. French design code
through elastic analysis.  k is used for computing ∈c and creep
RCC-MR provides rules for estimating fatigue damage from  el
damage in a load cycle.
and nL . The elastic strain range is computed as
Total strain range, ∈t = ∈1 + ∈2 + ∈3
 ∈ 1 = [2(1 + ) el ]/(3E) (A.1)
+∈4 + ∈c (A.4)
From ∈1, the plastic strain accumulation due to variation of
primary stresses (∈2 ), strain concentration (∈3 ), ∈4 the strain
enhancement due to the difference in the value of Poisson’s ratio Corresponding to ∈t , the permissible number of load cycles
(0.5 for plastic zone and 0.3 for elastic zone) and ∈c , creep strain Np is obtained from the design fatigue curve provided in the RCC-
developed during the hold time that is associated with the stabi- MR. Knowing the design load cycle nL and permissible cycle Np the
lized load cycle can be determined using the procedure of RCC-MR. fatigue damage (V) is computed as: V = nL /Np . If there are m number
For the plastic strain accumulation at the strain concentrated region of events, the above procedure is repeated to compute Vi for each
∈3 , Neuber’s rule is applied as follows: event knowing the ( el )i , (nL )i , the strain range and permissible
number of cycles that are associated with each cycle i.  Then the
elpl  ∈ elpl = invariant and hence = el el (A.2) cumulative fatigue damage V is summation of Vi , i.e., V = Vi .

The graphical construction for estimating the ∈elpl using the A.2. Creep damage of base metal
above Eq. (A.2) is illustrated in Fig. A1. Here ∈el = ∈1 + ∈2 and
∈elpl = ∈1 + ∈2 + ∈3 . The ∈c is the creep strain accumulated RCC-MR provides creep rupture curves which give the mini-
over the hold period in the given load cycle corresponding to  k mum time to rupture for the given stress and temperature. For
which is given by welds, creep strength reduction factor called Jr is applied, which
is provided as a function of temperature and life. Introducing a fac-
k = m + Ks s elpl (A.3)
tor of safety on sustained stress  k , the effective stress is defined
as  e =  k /0.9. The minimum time to rupture (tr ) is read from the
creep rupture curve corresponding to  e and appropriate temper-
ature. The creep damage of ith load cycle, Wi is equal to (ti /tri ). The
cumulative creep damage W is given by: W = Wi .

A.3. Creep and fatigue damage for welds

For the matching welds, the creep rupture as well as fatigue


curves corresponding to its base metal are altered using the weld
strength reduction factors as indicated below:
The minimum stress (Sr ), the ordinate of creep rupture curve
is multiplied by a factor called Jr . Since Jr is less than 1, the stress
rupture curve for weld is lower than the curve corresponding to
base metal. In the following Table A1, the Jr values provided in RCC-
MR: Appendix A9 as a function of time are reproduced for stainless
steel SS 316 LN (1S material) at 823 K. Typical Sr curve thus derived
for welds at 823 K are shown along with respective curve for base
metal for 1S material in Fig. A2.
The strain range (∈), the ordinate of design fatigue curve is
divided by a factor called Jf . The Jf value is equal to 1.25. Accord-
Fig. A1. Neuber’s rule for estimating elastoplastic stress–strains. ingly, the design fatigue curve for weld is lower than the curve
2764 P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765

Table A1
Jr values for ASS at 823 K recommended in RCC-MR.

Time 1 101 3 × 101 1 × 102 3 × 102 1 × 103 3 × 103 1 × 104 3 × 104 1 × 105 3 × 105
Jr 1.0 1.0 0.99 0.98 0.89 0.80 0.78 0.77 0.75 0.73 0.71

Descatha, Y.C., Devaux, J.C., Bernard, J.L., Pelissier-Tanon, A., 1980. A criterion for
analysing fatigue crack initiation in geometrical singularities. IMechE.
Durbay, B., Acker, D., 1994. Evaluation of thermal striping risks: limitation of crack
initiation and propagation, IWGFR/90. In: Specialists Meeting on “Correlation
between material properties and thermohydraulics conditions in LMFBRs, Aix-
en-Provence, France, November 22–24.
Gelineau, O., et al. Review of predictive methods applied to thermal striping problem
and recommendations. SMiRT 15, vol. IV, Div. F, Seoul, Korea, 15–20 August 1999.
Gelineau, O., Sperandio, M., 1994. Thermal fluctuation problems encountered in
LMFBRs, IAEA-IWGFR/90. In: Specialistic Meeting on “Correlation Between
Material Properties and Thermohydraulics Conditions in LMFBRs”, Aix-en-
Provence, France, November 22–24.
Gruter, L., Huget, W., 1982. Fatigue crack behaviour under thermal stresses. Journal
of Pressure Vessel Piping 110, 335–359.
Hu, L.-W., Kazimi, M.S., 2006. LES benchmark study of high cycle temperature fluc-
tuations caused by thermal striping in a mixing tee. Journal of Heat and Fluid
Flow 27, 54–64.
Jones, I.S., 1997. The frequency response model of thermal striping for cylindrical
geometries. Fatigue & Fracture of Engineering Materials & Structures 20 (6),
871–882.
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).
Kasahara, N. The structural response diagram approach for evaluation of thermal
striping phenomena. SMiRT 15, vol. IV, Div. F, Seoul, Korea, 15–20 August 1999.
Kasahara, N., 2001. Frequency response function of structures to fluid tempera-
Fig. A3. Design fatigue curves for base metal and weld for SS 316 LN at 823 K.
ture fluctuations under arbitrary constraint conditions. In: Fourth International
Congress on Thermal Stesses, Osaka, Japan.
Lejeail, Y., 1994. The working out of a design rule in case of structures submitted
corresponding to base metal. Typical design fatigue curve for welds to thermal striping. In: IAEA Specialist Meeting on “Correlation between mate-
rial properties and Thermohydraulics Conditions in Liquid Metal-cooled Fast
at 823 K are shown along with the respective curve for base metal
Reactors (LMFRs)”, Aix-en-Provence, France, November 22–24.
of 1S material in Fig. A3. Menant, B., Villand, M., 1994. Thermal fluctuations induced in conducting wall by
mixing sodium jets: an application of TRIO-VF using large eddy simulation
modelling. In: IAEA Specialist Meeting on “Correlation Between Material Prop-
A.4. RCC-MR: Appendix A16 procedure erties and Thermohydraulics Conditions in Liquid Metal-cooled Fast Reactors
(LMFRs)”, Aix-en-Provence, France, November 22–24.
Miller, A.G., 1980. Crack propagation due to random thermal fluctuations: effect of
A16 procedure provides rules for assessing creep-fatigue temporal incoherence. International Journal of Pressure Vessel and Piping l8,
damage in structures having crack like defects, through ‘␴-d 105–130.
approach’. The background to ‘␴-d approach’ is found in Roche Moulin. D., Autrusson. B., Barrachin, B. Fatigue analysis methods of crack like defects:
strain range evaluation. SMiRT 9, vol. E, Lausanne, 1987.
(1987), Moulin et al. (1987, 1995), and Autrusson et al. (1988). As
Moulin, D., Drubay, B., Acker, D., 1995. Practical method based on stress evaluation
per this procedure, the creep and fatigue damage accumulated till (␴d criterion) to predict initiation of crack under creep and creep fatigue con-
the end of the design life is calculated at the characteristic distance ditions, pressure vessel fracture, fatigue and life management, ASME. Journal of
Pressure Vessel Technology 117, 335–340.
d from the crack tip and it should satisfy the creep-fatigue damage
Muralidharan, G., Sivakumar, S.M. Yashpal, K. and Chellapandi, P. Studies on thermal
criteria described in the preceding sections. The recommended striping limits 2nd Intl Conf on Theoretical, Applied, Compl and Exptl. Mechanics
value for the characteristic distance (d) is equal to 50 ␮m for ASS in (ICTACEM2001), Dec27–30, 2001, India, pp. 81–87.
A16. In the present context, the interest is computation of fatigue Muramatsu, T., 1994a. Development of thermohydraulics computer programs
for thermal striping phenomena. In: IAEA Specialist Meeting on “Correlation
damage, for which there is a need to compute net strain range Between Material Properties and Thermohydraulics Conditions in Liquid Metal-
∈t at the distance d from the crack tip, influenced by stress field cooled Fast Reactors (LMFRs)”, Aix-en-Provence, France, November 22–24.
around crack tip. Muramatsu, T., 1994b. Investigation on the reduction measures of coolant tem-
perature fluctuations based on numerical methods in LMFR designs. In:
IAEA Specialist Meeting on “Correlation Between Material Properties and
Thermohydraulics Conditions in Liquid Metal-Cooled Fast Reactors (LMFRs)”,
References Aix-en-Provence, France, November 22–24.
Muramatsu, T., 1998a. Numerical analysis of non-stationary thermally response
ASME Sec III Div 1. Rules for construction of nuclear power plant components, characteristics for a fluid-structure interaction system. In: ASME/JSME Pressure
Subsection NH: Class 1 Components, 2007. Vessels and Piping Division, vol. 363, pp. 149–156.
Autrusson, B., Moulin, D., Barrachin, B., 1988. Fatigue analysis of crack like defects: Muramatsu, T., 1998b. Computer code developments and their validations for the
experimental verification of practical rules to predict initiation. In: 6th ICPVT, thermal striping phenomena in PNC. In: Third Research Coordination Meeting
Beijing, China. (RCM) of the IAEA Coordinated Research Programme (CRP) on “Harmonization
Bettes, C., Judd, A.M., Lewis, W.W.J., 1994. Avoiding thermal striping damage: exper- and Validation of Fast Reactor Thermomechanical and Thermohydraulic Codes
imentally based design procedures for high cycle thermal fatigue, IWGFR/90. In: and Relations Using Experimental Data”, France.
Specialists Meeting on “Correlation Between Material Properties and Thermo- Ohshime, H., et al., 1994. Current status of studies on temperature fluctuation phe-
hydraulics Conditions in LMFBRs, Aix-en-Provence, France, November 22–24. nomena in LMFBRs, IAEA-IWGFR/90. In: Specialists meeting on “Correlation
Buckthorpe, D., Debaene, J.P., Rathjen, P. Design and construction committee rules between material properties and thermohydraulics conditioning in LMFBRs”,
for the EFR-current status and development. SMiRT 14, vol. F, Lyon, France, Aix-en-Provence, France, November 22-24.
17–22 August 1997. Picker, C. UK development of a strain based creep-fatigue assessment procedure
Bueckner, H.F., 1973. In: Sih, G.C. (Ed.), Methods of Analysis of Solutions of Crack for fast reactor design, IAEA-TECDOC-933. Creep-Fatigue Damage Rules for
Problems, vol. 1. 239 Nordhoff International, Leyden, p. 239. Advanced Fast Reactor Design. Technical Committee Meeting, Manchester, UK,
Carslaw, H.S., Jaeger, J.C., 1959. Conduction of Heat in Solids, second ed. Oxford at 11–13 June 1996.
the Clarendon Press. Poette, C. CEA R&D Experimental programmes related to thermomechanics and
Clayton, A.M., Irvine, M.N., 1987. Structural assessment techniques for thermal strip- fracture mechanics for fast reactors. Final IAEA Meeting on Thermal Striping
ing. Journal of Pressure Vessel Technology 109, 305–309. Benchmark, 1998.
Czeslaw, O.P., Trass, O., 1991. Visualization of a free and impinging round jet. In: RCC-MR: Appendix A16. Guide for Leak Before Break Analysis and Defect Assess-
Experimental Thermal and Fluid Science 1991. Elsevier Science Publishing Co. ment, AFCEN, 2007.
P. Chellapandi et al. / Nuclear Engineering and Design 239 (2009) 2754–2765 2765

RCC-MR subsection NB for class 1 components. Design and construction rules for Tenchine, D., Moro, J.-P. Mixing of Coaxial Jets: Comparison of Sodium and Air Exper-
mechanical components of FBR Nuclear Islands. AFCEN, Paris, France, 2007. iments. CEA, France. Paper received through private communication.
Roche, R. The use of elastic computation for analysing fatigue damage. SMiRT 9, vol. Tokuhiro, A., Kimura, N., 1999. An experimental investigation on thermal striping
E, Lausanne, 1987. mixing phenomena of a vertical non-buoyant jet with two adjacent buoyant jets
Saanouni, K., Bathias, C., 1982. Study of fatigue crack initiation in the vicinity of as measured by ultrasound Doppler velocimetry. Journal of Nuclear Engineering
notches. Engineering Fracture Mechanics 16, 695–706. and Design 88, 49–73.
Sobolev, V.A., Kuzavkov, N.G., 1994. Identification of places with fluid temperatures Velusamy, K., Sundararajan, T., Chellapandi, P., Selvaraj, P., Chetal, S.C., 2005. Inves-
in BN 600 reactor and reactor systems, IAEA-IWGFR/90. In: Specialists meeting tigations of thermal striping in primary circuit of prototype fast breeder reactor.
on “Correlation Between Material Properties and Thermohydraulics Conditions In: Proceedings of the 13th International Conference on Nuclear Engineering,
in LMFBRs”, Aix-en-Provence, France, November 22–24. Beijing, China.
Tanaka, H., Toyoda, M., 1996. Random thermal fatigue in fast breeder reactor: Wakamatsu, M., et al., 1995. Attenuation of temperature fluctuations in thermal
a narrow-band spectrum. Journal of Nuclear Engineering and Design 160, striping. Journal of Nuclear Science and Technology 32 (8), 752–762.
333–345.

You might also like