Вы находитесь на странице: 1из 5

130

Journal

of Nuclear

Materials

1799181

(1991) 130-134 North-Holland

Low-temperature irradiation structural materials *


M.L. Grossbeck and L.K. Mansur

creep of fusion reactor

Oak Ridge National Laboratory,

P.O. Box 2008, Oak Ridge, TN 378314376,

USA

creep has been investigated in the Oak Ridge Research Reactor in an assembly spectrally tailored to achieve a in austenitic stainless steels. Temperatures of 60-4OOC were investigated to address the requirements of near term fusion devices. It was found that austenitic alloys, especially PCA, have higher creep rates at 60C than at 330 and 400C. Since this phenomenon could not be explained by existing theoretical models, a new mechanism was proposed and a corresponding theoretical model was developed. Since vacancy migration times can be a few orders of magnitude longer than the irradiation times in this temperature regime, the immobile vacancies do not cancel climb produced by mobile interstitials absorbed at dislocations. The result is a high climb rate independent of stress-induced preferred absorption (SIPA) mechanisms. Preliminary calculations indicate that this mechanism coupled with preferred-absorption-driven glide at higher temperatures predicts a high creep rate at low temperatures and a weak temperature dependence of irradiation creep over the entire temperature range investigated.

Irradiation

He: dpa ratio of 12-14: 1 appm/dpa

1. Introduction
Although irradiation creep has been known to occur below 200C for the past 20 years [l], little attention has been paid to it since most applications for reactor structural materials have been for fission power reactors that operate at higher temperatures. In addition, the effect was believed to be a transient of short duration. Near term fusion devices such as the International Thermonuclear Experimental Reactor (ITER), will operate with high neutron flux components at temperatures below 100C [2]. This temperature is below the swelling regime for all structural alloys, but our recent work has shown that irradiation creep can be as large an effect at 100C as at 500C. The effects of irradiation creep can be beneficial in a fission reactor in that it relieves stresses induced by swelling and allows fuel to flow into a central void rather than bulge its cladding. It is also possible for irradiation creep to induce plastic flow at the tip of a crack thus blunting and arresting its propagation. Nonetheless, irradiation creep must be considered in the design of fusion reactor components since, in addition to the beneficial effects noted above, it may also lead to bulging, bowing, or other unacceptable deformation of structural components. The magnitude of the effect of irradiation creep is illustrated in fig. 1 where the stress in a tightened bolt is shown as a function of displacement level. If a bolt is torqued to a stress of 250 MPa and placed in a neutron flux, irradiation creep, based on the data reported in this paper, will relieve the stress by the time a displacement level of 10 dpa has been attained. Of course, the

first wall of a fusion reactor will not be bolted, but the example illustrates that irradiation creep is an important source of deformation and must be considered in the design of fusion devices. 2. Experimental procedures Both austenitic and ferritic/martensitic alloys were investigated. AISI Type 316 stainless steel in 20% coldworked and in annealed conditions [3] as well as USPCA [4] and Japanese PCA [5] were investigated. The ferritic alloys Sandvik HT-9 [6], and a laboratory heat of the pure binary Fe-15Cr were also investigated to study irradiation creep in normally low swelling alloys. Tube specimens 4.57 mm in diameter and 25.4 mm long with a wall thickness of 0.25 mm (0.18 for HT-9) were pressurized with helium to exert desired stresses between 50 and 400 MPa. Atomic displacement levels of

1 * Research sponsored by the Office of Fusion Energy and the Division of Materials Sciences, US Department of Energy, under contract DE-AC05-840R21400 with Martin Marietta Energy Systems, Inc. 0022-3115/91/$03.50 0 1991 Elsevier Science Publishers DISPLACEMENT LEVEL (dpa)

Fig. 1. Relief of stress in a bolt, originally torqued to a stress of 250 MPa, exposed to a neutron flux.

B.V. (North-Holland)

M.L. Grossbeck, L.K. Mansur / Low-temperature


0.6

irradiation creep

131

ANNEALED 0.4

700

200

300
STRESS,

400
MPa

500
EFFECTIVE STRESS. MPa

EFFECTIVE

Fig. 2. Total irradiation creep deformation in austenitic alloys irradiated in a spectrally tailored neutron flux in the Oak Ridge Research Reactor to a displacementlevel of 8 dpa.

7-8 dpa were achieved and helium levels of 60 to 125 appm, depending upon the alloy, were produced by irradiation in the Oak Ridge research Reactor (ORR). The tubes were profiled with a non-contacting laser micrometer system with a precision of - 250 nm. Details of the experiments are given by Grossbeck et al. [71. 3. Results and discussion Figs. 2 and 3 show the measured creep of austenitic stainless and ferritic steels, respectively, for three temperatures: 60, 330, and 400C. The creep rates are shown in table 1. At 330 and 4OOC, the creep rates are nearly the same, in agreement with previous similar experiments indicating a very small temperature dependence in the range of 330 to 600C [3,8]. The creep rates for the ferritic steels are about a factor of 3 to 6 lower than for austenitic alloys at 330 and 400 o C. An exception is the binary Fe-15Cr alloy at 400C which has a creep rate comparable to the austenitic alloys. Since this alloy is a high purity laboratory heat, it is likely to be weak in thermal creep which could account for the large deformation at 400C. At 60C creep rates are higher than at 330 and 400C for all alloys. The most pronounced differences occur in US PCA and in the two ferritic alloys where creep rates 6 to 18 times higher than at 330 and 400C are observed at 60C. The weak temperature dependence of irradiation creep has been known for many years. Hudson et al. [9] observed a rather abrupt reduction in the temperature dependence of creep between 500 and 550C in Type

321 stainless steel using proton irradiation. The low temperature portion of the curve obtained for creep rate was weakly temperature dependent and was attributed to the phenomenon of irradiation creep. This temperature insensitivity below the thermal creep regime has been confirmed by Puigh [lo], Grossbeck and Horak [4], and cited in a review by Straalsund [ll]. Since most of the neutron irradiations have been done in relation to liquid metal breeder reactor development programs, temperatures below 400C were seldom investigated. However, perhaps the first observation of irradiation creep was reported by Konobeevsky et al. [12] and by Zaimovsky et al. [13] in uranium at a temperature of only 220C (fig. 4). Hesketh [l] later observed irradiation creep in nickel and Zircaloy at -195C. In a
1.2, I

::i:
0
700 200 EFFECTIVE 300 STRESS, 400 MPa 500

Fig. 3. Total irradiation creep deformation in fenitic/ martensitic alloys irradiated in the Oak Ridge Research Reactor to a displacementlevel of 8 dpa.

132

M. L. Grossbeck, L. K. Mansur / Low-temperature irradiation creep

1.6 55

E
6

1.2
0.8

w n 0.4

0.0
TIME (h) Fig. 4. Deformation in 220C with a neutron deformation ceases when is shutdown (dashed uranium under reactor irradiation at flux of 6X 1016 n/m2 s. Note that either stress is removed or the reactor lines). After Zaimovsky et al. [13].

comprehensive review of irradiation creep, Gilbert [14] suggested that the creep rate increases with decreasing temperature. Although irradiation creep has been observed at low temperatures, it is not understood in terms of currently accepted theory. Perhaps the most widely accepted mechanisms are stress induced preferential absorption (SIPA) [15-171 and the climb-enabled glide mechanism [18], which is driven by SIPA and other climb processes that enable glide and produce creep. Both of these mechanisms lead to a linear dependence on interstitial concentration and diffusivity. For SIPA, we have i = @LDiC,Zi and for SIPA-driven (= $E/b( io Iclimb-enabled-glide we have (2) (1)

where i is the creep rate, D is the atomic volume, L is the dislocation density, D, is the interstitial diffusivity, Ci, is the interstitial concentration, Zi, is the differential bias resulting from the preferred orientation, E is the elastic modulus, and b is the Burgers vector. The dependence on DiCi leads to a reduction in creep rate of about two orders of magnitude upon decreasing the temperature from 300 to 60C. Using analytical expressions for dislocation network and loop densities that have been fit to experimental data [7] in eq. (2) leads to the steady state curve plotted in fig. 5. A large decrease in irradiation creep is predicted as temperature decreases to 60C. This curve is based upon steady-state defect concentrations. However, at a temperature of 60X, the time for vacancies to diffuse to sinks and reach a steady state concentration is much longer than the time of an experiment, of the order of 100 years for the parameters used here. Thus steady-state concentrations of vacancies and, therefore interstitials, are never reached at low temperatures, and steady-state models must be discarded. A new analysis based on the more fundamental point defect concentration behavior is required. The rate equations for point defect concentrations are: SC,/& Xv/& = G - RC,C, - K,C,, = G - RCvCi - K,C,, (3)

aL)%DiCiZi,

TEMPERATURE

(C)

Fig. 5. Creep deformation per unit stress as a function of temperature for austenitic alloys using the conventional climbenabled glide model (steady-state) and using the proposed transient model. The sum of the two curves is also shown.

where G is the point defect production rate; R is the recombination coefficient; Ki v = Si,D,, is the rate constant for interstitial or vacancy absorption at sinks, depending upon the subscript i or v, respectively; S is the sink strength; and D is the diffusivity of interstitials or vacancies. For the steady state case, the time derivatives were set equal to zero. The two resulting algebraic equations were then solved for Ci, and C, which were substituted into eqs. (1) and (2). However, in the case of low temperatures, where steady state conditions are not met, eqs. (3) must be solved as time dependent differential equations. In order to make an assessment of the magnitude of low temperature transient creep, a conservative estimate of creep rate can be made by examining point defect behavior as their concentrations approach steady state. The approach to the steady state of defect concentrations is shown schematically in fig. 6 based on Mansur [I9]. The concentrations of vacancies and interstitials are equal to each other and increase at the same rate under irradiation until interstitials begin to be absorbed by sinks. This happens in a time 7i = l/K,, where Ki is the rate constant for loss due to absorption at sinks, which even at these low temperatures is of the order of microseconds or less. At this time, the loss rate of interstitials due to absorption at sinks balances the production rate, and a quasi-steady state is achieved with the interstitial concentration equal to G/K,. AS the

M.L. Grossbeck, L.K. Mansur / Low-temperature

irradiation creep

133

El

/I

ulus, b is the Burgers vector, and Ni, is the total number of interstitials absorbed by the dislocations [7]. The number of interstitials absorbed, Ni, is estimated in each numbered region of fig. 6 making the approximations given elsewhere [7]. The result is: Nr = Gru, - G/k,, N2 = 0.5G( TV - eat) - Ci( rn2), Ns = CyDiSi( T- rM), (5) (6) (7)

Ti

&tl

7ht2

7;

TIME

Fig. 6. Schematic representation of point defect concentrations during the transient following initiation of irradiation. After Mansur [19].

vacancy concentration continues to increase, a point, rut, is reached where recombination first becomes significant. The interstitials experience a new loss mechanism and their concentration now begins to decrease. The vacancy population continues to increase, although at a slower rate, until recombination becomes the dominant loss mechanism for defects at time rR2. Finally, at time rv = l/K,, where K, is the rate constant for absorption of vacancies at sinks, the vacancies are absorbed by sinks and this loss rate balances the vacancy production rate. At this time, both vacancy and interstitial concentrations reach steady state. It is only after this time that the theoretical models of irradiation creep previously advanced apply. At low temperatures, the time to reach steady state is far longer than the duration of the experiment so that the term transient should not be interpreted to mean short duration. Nonetheless, the interstitial concentration producing irradiation creep is higher than that predicted by steady state defect concentrations. In the SIPA mechanism, both vacancies and interstitials are absorbed by dislocations in nearly equal numbers. There is a slight preference for interstitials to be absorbed by dislocations with Burgers vectors aligned parallel to a stress axis. This slight imbalance gives rise to a net climb leading to deformation in the direction of the stress and, thus to irradiation creep. However, at low temperatures, the vacancies are essentially immobile so that only interstitials are absorbed by dislocations. Without the vacancies to cancel the climb induced by the interstitials, a greatly enhanced climb rate results. Once the dislocations climb past obstacles, they are free to glide and bow out in the direction of the stress resulting in irradiation creep by the climb-enabled glide mechanism. The deformation, cc-a, resulting from this mechanism is given by

where Cy is the steady state value of the interstitial concentration, and S, is the sink strength for interstitials at dislocations. Substituting eqs. (5)-(7) into (4) results in an expression for irradiation creep deformation per unit stress which is plotted in fig 5 together with steady-state climb-glide. [18] It is seen that the transient curve contributes primarily for low temperatures, and the steady state curve contributes primarily for high temperatures. When the two contributions are added as in fig. 5, irradiation creep is seen to be about the Same at 60C as above 300C. The peak at 200C requires further explanation. A rather primitive, although conservative, method was used to evaluate the irradiation creep deformation. A more accurate method, using a direct numerical solution to the differential equations for point defect concentrations was used by Staller et al. [20], the result of which is plotted in fig. 7. The curve for 8 dpa agrees rather well with the data for PCA but has a dip rather than a peak at 200C. This discrepancy is simply the result of inaccuracies in the first method of approximation. The above calculations use an interstitial migration energy of 0.15 eV. Recent measurements have suggested interstitial migration energies in stainless steels between

lo- :
E :;::

&./;--;q;
r i
._.-. ..*,_.,.. .--.-.-

E
0

1O-4

,_.-

CALCULATED 100 MPa 1om5

FOR 25% CW PCA AT 90 MPa. 8 dpa

----

80 dpa 8 dpa 0.8 dpa

200 TEMPERATURE

400

600

EC_,&7 =

(7rL)*(QNi/bEL),
L E is the elastic

(C)

where u is the external stress, density, D is the atomic volume,

is the dislocation mod-

Fig. 7. Creep strain as a function of temperature using the proposed transient mechanism and numerical solution of the point defect rate equations. After Staller et al. [20].

134

M. L. Grossbeck, L.K. Mansur / Low-temperature irradiation creep in terms of a new interstitial mechanism absorption. of low temperature

io-2 -

FERRITIC STEELS

transient (4) The levels

of irradiation creep observed in structural alloys can lead to significant deformations under primary loads and must be considered in the design of fusion reactors.

a ;

10-O
to-2

STEADY STATE CLIMB-GLIDE

Acknowledgements The authors wish to thank L.J. Turner, N.H. Rouse and L.T. Gibson for their assistance in making the physical measurements and in preparing the specimens. The authors are also most grateful for many helpful discussions with E.R. Gilbert and A.D. Brailsford.
600 700

-. .-

----

100

200

300

400

500
(C)

TEMPERATURE

References
[l] R.V. Hesketh, in: Proc. on Solid State Physics Research with Accelerators (Brookhaven National Laboratory, Upton, New York, 1967) p. 389. [2] ITER Design Team, ITER Concept Definition (IAEA, Vienna, 1989) p. 296. [3] S. Hamada, P.J. Maziasz, M.P. Tanaka, M. Suzuki and A. Hishinuma, J. Nucl. Mater. 155-1 57 (1988) 838. [4] M.L. Grossbeck and J.A. Horak, J. Nucl. Mater. 155-157 (1988) 1001. [5] M.P. Tanaka, S. Hamada, A. Hishinuma and P.J. Maziasz, J. Nucl. Mater. 155-157 (1988) 801. [6] R.J. Puigh, in: Proc. Effects of Radiation on Materials (ASTM, Philadelphia, 1984) p. 7. [7] M.L. Grossbeck, M. Suzuki and L.K. Mansur, in: Proc. Effects of Radiation on Materials, ASTM-STP 1046, Vol. II (ASTM, Philadelphia, 1989) p. 537. [8] M.L. Grossbeck and J.A. Horak, Trans. ANS 55 (1987) 299. [9] J.A. Hudson, R.S. Nelson and R.J. McElroy, J. Nucl. Mater. 65 (1977) 279. [lo] R.J. Puigh, J. Nucl. Mater. 141-143 (1986) 954. [II] J.L. Straalsund, in: Proc. Radiation Effects in Breeder Reactor Structural Materials (AIME, New York, 1977) p. 191. [12] ST. Konobeevsky, N.F. Pravdyuk and V.I. Kutaitsev. in; Proc. Peaceful Uses of Atomic Energy, Vol. 7 (United Nations, Geneva, 1955) p. 433. G.Y. Sergeev, V.V. Titova, B.M. Levitsky 1131 AS. Zaimovsky, and Y.N. Sokursky, in: Proc. 2nd Conf. on Peaceful Uses of Atomic Energy (United Nations, Geneva, 1958) p. 566. 1141 E.R. Gilbert, Reactor Technol. 14 (1971) 258. [I51 W.G. Wolfer and M. Ashkin, J. Appl. Phys. 47 (1976) 791. Philos. Mag. 29 (1974) [161 P.T. Heald and M.V. Speight, 1075. [I71 R. Bullough and J.R. Willis, Philos. Mag. 31 (1975) 855. L.K. Mansur, Philos. Mag. A 39 (1979) 497. WI [I91 L.K. Mansur, Nucl. Technol. 40 (1978) 5. R.E. Stoller, M.L. Grossbeck and L.K. Mansur, in: Proc. WI Effects of Radiation on Materials (ASTM, Philadelphia, 1990) to be published. 0. Dimitrov and C. Dimitrov, J. Nucl. Mater. 105 (1982) WI 39. M.L. Grossbeck, L.L. Horton and L.K. Mansur, Trans. WI ANS 60 (1989) 295.

Fig 8. Creep deformation per unit stress as a function of temperature for ferritic/martensitic alloys. Curves for the climb-enabled glide model (steady-state), the proposed transient model, and the sum of the two curves are shown

0.9 and 1.0 depending upon the composition [21]. Using interstitial migration energies in this range in our calculations was found to make no perceptible difference in the calculated irradiation creep rates. The reason is that the migration times of interstitials to sinks remain sufficiently short in the temperature range of interest to insure that interstitials begin arriving at dislocations in the early stages of an irradiation. A similar analysis can be done for the ferritic steels investigated, and fig. 8 shows the results using the previously described method of estimation of absorbed interstitials [22]. A curve qualitatively similar to that for austenitic stainless steels was obtained. Since there are significant uncertainties in measuring dislocation loop concentrations in the microstructures formed during low temperature irradiations, the dislocation loop density was varied by a factor of 100 in each direction. Reducing the loop concentration made little difference, but increasing the loop concentration showed a higher level of irradiation creep at 60C than at 300C which was observed experimentally. A more accurate analysis will be carried out based on the numerical solution of the defect rate equations using measured microstructural data and, perhaps, time dependent microstructures when the actual specimens are sectioned and examined by electron microscopy. 4. Conclusions
(1)

High levels of irradiation creep have been observed in austenitic and ferritic alloys at temperatures in the operating range of near-term fusion reactors
such as the ITER. creep at 60C was greater creep than at 330 and

(2) The 400C. (3) Low

temperature

irradiation

can be explained

Вам также может понравиться