Skip to main content
Log in

Mechanisms of Neutron Irradiation Hardening in Impurity-Doped Ferritic Alloys

  • Published:
Metallurgical and Materials Transactions A Aims and scope Submit manuscript

Abstract

Mechanisms of neutron irradiation hardening in phosphorus (P)-doped, sulfur (S)-doped, and copper (Cu)-doped ferritic alloys have been studied by applying a rate theory to the temperature dependence of the yield strength. Hardening behavior induced by neutron irradiation at various temperatures (473 to 711 K) is characterized in terms of the variations in athermal stress and activation energy for plasticity controlled by precipitation or solid solution, and kink-pair formation with the content and type of impurities. In P-doped alloys, neutron irradiation below 563 K brings about a remarkable increase in the athermal stress and activation energy, due to the dispersion of fine (∼1.7-nm) P-rich precipitates that is more extensive than that for the Cu-rich precipitates reported in irradiated steel. During neutron irradiation above 668 K, precipitation hardening occurs to some extent in Cu-doped and S-doped alloys, compared to small or negligible hardening in the P-doped alloys. In alloys with a low to moderate content of various dissolved impurities subjected to high-temperature irradiation, the formation of kink pairs becomes considerably difficult. Differing dynamic interactions of dissolved and precipitated impurities, i.e., P and Cu, with the nucleation and growth of dislocations are discussed, giving rise to irradiation hardening.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22

Similar content being viewed by others

Notes

  1. DO-controlled hardening has usually been characterized by the yield strength at room temperature ((σ y ) RT ). Two linear relationships between the \( \sigma ^{o}_{a} \) and (σ y ) RT were found, depending on the activation energy. The (σ y ) RT is 75 pct of the \( \sigma ^{o}_{a} \) in the group I and II alloys with ΔF o below ∼600 kJ/mole and 87 pct of the \( \sigma ^{o}_{a} \) in the group III alloys irradiated below 563 K, with ΔF o above ∼800 kJ/mole.

Abbreviations

P y :

yield load

t o :

thickness of small punch specimen

δ :

displacement

t :

time

σ y, :

tensile yield stress

τ s :

shear flow stress

ε and γ :

tensile and shear strain, respectively

/dt and /dt :

tensile and shear strain rate, respectively

\( \sigma ^{o}_{a} \) and \( \sigma ^{l}_{a} \) :

DO- and LR-controlled athermal stress, respectively

ΔF o and ΔF l :

DO- and LR-controlled activation energy, respectively

\( \Delta \sigma ^{{\text{o}}}_{a} \) and \( \Delta \sigma ^{l}_{a} \) :

DO- and LR-controlled hardening, respectively

\( \Delta \sigma ^{s}_{y} \) and \( \Delta \sigma ^{p}_{y} \) :

hardening caused by solid solution and precipitation, respectively

μ :

shear modulus

ν :

Poisson’s ratio

α, β, φ, and χ :

dimensionless constant

ψ :

gradient of yield strength with respect to absolute temperature

R:

gas constant

T :

absolute temperature

C P, C Cu, and C S :

atomic percent or fraction of P, Cu, and S

b :

magnitude of Burgers vector

ρ m :

density of mobile dislocations

V :

average velocity of dislocations

ΔG :

Gibbs free energy

κ s , κ p :

dimensionless dislocation-pinning factor related to solid and precipitation hardening

ω :

critical angle at which a dislocation cut an obstacle

L :

spacing of precipitates or impurities

d :

diameter of precipitate

f v :

volume fraction of precipitate

q :

scattering vector

I(q):

neutron scattering intensity

Q :

invariant

λ :

neutron wavelength

θ :

neutron scattering angle

ϕ :

angle with respect to the direction of applied magnetic field

Δρ :

difference in scattering length density in the matrix and precipitate

References

  1. G.E. Lucas, G.R. Oddete, P.M. Lombrozo, and J.W. Seecked: in Effect of Radiation on Materials, ASTM STP 870, F.A. Garner and J.S. Perrin, eds., ASTM INTERNATIONAL, West Conshohocken, PA, 1985, pp. 900–30

  2. W.J. Phythian, C.A. English: J. Nucl. Mater., 1993, vol. 205, pp. 162–70

    Article  CAS  Google Scholar 

  3. K. Fukuya, K. Ohno, and H. Nakata: Microstructural Evolution in Reactor Vessel Steels, INSS Monograph No. 1, Institute of Nuclear Safety, Inc., Mihama, Fukui, Japan, 2001

  4. K. Onizawa and M. Suzuki: Effects of Radiation on Materials: 20th Int. Symp., ASTM STP 1405, R.S. Rosinski, M.L. Grossbeck, T.R. Allen, and A.S. Kumar, eds., ASTM INTERNATIONAL, West Conshohocken, PA, 2001, pp. 79–96

  5. J. Kameda, A.J. Bevelo: Acta Metall., 1989, vol. 12, pp. 3283–96

    Google Scholar 

  6. Y. Nishiyama, K. Onizawa, and M. Suzuki: J. ASTM Inter., 2007, vol. 4, no. 8, JAI100690

  7. A.V. Nikolaeva, Y.A. Nikolaeva, A.M. Kryukov: J. Nucl. Mater., 1994, vol. 218, pp. 85–93

    Article  Google Scholar 

  8. G. Faulkner, S. Song, P.E.J. Flewitt: Metall. Mater. Trans. A, 1996, vol. 27A, pp. 3381–90

    Article  CAS  Google Scholar 

  9. A.V. Barashev, S.I. Gulubov, D.J. Bacon: in Microstructure Processes in Irradiated Materials, G.E. Lucas, L.L. Snead, M.A. Kirk Jr., R.G. Elliman, eds., Materials Research Society, Pittsburgh, PA, 2001, p. R6.8

    Google Scholar 

  10. J. Kameda, X. Mao: Mater. Sci. Eng., 1989, vol. A112, pp. 143–49

    CAS  Google Scholar 

  11. J. Kameda, Y. Nishiyama, T.E. Bloomer: Surf. Int. Anal., 2001, vol. 31, pp. 522–31

    Article  CAS  Google Scholar 

  12. Y. Nishiyama, T.E. Bloomer, J. Kameda: in Microstructure Processes in Irradiated Materials, G.E. Lucas, L.L. Snead, M.A. Kirk Jr., R.G. Elliman, eds., Materials Research Society, Pittsburgh, PA, 2001, p. R6.10

    Google Scholar 

  13. Y. Nishiyama and J. Kameda: Japan Atomic Energy Agency, Tokai, Ibaraki, Japan, unpublished research, 2007

  14. Y. Nishiyama, K. Onizawa, M. Suzuki, J.W. Anderegge, Y. Nagai, T. Toyama, M. Hasegawa, and J. Kameda: submitted for publication

  15. H. Frost, M.F. Ashby, Deformation-Mechanism Maps—The Plasticity and Creep of Metals and Ceramics, Pergamon Press, Oxford, United Kingdom, 1982, pp. 6–9

    Google Scholar 

  16. P. Guyot, J.E. Dorn: Can. J. Phys., 1976, vol. 45, pp. 983–1016

    Google Scholar 

  17. U.F. Kocks, A.S. Argon, and M.F. Ashby: Progr. Mater. Sci., 1975, vol. 19

  18. K.C. Russell, L.M. Brown: Acta Metall., 1972, vol. 20, pp. 969–74

    Article  CAS  Google Scholar 

  19. A. Fujii, M. Nemoto, H. Suto, K. Monma: Trans. Jpn. Inst. Met. Suppl., 1968, vol. 9, pp. 374–80

    CAS  Google Scholar 

  20. R. Kasada, T. Kitao, K. Morishita, H. Matsui, and A. Kimura: Effects of Radiation on Materials: 21st Int. Symp., ASTM STP 1405, R.S. Rosinski, M.L. Grossbeck, T.R. Allen, and A.S. Kumar, eds., ASTM INTERNATIONAL, West Conshohocken, PA, 2001, pp. 237–46

  21. L.M. Brown, R.K. Ham: in Strengthening Method in Crystals, A. Kelly, R.B. Nicholson, eds., Elsevier, London, 1971, pp. 9–136

    Google Scholar 

  22. J.M. Baik, J. Kameda, O. Buck: Scripta Metall., 1983, vol. 17, pp. 1443–47

    Article  CAS  Google Scholar 

  23. X. Mao, H. Takahashi: J. Nucl. Mater., 1987, vol. 150, pp. 42–50

    Article  CAS  Google Scholar 

  24. J. Kameda, X. Mao: J. Mater. Sci., 1992, vol. 27, pp. 983–89

    Article  CAS  Google Scholar 

  25. H. Matsui, S. Moriya, S. Takaki, H. Kimura: Trans. JIM, 1978, vol. 19, pp. 163–70

    CAS  Google Scholar 

  26. C.G. Glinka, J.C. Barker, B. Hammounda, S. Kyueger, J.S. Moyer, W.J. Orts: J. Appl. Cryst., 1998, vol. 31, p. 430

    Article  CAS  Google Scholar 

  27. Neutron , X-Ray and Light Scattering, P. Lindner and T.H. Zemb, eds., Elsevier Science Publishers B.V., Amsterdam, Netherland, 1991, p. 42

  28. A.J. Allen, D. Gavillet, J.R. Weertman: Acta Metall., 1993, vol. 41, p. 1869–84

    Article  CAS  Google Scholar 

  29. K. Kitao, R. Kasada, A. Kimura, H. Nakata, K. Fukaya, K. Matsui, and M. Narui: in Effects of Radiation on Materials, ASTM STP 1447, M.L. Grossbeck, T.R. Allen, R.G. Lott, and A.S. Kumar, eds., ASTM INTERNATIONAL, West Conshohocken, PA, 2004, pp. 365–75

  30. J.P. Hirth and J. Lothe: in Theory of Dislocations, John Wiley & Sons, Inc., New York, NY, 1982, pp. 537–39

  31. V.N. Svechnikov, V.M. Pan, A.K. Shurin: Fiz. Metal. Metalloved., 1958, vol. 6, pp. 662–64

    CAS  Google Scholar 

  32. W.H. Herrnstein III, F.H. Beck, M.G. Fontana: Trans. TMS-AIME, 1968, vol. 242, pp. 1049–56

    CAS  Google Scholar 

  33. M. Peraz, F. Ferrard, V. Massardier, X. Kleber, A. Deschamps, H. Demonestrol, P. Pareige, and G. Covarel: Phil. Mag., 2005, vol. 85, pp. 2197–2210

    Article  Google Scholar 

  34. S.L. Maydet and K.C. Russell: J. Nucl. Mater., 1977, vol. 64, p. 101

    Article  CAS  Google Scholar 

  35. J. Kameda, T.E. Bloomer: Acta Mater., 1999, vol. 47, pp. 893–903

    Article  CAS  Google Scholar 

  36. R.O. Scattergood, D.J. Bacon: Acta Metall. 1982, vol. 30, pp. 1665–77

    Article  Google Scholar 

  37. Y.N. Osetsky, D.J. Bacon, V. Mohles: Phil. Mag. 2003, vol. 83, pp. 3623–41

    Article  CAS  Google Scholar 

  38. H.P. Scott, S. Huggins, M.R. Frank, S.J. Maglio, C.D. Martin, Y. Meng, J. Santillan, and G. Williams: Geophys. Res. Lett., 2007, vol. 34, p. L06302

    Article  Google Scholar 

  39. K. Onizawa: Japan Atomic Energy Agency, Tokai, Ibaraki, Japan, private communication

  40. A. Sato, M. Meshii: Phys. Status Solidi A, 1975, vol. 28, pp. 561–69

    Article  CAS  Google Scholar 

  41. M. Guttmann, D. McLean: in Interfacial Segregation, C. Johnson, J.M. Blakely, eds., ASM, Metals Park, OH, 1979, pp. 261–348

    Google Scholar 

  42. M. Yamaguchi, Y. Nishiyama, H. Kaburaki: Phys. Rev. B, 2007, vol. 76, pp. 035418–22

    Article  Google Scholar 

  43. D. McLean: Grain Boundaries in Metals, Oxford University Press, London, 1957, pp. 116–49

    Google Scholar 

Download references

Acknowledgments

This research was partly supported by the United States Department of Energy, the Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, while one of the authors (JK) was affiliated with Ames Laboratory operated by Iowa State University under Contract No. W-7405-ENG-82. The authors also acknowledge the support of the National Institute of Standard and Technology, United States Department of Commerce, in providing the neutron research facilities used in this experiment.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. Kameda.

Additional information

Manuscript submitted September 17, 2007.

Appendices

Appendix I

1.1 Rate theory

A fundamental principle of rate controlling theory is that the shear strain rate (/dt) is proportionally related to the average velocity of dislocations (V) and the density of mobile dislocations (ρ m ):[15]

$$ d\gamma /dt = \rho _{m} {\mathbf{b}}V $$
(AI-1)

where b is the magnitude of the Burgers vector. Under steady state, the value of ρ m parabolically increases with increasing shear flow stress (τ s ); i.e.,

$$ \rho _{m} = \alpha {\left( {\tau _{s} /\mu {\mathbf{b}}} \right)}^{2} $$
(AI-2)

where μ is the shear elastic modulus and α is a dimensionless constant. The dislocation mobility is thermally activated under the shear flow stress and is appreciably changed depending on two types of barriers, such as discrete obstacles and lattice resistance. The rate of dislocation motion can be described in terms of the Gibbs free energy (ΔG(τ s )) that depends on the flow stress

$$ V = \beta {\mathbf{b}}\nu {\text{ exp }}{\left[ { - \Delta G{\left( {\tau _{s} } \right)}{\text{/R}}T} \right]} $$
(AI-3)

where β is a dimensionless constant, R is the gas constant, ν is the Debye frequency, and T is the absolute temperature. The magnitude of activation energy is related to how the nucleation and growth of dislocations are influenced by discrete obstacles such as solute atoms, precipitates, and forest dislocations, and by lattice resistance. Although the activation area is theoretically introduced to describe the effect of flow stress, the Gibbs free energy is phenomenologically described in Eq. [AI-4], because of ambiguous physical meanings of the activation area indicating the temperature dependence:[15,16]

$$ \Delta G{\left( {\tau _{s} } \right)} = \Delta F^{i} {\left\{ {1 - {\left( {\tau _{s} /\tau ^{i}_{a} } \right)}^{p} } \right\}}^{q} $$
(AI-4)

where ΔF i is the Hermholtz free energy not influenced by the stress; \( \tau ^{i}_{a} \) is the athermal shear stress at 0 K; p and q are empirical factors that would depend on the shape of dislocation barriers and the superscript of the activation energy and athermal stress; and subscript i = o and l, for DO- and LR-controlled plasticity. Using Eqs. [AI-1] through [AI-4], the rate equation for the dislocation motion is given by

$$ d\gamma /dt = \alpha \beta \nu {\left( {\tau _{s} /\mu } \right)}^{2} {\text{exp }}{\left[ { - {\left( {\Delta F^{i} /{\text{R}}T} \right)}{\left\{ {1 - {\left( {\tau _{s} /\tau ^{i}_{a} } \right)}^{p} } \right\}}^{q} } \right]} $$
(AI-5)

In the case of the DO-controlled glide, (τ s /μ)2, p and q can be unity because of the large value of ΔF o. The pre-exponential term [(/dt) o , = αβν] is empirically given by 106 s−1. Converting the shear flow stress and strain components into the tensile yield stress (σ y ) with σ y  = √3τ s and ε = γ/√3, the rate equation for DO-controlled plasticity using the tensile stress and strain components is given by

$$ d\varepsilon /dt = {\left\{ {10^{6} /\surd 3} \right\}}{\text{exp }}{\left[ { - {\left( {\Delta F^{o} /{\text{R}}T} \right)}{\left\{ {1 - {\left( {\sigma _{y} /\sigma ^{o}_{a} } \right)}} \right\}}} \right]} $$
(AI-6)

For LR-controlled plasticity with low ΔF l, p = 3/4, q = 4/3, and (/dt) o  = 1011 s−1 are experimentally used, and the rate equation is

$$ d\varepsilon /dt = {\left\{ {{\left( {10^{{11}} /3\surd 3} \right)}} \right\}}{\left( {\sigma _{y} /\mu } \right)}^{{\text{2}}} {\text{exp }}{\left[ { - {\left( {\Delta F^{l} /{\text{R}}T} \right)}{\left\{ {1 - {\left( {\sigma _{y} /\sigma ^{l}_{a} } \right)}^{{3/4}} } \right\}}^{{4/3}} } \right]} $$
(AI-7)

The temperature dependence of the shear modulus of iron is given by

$$ \mu {\text{ (MPa)}} = 35.36{\left( {2050 - 0.81T} \right)} $$
(AI-8)

Plasticity controlled by obstacles and lattice resistance, which are affected by the type and quantity of impurities and neutron irradiation, can be characterized in terms of the athermal stress and activation energy, determined using Eqs. [AI-6] through [AI-8].

According to Reference 15, the dislocation mobility through discrete obstacles and lattice resistance can be categorized into three groups with different activation energies: (1) for weak obstacle strength (lattice resistance and solid solution), ΔF < 0.2 μ b 3 (120 kJ/mole); (2) for a medium strength (dislocations, defect clusters, and small or weak precipitates), 0.2 μ b 3 (120 kJ/mole) <ΔF < 2 μ b 3 (1200 kJ/mole); (3) for a strong strength (large or strong precipitates), ΔF > 2 μ b 3 (1200 kJ/mole).

The athermal stress is inversely related to the obstacle spacing (L) by

$$ \sigma _{a} = \kappa _{o} \mu {\mathbf{b}}/L $$
(AI-9)

where κ o is a dimensionless dislocation-pinning factor depending on the type of obstacles, i.e., solute solution (κ s ) and precipitate (κ p ).[15,17,30] The inverse obstacle spacing (1/L) is given by χC i /b and φf v /d for solid solution and precipitation, where χ and φ are constant, C i is the solute content, and f v and d are the volume fraction and diameter of precipitates, respectively.[21] Then σ a is given by

$$ \sigma _{a} = \kappa _{s} \chi \mu \surd C_{i} {\text{ for solid solution}} $$
(AI-10a)
$$ \sigma _{a} = \kappa _{p} \varphi \mu {\left( {{\mathbf{b}}{\text{/d}}} \right)}\surd f_{v} {\text{ for precipitation}} $$
(AI-10b)

The pinning strength changes as a function of the type and size of the medium-to-strong precipitates, that is, κ p is proportionately related to the critical angle (ω) at which a dislocation can cut an obstacle:[18,36,37]

$$ \kappa _{p} \propto {\text{cos }}{\left( {\omega /2} \right)} = {\left\{ {1 - {\left( {\mu _{p} {\text{/}}\mu _{{{\text{Fe}}}} } \right)}^{{\text{2}}} } \right\}}^{{1/2}} $$
(AI-11a)
$$ \kappa _{p} \propto {\text{cos }}{\left( {\omega /2} \right)} = {\left\{ {{\text{ln (}}d{\text{)}} + 0.7} \right\}}/2\pi {\text{ when }}d\ll{\text{ }}L $$
(AI-11b)

Appendix II

1.1 Analysis of small-angle neutron scattering

The neutron scattering intensity can be given as a function of the scattering vector (q), which is related to the neutron wavelength (λ) and the neutron scattering angle (θ):

$$ {\mathbf{q}} = 4\pi {\text{ sin }}{\left( {\theta /2} \right)}/\lambda $$
(AII-1)

To analyze precipitates induced by neutron irradiation, the difference in the neutron scattering intensity between the irradiated and unirradiated alloys (ΔI(q)) was measured. The volume fraction of precipitates (f v ) is related to the invariant of neutron scattering intensity (Q) that is given by q 2ΔI(q)d q and the difference in the scattering length density in the matrix and precipitate (Δρ) as given in[26,27]

$$ f_{v} = Q/2(\pi \Delta \rho )^{2} $$
(AII-2)

Assuming the atomic mass is constant, the Δρ can be given by superimposing the nuclear and magnetic scattering, is

$$ {\left( {\Delta \rho } \right)}^{{\text{2}}} = {\left\{ {{\left( {\rho _{n} } \right)}_{p} - {\left( {\rho _{n} } \right)}_{{{\text{Fe}}}} } \right\}}^{{\text{2}}} + {\left\{ {{\left( {\rho _{m} } \right)}_{p} - {\left( {\rho _{m} } \right)}_{{{\text{Fe}}}} } \right\}}^{{\text{2}}} {\text{ sin}}^{{\text{2}}} \phi $$
(AII-3)

where ϕ is the angle with respect to the direction of an applied magnetic field. Then

$$ {\left( {\Delta \rho } \right)}^{2}_{h} = {\left\{ {{\left( {\rho _{n} } \right)}_{p} - {\left( {\rho _{n} } \right)}_{{{\text{Fe}}}} } \right\}}^{{\text{2}}} {\text{ for the horizontal direction: }}\phi = 0{\text{ deg}} $$
(AII-4a)
$$ {\left( {\Delta \rho } \right)}^{2}_{v} = {\left\{ {{\left( {\rho _{n} } \right)}_{p} - {\left( {\rho _{n} } \right)}_{{{\text{Fe}}}} } \right\}}^{{\text{2}}} + {\left\{ {{\left( {\rho _{m} } \right)}_{p} - {\left( {\rho _{m} } \right)}_{{{\text{Fe}}}} } \right\}}^{{\text{2}}} {\text{ for the vertical direction: }}\phi = 90{\text{ deg}} $$
(AII-4b)

where the nuclear and magnetic scattering length density of the precipitate were estimated using (ρ n ) p  = Σc i (ρ n ) i and (ρ m ) p  = Σc i (ρ m ) i , where c i is the atomic fraction of each element.[28] The average volume fraction of precipitates was determined by the invariants in the horizontal and vertical direction (Q h and Q v ), experimentally measured together with (Δρ) h and (Δρ) v . In addition, the average diameter of precipitates (d) is derived using

$$ \Delta I({\mathbf{q}}) \propto {\text{exp }}( - {\mathbf{q}}^{2} d^{2} /20) $$
(AII-5)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Nishiyama, Y., Liu, X. & Kameda, J. Mechanisms of Neutron Irradiation Hardening in Impurity-Doped Ferritic Alloys. Metall Mater Trans A 39, 1118–1131 (2008). https://doi.org/10.1007/s11661-008-9482-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11661-008-9482-9

Keywords

Navigation