You are on page 1of 10

International Journal of Engineering Science 105 (2016) 28–37

Contents lists available at ScienceDirect

International Journal of Engineering Science


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

On the connections between plasticity parameters and


electrical conductivities for austenitic, ferritic, and
semi-austenitic stainless steels
Ahmed Kanaan, Aref Mazloum, Igor Sevostianov∗
Department of Mechanical and Aerospace Engineering, New Mexico State University Las Cruces, NM 88001, USA

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

Article history: This paper focuses on cross-property connections between plasticity parameters (yield
Received 10 February 2016 limit and hardening coefficient) and electrical conductivity of stainless steels. Comparative
Revised 24 April 2016
analysis of such connections is done for four materials that differ by their microstructure
Accepted 25 April 2016
and chemical content. The possibility of cross-property connection is provided by the fact
Available online 12 May 2016
that both plasticity parameters and electrical conductivity are governed by the same mi-
Keywords: crostructural parameter, which is the dislocations density. The cross-property connections
Stainless steel are obtained in explicit analytical form. Experimental observations are in good agreement
Electrical resistivity with theoretical results for three of the considered materials (ferritic and austenitic steels).
Yield Behavior of semi-austenitic low carbon steel 17-7 PH, however, is completely different, that
Hardening can be explained by the specific character of its microstructure. The results can be used for
Cross-property connection development of a new methodology to estimate mechanical performance of austenitic and
ferritic stainless steels with high carbon content.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction

The present research is motivated by the needs of pipe-line industry, where increasing the pipelines’ life, governed by
the development of various defects, is one of the main challenges. The monitoring and control of microstructure changes -
formation and development of dislocations, foreign particles, cracks etc. - during working life of the structural elements is
still an open problem.
The main material used to construct pipelines is stainless steel. Changes in material microstructure depend mostly on
pipelines’ installation type. In the case of offshore pipelining, pipelines are affected by corrosion, so the cathode’s protection
is usually used to control such corrosion. In the case of onshore pipelining (above-ground), pipelines are subjected to ther-
mal fatigue that yields the increase in dislocation density. The latter, in turn, leads to change in the macroscopic residual
stresses (Revie, 2015) and increase in the electrical resistivity (Watts, 1988a, 1988b).
In the present paper, we provide a comparative analysis of mechanical and electrical behavior of four stainless steels –
semi-austenitic stainless steel 17-7 PH, ferritic stainless steels 430 and austenitic stainless steels 302 and 310 - typically used
as construction material in pipelines’ industry,. These stainless steels differ from each other by microstructure and chemical
content – most importantly the carbon content which governs, in particular, the change in the yield limit during the cyclic
plasticity process (Brown, 1977; Seeger, 1958). We also propose, a methodology to control changes in the yield limit and


Corresponding author.
E-mail address: igor@nmsu.edu (I. Sevostianov).

http://dx.doi.org/10.1016/j.ijengsci.2016.04.012
0020-7225/© 2016 Elsevier Ltd. All rights reserved.
A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37 29

hardening coefficient of metals using electrical conductivity measurements. The possibility of such an approach is based on
the fact that these parameters are affected by the same microstructural parameter, which is the dislocation density (note
that the effect of the dislocation density on the said properties has been studied in literature, mostly, unconnectedly).
Dislocation density is defined as the number of dislocation lines crossing the unit area (Seeger, 1980). This parameter de-
termines the character of plastic deformation of metals (Honeycombe, 1984; Kuhlmann-Wilsdorf, 1989; Seeger, 1958). When
a metal is subjected to plastic deformation, the stress needed to endure the deformation increases with strain growth, i.e.
the yield stress increases. This occurrence can be called as work hardening (Clarebrough & Hargreaves,1959) and its nature
used to be a prominent problem, in the study of the crystals’ plasticity, starting from the earliest work in this field. After all,
substantive progress has been achieved since the entity of dislocations and the prevailing role in plastic deformation have
been elucidated. Review studies for experimental and theoretical results was made by Cottrell (1949) and (1953a). He, also
indicated (Cottrell, 1953b) that work hardening embraces the group behavior of massive number of dislocations rather than
the behavior of isolated dislocations.
Since 1960 s different aspects of the relation between the dislocation density and electrical conductivity have been dis-
cussed. It has been shown that the obtained dislocations’ resistivity for simple metals are harmonic according to the occur-
rence of resonances in the scattering near the Fermi surface (Brown, 1967a) and it can be expected that resonance happens
somewhere in the conduction band (Brown, 1967b). Watts (1987) suggested that if the dominant conduction electron scat-
tering takes place in the core regions of dislocations, then it is essential to include crystalline structure in the scattering
model – otherwise, the measured degree of isotropy of the resistivity cannot be accounted for.
To the best of our knowledge, the first attempt to relate plastic yield limits to the electrical conductivity of metals was
done by Bell, Latkowski, and Willoughby (1966), who showed how electrical conductivity of indium antimonite single crys-
tal changes due to plastic bending. Actually, it was the first observation of the cross-property connection between the yield
limit and electrical properties of metals. Generally, cross-property connections for heterogeneous materials belong to the
fundamental problems of engineering science and physics. They relate changes in different physical properties caused by
the presence of certain microstructure. Their practical usefulness lies in the fact that one physical property (say, electrical
conductivity) may be easier to measure than the other. Cross-property connections have been discussed in literature for
about half a century. Most of them had a character of qualitative observations. First quantitative theoretical results on cross-
property connections appeared in the classical work of Bristow (1960), who derived explicit elasticity-conductivity connec-
tion for a micro-cracked material in the isotropic case of random crack orientations and low crack density. Levin (1967) in-
terrelated the effective bulk modulus and the effective thermal expansion coefficient of a general two phase isotropic com-
posite. Milton (1981) established cross-property bounds for the transport and the optical constants of isotropic composites.
Similar bounds for the electrical and the magnetic properties were given by Cherkaev and Gibiansky (1992). The general
approach to establish various cross-property correlations was outlined by Milton (1996). The conductivity-elasticity cross-
property bounds have been derived in works of Berryman and Milton (1988) and Gibiansky and Torquato (1995, 1996a,
1996b). Sevostianov and Kachanov (see their review 2009 for details) established approximate cross-property connections
between elastic and conductive properties of heterogeneous materials. They also shown connections between two different
physical properties can be established if and only if these properties are governed by the same (or similar) microstructural
parameter (Kachanov & Sevostianov, 2005). This approach has been used to connect yield limit and electrical conductivity in
the papers of Dominguez and Sevostianov (2011), Omari and Sevostianov (2013) and Omari, Balázs, and Sevostianov (2014).
In these works, however, the authors did not compare different materials and did not study effect of any other parameters
(like carbon content or original microstructure, for example). We do it in the present work. We used the approach devel-
oped by Dominguez and Sevostianov (2011) and Omari and Sevostianov (2013) for quasi-static loading of stainless steel
specimens.

2. Experimental procedure

We examine, four types of stainless steels: semi-austenitic 17-7 PH, austenitic 310, ferritic 430, and austenitic 302
(see Table 1 for their chemical content and material properties) and follow the approach of Dominguez and Sevostianov
(2011) and Omari and Sevostianov (2013). For each type of stainless steel, six specimens were studied.

2.1. Specimens preparation

A water jet cutting with abrasive particles (see Fig. 1) was used to cut the different stainless steel specimens, in order
to avoid any changes in microstructure and mechanical properties. This method allows one to cut the specimens without
deforming or altering intrinsic properties. In addition, the need post-processing operations are mostly eliminated (Lorincz,
2009). The specimens were cut according to ASTM standards as shown is Fig. 2.

2.2. Cyclic plasticity tests

Mechanical tests have been performed according to ISO 6892 standards using Instron 5582 testing machine and Bluehill
software. The applied extension rate was 500 N/min, which is in the range of quasi-static loading (Dominguez & Sevostianov,
2011). The loading-unloading process were repeated four times on each specimen to follow conditions of cyclic plasticity.
30 A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37

Table 1
Properties and contents of steels used in the experiment.

Properties/contents Stainless Steels

17–7 PH 430 302 310

Material Meets These Standard(s) AMS 5528 ASTM A240 ASTM A240 AMS 5521
Yield Strength, psi 40,0 0 0 50,0 0 0 39,900 45,0 0 0
Ultimate Tensile Strength, psi 130,0 0 0 75,0 0 0 89,900 89,900
Elongation 35% 25% 55% 45%
Resistivity (microhm-cm) 80 60 72 78
Carbon (C) 0.09% max 0.12% max 0.15% max 0.25%
Iron (Fe) 70.59–76.75% 87% 70% 48–53%
Chrome (Cr) 16–18% 11% 18% 26%
Aluminum (Al) 0.75–1.5% max – – –
Manganese (Mn) 1% max 1% max 2% max 2%
Nickel (Ni) 6.5–7.75% – 9% 19–22%
Phosphorus (P) 0.04% max 0.04% max 0.045% max 0.045% max
Sulphur (S) 0.03% min 0.03% max 0.03% max 0.03%
Silicon (Si) 1% max 1% max 1% max 1.5%

Fig. 1. Illustration of abrasive water-jet cutting.

Fig. 2. Geometry and dimensions in millimeters of specimens.

The effects of load history in cyclic plasticity of metals and alloys are taken over from material strain hardening. Two types
of material hardening can be classified in tension-compression tests - kinematic and isotropic hardening (Pommier, 2003).
We can illustrate the difference between them as follows:

(I) Kinematic hardening: 302 stainless steel was tested under uniaxial and reversed cyclic strain. The material has an elastic
domain, which is midpoint on an internal stress σ . After the specimen was plastically deformed at ε = 1.5%, the corre-
sponding value of the internal stress is 175 MPa. Then, in the process of reversed plastic strain, the internal stress level
is lessened to reach σ = −175 MPa at ε = −1.5%. This effect is called kinematic hardening and can be treated as a dis-
placement of materials’ yield surface, corresponding to the direction of plastic strain (Mroz, 1967; Chaboche, 1977). In
this paper, only forward plasticity (uniaxial cyclic strain) was considered.
A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37 31

Fig. 3. Illustration of Four-point probe method.

(II) Isotropic hardening can be illustrated on example of 310 stainless steel. At ε = 0.6%, the size of the elastic domain is equal
to 290 MPa after the first elongation and reaches 390 MPa after the fourth cycle with ε = 3.65%. This effect is described
as an expansion of materials’ yield surface (Chaboche, 1977).

Most materials show both types of hardening with one of them being dominant – for austenitic stainless steels, the
dominant mechanism is isotropic hardening, while kinematic hardening is dominant for low carbon and semi-austenitic
steels stainless steel (Pommier, 2003).

2.3. Electrical resistivity measurements

Electrical resistivity was measured using HP 433B Milliohmmeter that is an accurate, dependable, and high speed test
tool to measure low resistance. Its relatively high resolution enables us to determine the slightest differences in examining of
contact resistance for stainless steel specimens using the Four-point probe method. This method is usable when the distance
between the probes is comparably small to the size of the specimen. None of the probe should be so close to the edge of
the specimen (Singh, 2013). The arrangement of probes is shown in Fig. 3 detailed description of the method is given by
Valdes (1954). The probes were locating 20 mm away from each other (a = 20 mm), in order to measure specimen resistance
after each plasticity cycle for the different stainless steel materials.

2.4. AFM study of the specimens’ surfaces

To study changes in the surface topography of the specimens due to loading, BRUKER Atomic Force Microscope with
contact imaging mode in air was used. The tip used was SNL-10 (Sharp Nitride Lever) probe. The max radius of curvature
of those probes is 12 nm. Photos were taken for the four types of stainless steels specimens.

3. Connection between plasticity parameters and electric properties

The cross-property connection between isotropic hardening coefficient of stainless steel specimens and electrical resis-
tivity is derived in this section following approach proposed by Dominguez and Sevostianov (2011) and Omari et al. (2013).
In literature, dislocations and its relationships to material’s resistance were discussed for decades and had complications
in correlating them accurately. In 1993 Karolik and Luhvich have made calculations of the residual resistivity accompanied
with dislocations and grain boundaries for metals ended up with 1-2 correspondence regarding to the experimental data.
Considering the lattice dilatation during cyclic plasticity and distortion, which is condensed primarily in the dislocation
core yielded the right correspondence of the residual resistivity produced by dislocations, Karolik and Luhvich (1993) have
derived an expression to correlate the residual resistivity due to a dislocation per unit of its density and calculated the
scattering cross section for line defects performed in the free-electron approximation using the partial-wave method. They
were also able to conclude that by using the proposed model of defects and considering the lattice dilatation in the disloca-
tion core and the being of resonance electron locates near the Fermi energy can explain the experimental results for a wide
range of metals. In 1988, Watts discussed how atomic disorder close to dislocation lines yields enough large-angle scattering
to consider for the measured values of electrical resistivity of metals. Then he showed that these discrete models can also be
considered full-filling generally for the alternation of particular dislocation resistivity among metals. As a conclusion to that,
the electrical resistivity of dislocations is comprehensible in terms of scattering from the assembly of displaced atoms in a
region within a few atomic lengths of dislocation lines. Lately, it was shown that the influence of dislocations on electrical
resistivity is associated with virtual quasi –stationary states. This approach has been used to calculate particular resistivi-
ties of dislocations that are close to experimentally obtained results. The increase in dislocation density and formation of
new grain boundaries due to the cyclic plasticity leads to change in the entire electrical resistivity of the specimenRi j − R0i j
32 A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37

. Considering the growth in the whole dislocation density tensor to be ρ ij , the change in electrical resistivity can be
expressed as:
Ri j ≡ Ri j − R0i j = ρi j Rd /N, (1)

where Rd /N represents dislocations’ resistivity per dislocation density unit, that is experimentally measured and theoretically
derived in literature (Brown, 1977; Karolik & Luhvich, 1994). Steels specimens’ microstructure is isotropic, so Rij can be
written asRδ ij and the growth in scalar dislocation density ρ = tr (ρi j ) has the following form:
 
N R − R0
ρ = (2)
Rd
Note that this expression is similar to one proposed by Kocer, Sachslehner, Müller, Schafler, and Zehetbauer (1996) to
estimate dislocation density in aluminum using electrical conductivity measurements. The estimates of this method were
compared to Transmission Electron Microscopy’s (TEM) data and showed good consistency. Particularly, this method is im-
portant when dislocation density is high (up to 1018 m−2 ), because TEM cannot be used for dislocation densities higher than
1014 m−2 .
The work-hardening theory, described, for example, by Seeger (1958); Honeycombe (1984); Kuhlmann-Wilsdorf (1989) for
low energy dislocation structures was used to determine dislocation density effect on the work-hardening coefficient θ =
dσ /dε . Counterpart dislocation density yields a decrement in the yield stress as:
√ √
σY = α Gb( ρ − ρ0 ) (3)
where ρ 0 , ρ and are initial dislocation density, current dislocation density and length of the Burgers’ vector. G is the shear
modulus and parameter α is a constant measuring the dislocation strengthening efficiency. α ≈ 1 and depends on the
mutual positions of the dislocations. In a simplest case, the relationship between the dislocation density and the plastic
strain can be expressed as:
ρ = ε /λb (4)
where λ represents the length of the dislocation free path.
Expression (3) can be solved for current dislocation density asas
 2
σY √
ρ= + ρ0 (5)
α Gb
The dislocation density varies with strain (Taylor, 1934). Generally, four main processes determine the variation of to-
tal dislocation density with strain: creation, immobilization, remobilization, and annihilation of dislocations (Tóth, 1979).
Expression (5) is similar to one proposed by Reid, Gilbert, and Rosenfeld (1965).
Dependence of the material resistivity and plastic parameters on dislocation density lead to the explicit cross-property
connections. Indeed, substitution of (2) in (3) yields
 
N √ 
σY = α Gb d
R − R0 (6)
R
The fact that dislocations cannot move freely means that plastic deformation cannot extend until the applied stresses
are increased. This is how a material is then work hardened. A material can only be strain hardened so much before it
is saturated. With cyclic plasticity, the rate of hardening progressively diminishes and a quasi-steady state of deformation,
known as “saturation”, is reached (Suresh, 1998). It is apparent that a strain hardened material is due to an increase in
dislocations within its structure and one can relate resistivity to the hardening coefficient. The latter can be represented in
terms of dislocation density as:

α G b/λ αG
θ = d σ /d ε = √ = √ (7)
2 ε 2λ ρ
Substituting (2) into (7) gives explicit expression for the hardening coefficient in terms of the resistivity change:

αG Rd
θ= (8)
2λ N ( R − R0 )
For small θ , it makes sense to invert expression (8) as

2λ (R − R0 )N
1/θ = (9)
αG Rd
In the next section we verify expressions (6) and (8) by comparison with experimental data for stainless steel. These
relations can be used

(A) As a simple method to estimate the stage of the plastic deformation from electrical measurements
(B) To recover information on microstructural parameters α and λ from combining mechanical and electrical measurements.
A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37 33

Fig. 4. Stress-strain curves for four studied steels (averaged over six specimens): (a) semi-austenitic stainless steel 17-7 PH; (b) austenitic stainless steel
310; (c) ferritic stainless steel 430; (d) austenitic stainless steel 302.

4. Results

Stress strain curves (averaged over six specimens) are presented in Fig. 4 for four different stainless steels. Note that
the curve for steel 17-7 PH substantially differs from the other ones. It may be explained by the material microstructure.
Steel 17-7PH is a precipitation hardening stainless steel, which was thermally hardened in order to precipitate carbon. This
process has increased toughness, ductility, hardness and corrosion resistance and lead to formation of very complex mi-
crostructure.
For austenitic stainless steels, the relation between applied stress and formations of microstructure has been proposed
by Byun, Lee, and Hunn (2003). The microstructure depends on altering material and testing conditions (Lee, Byun, et al.,
2001; Lee, Yoo, et al., 2001; Lee, Byun, Hunn, Farrell & Mansur, 2001; Farrell et al., 2003). Generally, the dependence of
the microstructure on applied stress can be described as follows: for stresses smaller than 400 MPa, dislocation tangles
are dominant microstructural features. For stresses in the range between 400 MPa and 600 MPa isolated and small stacking
faults are formed. The size for these stacking is less than 1 μm. For applied stress higher than 600 MPa, large stacking faults
were taking control over other microstructural formations. In addition to that, ordinary dislocations have been observed
along with large stacking faults in the same microstructures.
For ferritic stainless steels, relationships between strain amplitudes and microstructural parameters have been reported
by Kutka (1977) and Tjong (1991). They can be summarized as follows: at low strain amplitudes (0.3%), dislocation substruc-
tures are formed. These substructures are mostly composed of dislocation bundles that subsequently merged and condensed
into dipolar or multipolar walls. At intermediate strain amplitudes (0.6%), the formation of intricacy structures along with
extended wall structures is observed. These extended walls and intricacy structures are typical substructures promoted in
fatigued metals. At high strain amplitudes (2%), dislocation cells are formed inside the extended wall structures’ channels.
In the case of 17-7PH semi-austenitic stainless steel, which contains both martensitic and austenitic domains since its
chromium-nickel ratio does not allow the full formation of austenitic phase, metallurgists faced a very challenging and
complex microstructure. The existence of martensitic, carbide particles and the precipitation of intermetallic all participate
in strengthening the steel (Sarosiek & Owen, 1983). In particular, the carbide particles are so bulky to contribute in blocking
dislocation movement with the typical precipitation strengthening mechanism, which can powerfully confine the austenitic
matrix by producing stress around the precipitates and raise up the elastic limit of the material. Moreover, during uniaxial
tension for 17-7PH stainless steel, very little discontinuities yielding were noticed (Tomota, 1987). The martensite phase
occurs to restrain the occurrence of discontinuous yielding in 17-7 PH stainless steel. This restraining becomes more effective
with raising up austenitic matrix’s volume fraction and/or martensite’s hardness (Sarosiek & Owen, 1984; Rigsbee et al.,
1979).
34 A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37

Fig. 5. Resistivity-strain curves for four studied steels (averaged over six specimens): (a) semi-austenitic stainless steel 17-7 PH; (b) austenitic stainless
steel 310; (c) ferritic stainless steel 430; (d) austenitic stainless steel 302.

Fig. 6. Microscopic photos of stainless steels specimens’ surfaces before (on the left) and after (on the right) cyclic plasticity process: (a) semi-austenitic
stainless steel 17-7 PH; (b) austenitic stainless steel 310; (c) ferritic stainless steel 430; (d) austenitic stainless steel 302.

The resistivity – strain curves are given in Fig. 5. Note that stainless steel 17-7 PH’s curve is again substantially different
from ones for other stainless steels which may be explained by the material microstructure. AFM images of the specimens
before and after the loading are given in Fig. 6. Note that formation of dislocation lines can be recognized in images (a)
and (d) only. It may be explained by insufficient processing of the specimens’ surfaces. Fig. 7 shows the dependence of the

hardening coefficient on 1/ ε. Note the linear dependence (in accordance to expression (7) for all the curves except the
one for stainless steel 17-7 PH. Fig. 8. Illustrates dependence of the work-hardening coefficient on the normalized resistivity
N (R−R0 )
change Rc, where Rc= Rd . Note that linear dependence according to expression (9) is appropriate for ferritic and
A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37 35

Fig. 7. Work-hardening - Strain curves for four studied steels (averaged over six specimens): (a) semi-austenitic stainless steel 17-7 PH; (b) austenitic
stainless steel 310; (c) ferritic stainless steel 430; (d) austenitic stainless steel 302.


Fig. 8. Work-hardening- Resistivity change (Rc = N (R − R0 )/Rd ) curves for four studied steels (averaged over six specimens) (a) semi-austenitic stainless
steel 17-7 PH; (b) austenitic stainless steel 310; (c) ferritic stainless steel 430; (d) austenitic stainless steel 302.

austenitic steels. Steel 17-7 PH does not show the same tendency. It is consistent with the curves shown in Figs. 4 and 5.
The cross-property coefficient α2λG in expression (9) for all three cases (b)–(d) is about 10−19 Pa−1 .
Fig. 9. Illustrates dependence of the yield stress change in expression (6) on the normalized resistivity change Rc, where
√ 
Rc= N
Rd
( R − R0 ). The linear dependence according to expression (6) is also appropriate for ferritic and austenitic steels.
Steel 17-7 PH behaves differently. The cross-property coefficient α Gb in expression (6) for all three cases (b)-(d) is of or-
der10−25 Pa.
36 A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37


Fig. 9. Yield stress change - Resistivity change (Rc = N (R − R0 )/Rd ) curves for four studied steels (averaged over six specimens) (a) semi-austenitic stain-
less steel 17-7 PH; (b) austenitic stainless steel 310; (c) ferritic stainless steel 430; (d) austenitic stainless steel 302.

5. Concluding remarks

In the present work we focused on the cross-property connections between plastic properties of stainless steels and
electrical conductivity. It is shown that both properties are changed in the process of cyclic plasticity deformation. It is
well explained by classical theories that expressing the two properties in terms of the same micro-structural parameter
-dislocation density, which increases in the process of cyclic plasticity.
It is shown, that for austenitic and ferritic steels, the cross- property connection is in a good agreement with analytical
estimates. Experimental result, in particular allows to evaluate the parameters entering expressions for hardening coefficient
and the ultimate yield strength. For semi-austenitic steel 17-7 PH, the cross-property connection cannot be established. It
can be explained by high density of internal obstacles complicating dislocations gliding. The obtained results allow develop-
ment of a new methodology to estimate changes in plastic properties of metals using non-destructive electrical resistivity
monitoring.

Acknowledgement

Financial support from the FP7 Project TAMER IRSES-GA-2013-610547 and New Mexico Space Grant Consortium con-
tained in the NASA Cooperative Agreement NNX13AB19A to New Mexico State University are gratefully acknowledged.

References

Bell, R. L., Latkowski, R., & Willoughby, A. F. W. (1966). The effect of plastic bending on the electrical properties of indium antimonide. Journal of Materials
Science, 1(1), 66–78.
Berryman, J. G., & Milton, G. W. (1988). Microgeometry of random composites and porous media. Journal of Physics D: Applied Physics, 21(1), 87–94.
Bristow, J. R. (1960). Microcracks, and the static and dynamic elastic constants of annealed and heavily cold-worked metals. British Journal of Applied Physics,
11(2), 81.
Brown, R. A. (1967a). Resonance scattering and the electrical and thermal resistivities associated with extended defects in crystals. Physical Review, 156(3),
692.
Brown, R. A. (1967b). Electron and Phonon bound states and scattering resonances for extended defects in crystals. Physical Review, 156(3), 889.
Brown, R. A. (1977). Electrical resistivity of dislocations in metals. Journal of Physics F: Metal Physics, 7(7), 1283.
Byun, T. S., Lee, E. H., & Hunn, J. D. (2003). Plastic deformation in 316LN stainless steel–characterization of deformation microstructures. Journal of nuclear
materials, 321(1), 29–39.
Chaboche, J. L. (1977). Viscoplastic constitutive equations for the description of cyclic and anisotropic behaviour of metals. Bulletin of the Polish Academy of
Science, Series on Science and Technology, 25(1), 33–42.
Cherkaev, A. V., & Gibiansky, L. V. (1992). The exact coupled bounds for effective tensors of electrical and magnetic properties of two-component two-di-
mensional composites. Proceedings of the Royal Society of Edinburgh: Section A Mathematics, 122(1-2), 93–125.
Clarebrough, L. M., & Hargreaves, M. E. (1959). Work hardening of metals. Progress in Metal Physics, 8, 1–103.
Cottrell, A. H. (1949). Theory of dislocations. Progress in Metal Physics, 1, 77.
Cottrell, A. H. (1953a). Theory of dislocations. Progress in Metal Physics, 4, 205–264.
Cottrell, A. H. (1953b). Dislocations and plastic flow in crystals. Oxford: Clarendon Press.
A. Kanaan et al. / International Journal of Engineering Science 105 (2016) 28–37 37

Dominguez, D., & Sevostianov, I. (2011). Cross-property connection between work-hardening coefficient and electrical resistivity of stainless steel during
plastic deformation. International Journal of Fracture, 167(2), 281–287.
Farrell, K., Byun, T. S., & Hashimoto, N. (2003). ORNL/TM-2003/63 p. 471.
Gibiansky, L. V., & Torquato, S. (1995). Connection between the conductivity and bulk modulus of isotropic composite materials. Physics Transaction of the
Royal Society London A, 353, 243–278.
Gibiansky, L. V., & Torquato, S. (1996a). Rigorous link between the conductivity and elastic moduli of Bber reinforced materials. Proceedings of the Royal
Society A, 452, 253–283.
Gibiansky, L. V., & Torquato, S. (1996b). Bounds on the e1ective moduli of cracked materials. Journal of the Mechanics and Physics of Solids, 44(2), 233–242.
Honeycombe, R. W. K. (1984). The plastic deformation of metals (pp. 231–232). London: Edward Arnold Publ. Ltd.
Kachanov, M., & Sevostianov, I. (2005). On quantitative characterization of microstructures and effective properties. International Journal of Solids and Struc-
tures, 42(2), 309–336.
Karolik, A. S., & Luhvich, A. A. (1994). Calculation of electrical resistivity produced by dislocations and grain boundaries in metals. Journal of Physics:
Condensed Matter, 6(4), 873.
Kocer, M., Sachslehner, F., Müller, M., Schafler, E., & Zehetbauer, M. J. (1996, May). Measurement of dislocation density by residual electrical resistivity. In
Materials science forum: Vol. 210 (pp. 133–140).
Kuhlmann-Wilsdorf, D. (1989). Theory of plastic deformation:-Properties of low energy dislocation structures. Materials Science and Engineering: A, 113,
1–41.
Kutka, J. (1977). The metallurgy of Al-containing stainless steels. Stainless Steel’77, 49–56.
Lee, E. H., Byun, T. S., Hunn, J. D., Yoo, M. H., Farrell, K., & Mansur, L. K. (2001a). On the origin of deformation microstructures in austenitic stainless steel:
Part I—microstructures. Acta materialia, 49(16), 3269–3276.
Lee, E. H., Yoo, M. H., Byun, T. S., Hunn, J. D., Farrell, K., & Mansur, L. K. (2001b). On the origin of deformation microstructures in austenitic stainless steel:
Part II—mechanisms. Acta materialia, 49(16), 3277–3287.
Lee, E. H., Byun, T. S., Hunn, J. D., Farrell, K., & Mansur, L. K. (2001). Origin of hardening and deformation mechanisms in irradiated 316 LN austenitic
stainless steel. Journal of nuclear materials, 296(1), 183–191.
Levin, V. M. (1967). On the coefficients of thermal expansion of heterogeneous material. Mechanics Solids, 2, 58–61.
Lorincz, J. (2009). Waterjets: Evolving from macro to micro. Manufacturing Engineering, 143(5), 47–53.
Milton, G. W. (1981). Bounds on the electromagnetic, elastic, and other properties of two-component composites. Physical Review Letters, 46(8), 542.
Milton, G. W. (1996). Composites: A myriad of microstructure independent relations. In T. Tatsumi, et al. (Eds.), Theoretical and Applied Mechanics
(pp. 443–459). Elsevier.
Mroz, Z. (1967). On the description of anisotropic workhardening. Journal of the Mechanics and Physics of Solids, 15(3), 163–175.
Omari, M. A., & Sevostianov, I. (2013). Evaluation of the growth of dislocations density in fatigue loading process via electrical resistivity measurements.
International Journal of Fracture, 179(1-2), 229–235.
Omari, M. A., Balázs, T., & Sevostianov, I. (2014). Evaluation of Changes in Plastic Yield Parameters of Titanium CP-2 using Electrical Resistivity Measure-
ments. International Journal of Fracture, 187(1), 179–186.
Pommier, S. (2003). Cyclic plasticity and variable amplitude fatigue. International Journal of Fatigue, 25(9), 983–997.
Reid, G. N., Gilbert, A., & Rosenfeld, A. R. (1965). Phyl. Mag, 42, 409.
Revie, R. W. (2015). Oil and gas pipelines: Integrity and safety handbook. Hoboken, NJ: John Wiley & Sons.
Rigsbee, J. M., VanderArend, P. J., & Davenport, A. T. (1979). Formable HSLA and dual-phase steels (p. 56). Warrendale, PA: TMS-AIME.
Sarosiek, A. M., & Owen, W. S. (1983). On the importance of extrinsic transformation accommodation hardening in dual-phase steels. Scripta Metallurgica,
17(2), 227–231.
Sarosiek, A. M., & Owen, W. S. (1984). The work hardening of dual-phase steels at small plastic strains. Materials Science and Engineering, 66(1), 13–34.
Seeger, A. (1958). Kristallplastizität. Bd: Handbuch d. Physik VII/2.
Seeger, A. K. (1980). Early work on imperfections in crystals, and forerunners of dislocation theory. In Proceedings of the royal society of london. series a,
mathematical and physical sciences (pp. 173–177).
Sevostianov, I., & Kachanov, M. (2009). Connections between elastic and conductive properties of heterogeneous materials. Advances in Applied Mechanics,
42, 69–252.
Singh, Y. (2013). Electrical resistivity measurements: A review. In International journal of modern physics: Conference series: Vol. 22 (pp. 745–756). World
Scientific Publishing Company.
Suresh, S. (1998). Fatigue of materials. Cambridge: Cambridge University Press.
Taylor, G. I. (1934). The mechanism of plastic deformation of crystals. Part I. Theoretical. In Proceedings of the royal society of London. Series A, containing
papers of a mathematical and physical character (pp. 362–387).
Tjong, S. C. (1991). Transmission electron microscope observations of the dislocation substructures induced by cyclic deformation of the ferritic Fe%
25Cr%(2–4) Al and Fe% 19Cr%4Ni%2Al alloys. Materials Characterization, 26(2), 109–121.
Tomota, Y. (1987). Effects of morphology and strength of martensite on cyclic deformation behaviour in dual-phase steels. Materials science and technology,
3(6), 415–421.
Tóth, L. (1979). Note on the work-hardening of metals. Scripta Metallurgica, 13(2), 145–148.
Valdes, L. B. (1954). Resistivity measurements on germanium for transistors. Proceedings of the IRE, 42(2), 420–427.
Watts, B. R. (1987). The structure factor of a dislocated metal and why its electrical resistivity may be approximately isotropic. Journal of Physics F: Metal
Physics, 17(8), 1703.
Watts, B. R. (1988a). The contribution of the long-range strain field of dislocations in metals to their electrical resistivity. Journal of Physics F: Metal Physics,
18(6), 1183.
Watts, B. R. (1988b). Calculation of electrical resistivity produced by dislocations in various metals. Journal of Physics F: Metal Physics, 18(6), 1197.

You might also like