Abstract

Plastic deformation in metals is dominated by the interactions among dislocations and other defects inside the crystal. A large number of dislocation multipoles (dipoles, tripoles, quadrupoles, etc.) can form during plastic deformation. Depending on the relative position and the orientation of the dislocations, interactions in and between multipoles can change the elastoplastic properties of a material. The authors of this article investigate the effect of dislocation multipoles on the elastoplastic properties of a material. This is performed analytically under different multipole configurations (i.e., the distance between active glide planes and the signs of the dislocations) as well as using a three-dimensional discrete dislocation dynamics (DDD) code. The simulations show that multipoles exhibit a hardening/softening effect when the sign of the dislocations involved is the same, and a hardening effect only when the dislocations are of opposite sign to nearby ones. The distance between the two neighboring dislocations was also affecting the proportional limit (PL) for the material. Such hardening or flow stress (FS) results, as in this study, can be incorporated into larger-scale modeling work.

Introduction

Dislocations are a type of defect in crystalline materials. They are labeled as “line defects.” In a typical crystal, many dislocations can exist and the dislocation density and structure will have an effect on the physical properties of the material. When two or more dislocations interact closely from same/different slip planes, they are labeled as “multipoles.” Multipoles can be of edge or screw type. Depending on the proximity of the dislocations in the multipoles, interactions among the dislocations can affect the elastoplastic properties [1]. A material can be hardened or softened as a consequence of these interactions.

Stokes and Olsen [2] and Kroupa [3] described the mechanism of dipole formation for both screw and edge dislocations. Gilman [1] and Neumann [4] studied the interaction between dislocations and dislocation dipoles, tripoles, and quadrupoles. In the two-dimensional study by Neumann [4], it was found that decomposition, i.e., splitting of the dipole usually occurs more often than trapping of the approaching dislocations.

To understand dislocation phenomena in the plastic regime, researchers and scientists developed several codes that can simulate the dynamic behavior of dislocation interactions under simple to complex configurations, which are often not captured in real experiments. Kubin et al. [5] introduced a basic framework for three-dimensional (3D) dislocation dynamics simulations and various features were introduced in more recent dislocation dynamics codes. In the current study, the simulations are performed on a dislocation dynamics code developed by Refs. [614]. The first work to look at dipole interactions using discrete dislocation dynamics (DDD) simulations is by Ref. [15]. In such work, it showed dynamic zipping and unzipping, i.e., dynamic coupling/de-coupling of dipole dislocations. Such dipole sources were initially remote from one another unlike the initial equilibrium stacking studied herein.

The nature (stable/unstable) of the equilibrium points in a multipole configuration plays a vital role in the elastoplastic behavior. The authors of this article present an analytical/theoretical model for stability analysis for any number of dislocations in a multipole configuration. This is something that was lacking from the literature.

All the analytical/theoretical model cases presented in this article consider the dislocations as infinitely long. Although such studies can provide an understanding of multipole stability, the infinite length of the dislocation preclude the line tension effect, and the interplay between line tension and remote interactions, on elastoplasticity. To address this lacking knowledge, we therefore also show DDD simulation results of staggered Frank–Read (FR) sources [16] (pinned dislocation segments on different slip plane) in a 3D finite volume. These represent multipoles in real-world problems since both the medium and the dislocations are finite. Such simulations capture both the line tension and the interaction effects on a material's elastoplasticity.

Detailed development of the interaction equations among the dislocations in a multipole configuration is shown in the Theory section, and the stability analysis is shown in the Stability of Multipoles section. The simulation setup is discussed in the Simulation Methodology section for the reproducibility of the results. A comprehensive discussion of the simulation results is presented in the Results and Discussion section.

Theory

The thermodynamic force F (also known as the Peach–Koehler (PK) force) acting on a dislocation can be calculated by [17]
F=(σb)×e^
(1)
where σ is the stress tensor at a point on a dislocation line due to internal and externally applied stresses, b is the magnitude and direction of the crystal distortion associated with that dislocation (also known as the Burgers vector) and e^ is a unit vector along the dislocation line (i.e., the line sense vector). Dislocations move in a slip plane when the PK force is high enough to overcome the internal friction in the crystal. A dislocation source multiplies when this force is large enough to overcome the line tension of the dislocation. Line tension T of a dislocation is given by [18]
T=αGb2
(2)
where α ≈ 0.5 − 1.0 and b is the magnitude of the Burgers vector. In the multiplication of an initially straight Frank–Read source, it is assumed that plastic flow occurs when the dislocation bows a half-circle [18]. This is also known as critical bowing. For critical bowing, the critical shear stress is calculated as
τcrit=αGbR
(3)
where R is the radius of the bowed dislocation. A similar derivation for the line tension and τcrit is given in Ref. [19].

Dipole

Consider two infinite (along z) edge dislocations laying on slip planes parallel to the xz plane and separated by a distance d, as shown in Figs. 1 and 2. The glide force, Fx (PK force) acting on either dislocation due to their interaction when they both have the same sign (Fig. 1) can be derived by inserting b=(bx,0,0) and e^=(0,0,1) in Eq. (1)
Fx=σxybx
(4a)
Fx=Gbx22π(1ν)x(x2d2)(x2+d2)2
(4b)
Fxκ=(xd)[(xd)21][(xd)2+1]2
(4c)
where κ=Gbx2/(2πd(1ν)), G and ν are the shear modulus and the Poisson's ratio of the material. Equation 4(b) was also provided in Refs. [18,20,21].
Fig. 1
Dipole stable equilibrium (same-sign dislocations)
Fig. 1
Dipole stable equilibrium (same-sign dislocations)
Close modal
Fig. 2
Dipole stable equilibrium (opposite sign dislocations)
Fig. 2
Dipole stable equilibrium (opposite sign dislocations)
Close modal
When both dislocations have an opposite sign as in Fig. 2 (e.g., same Burgers vector but opposite line sense) one can insert e^=(0,0,1) in Eq. (1) and then write
Fxκ=(xd)[(xd)21][(xd)2+1]2
(5)

Tripole

Tripoles can be formed in different configurations. Figure 3 shows two tripole arrangements where the dislocations lie on different slip planes.

Fig. 3
Tripole configurations: (a) same sign and (b) opposite sign in zigzag pattern
Fig. 3
Tripole configurations: (a) same sign and (b) opposite sign in zigzag pattern
Close modal
When all dislocations have the same sign (Fig. 3(a)), the glide force Fx at the top (or bottom) dislocation due to the interaction among the dislocations can be expressed as
Fxκ=(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)24][(xd)2+4]2
(6)
And for the dislocation in the middle
Fxκ=2(xd)[(xd)21][(xd)2+1]2
(7)
When the tripole dislocations are patterned in a zigzag (Fig. 3(b)) and are alternating in sign, the resulting glide force Fx at the top (or bottom) dislocation due to dislocation interactions is given by
Fxκ=(xd1)[(xd1)21][(xd1)2+1]2+(xd)[(xd)24][(xd)2+4]2
(8)
And for the dislocation in the middle
Fxκ=2(xd)[(xd)21][(xd)2+1]2
(9)

General Multipole

In general, when all the dislocations have the same sign and are all lined up vertically, the glide force Fx on the top (or bottom) dislocation resulting from the interaction among the dislocations is given by
Fxκ=i=1n1(xd)[(xd)2i2][(xd)2+i2]2
(10)
And for the dislocation in the middle when the number of dislocations in the multipole n is an odd number
Fxκ=2i=1(n1)/2(xd)[(xd)2i2][(xd)2+i2]2
(11)
when n is even (middle is dislocation number n/2 + 1 counting from the bottom dislocation)
Fxκ=(xd)[(xd)2(n2)2][(xd)2+(n2)2]2+2i=1n/21(xd)[(xd)2i2][(xd)2+i2]2
(12)
As for the top (or bottom) dislocation when all the dislocations in the multipole have a zigzag pattern (similar to Fig. 3(b)) and alternating in sign, the force is given by
Fxκ=i=1n1(1)i(xdξi)[(xdξi)2i2][(xdξi)2+i2]2
(13)
where ξi = 1/2 + 1/2( − 1)ni.
And for the dislocation in the middle when n is an odd number
Fxκ=2i=1(n1)/2(1)i(xdξi)[(xdξi)2i2][(xdξi)2+i2]2
(14)
where ξi = 1/2 − 1/2(−1)(n−1)/2−i.
When n is even and for the middle dislocation (middle is dislocation number n/2 + 1 counting from the bottom dislocation)
Fxκ=(1)n/2(xd)[(xd)2(n2)2][(xd)2+(n2)2]2+2i=1n/21(1)i(xdξi)[(xdξi)2i2][(xdξi)2+i2]2
(15)
where ξi = 1/2 − 1/2(−1)n/2−i.

For a quadrupole and pentapole configurations, these are provided in the  Appendix (the configuration, the glide force, and the glide force versus x).

Stability of Multipoles

Figure 4 shows the glide forces Fx for dipoles, as a function of x, and the equilibrium points (where Fx = 0) [18,19]. For a dipole having same-sign dislocations (Fig. 1), there exists two unstable equilibrium points (at x = ±d) and one stable (middle) equilibrium point at the origin (at x = 0). The two unstable equilibrium points occur when the relative angle between the two dislocations is 45deg, and the stable equilibrium point occurs when the relative angle between the two dislocations is 0deg (i.e., they are lined up vertically).

Fig. 4
Normalized glide force between two parallel infinite edge dislocations of same and opposite signs
Fig. 4
Normalized glide force between two parallel infinite edge dislocations of same and opposite signs
Close modal

For the dipole with two oppositely signed dislocations (Fig. 2), the dipole has two stable equilibrium points (at x = ±d) and one unstable (middle) equilibrium point at the origin (at x = 0). The two stable equilibrium points occur when the relative angle between the two dislocations is 45deg and the unstable equilibrium point occurs when the relative angle between two dislocations is 0deg (i.e., they are lined up vertically).

For any perturbation from the equilibrium points, the glide force is directed toward the stable equilibrium points and away from the unstable equilibrium points. In other words, when any dislocation in a dipole configuration moves a little away from the stable equilibrium position, it feels a force that tends to bring it back to that stable equilibrium position. For an unstable equilibrium point, the opposite phenomenon occurs.

The region in space where a perturbation of the dislocation position tends to return it to an equilibrium point is termed a stable region, whereas a region in space where a perturbation of the dislocation position tends to move it away from an equilibrium point is termed an unstable region. The same-sign dipole has stable and unstable regions (the stable region is |x| < d and the unstable regions are |x| > d) but the opposite sign dipole has a stable region only (the stable region is − < x < +). Therefore, the opposite sign dipole should be resistant to dislocation flow on an extended range whereas the same-sign dipole is only resistant to flow on a shorter range, and after that, it would not resist but rather assist in the dislocation flow.

Figure 5 shows the glide forces Fx for tripoles, as a function of x, and the equilibrium points (where Fx = 0). For a tripole having same-sign dislocations (Fig. 3(a)), there exists two unstable equilibrium points and one stable (middle) equilibrium point at the origin. This is true for both the top (or bottom) dislocation and the middle dislocation although both have a relatively smaller region of stability. For Fig. 3(b), there are also three equilibrium points with two being stable but with an infinite stability region. This is true for both the top/bottom and middle dislocations. Note that the force/stability curves for the tripole are more complex than for the dipole. For one reason, there are more of them to consider. Also for the tripole (or higher-order poles), the curves are not always anti-symmetric which requires careful identification of peak or absolute peak values as shown in the  Appendix.

Fig. 5
Normalized glide force for the tripole configurations in Fig. 3. “Same” refers to all dislocations in the multipole being of the same sign. “Opp.” Refers to a zigzag pattern of the multipole where the dislocations alternate in sign. “Top” refers to the top dislocation and “Mid” refers to the middle dislocation.
Fig. 5
Normalized glide force for the tripole configurations in Fig. 3. “Same” refers to all dislocations in the multipole being of the same sign. “Opp.” Refers to a zigzag pattern of the multipole where the dislocations alternate in sign. “Top” refers to the top dislocation and “Mid” refers to the middle dislocation.
Close modal

From the stability figures (Fig. 5 and the  Appendix), it is evident that the top or outer dislocation in the multipole configuration is weakly coupled with other dislocations in the multipole and therefore subject to flowing first, i.e., detaching away first from the multipole. In other words, it is the weakest link in the multipole. The middle dislocation on the other hand seems to be more stable. With focus thus on the top/outer dislocation, Figs. 6 and 7 show surface plots based on Eqs. (10) and (13). Each surface point shows the PK force on the top dislocation of the multipole configurations. The force is a function of n and x. The figures show the complex dependence on n and x.

Fig. 6
Surface plot of the top dislocation when all dislocations in the multipole are vertically stacked and of the same sign (Eq. (10))
Fig. 6
Surface plot of the top dislocation when all dislocations in the multipole are vertically stacked and of the same sign (Eq. (10))
Close modal
Fig. 7
Surface plot of the top dislocation when all dislocations in the multipole are arranged in a vertical zigzag with their sign alternating between plus and minus (Eq. (13))
Fig. 7
Surface plot of the top dislocation when all dislocations in the multipole are arranged in a vertical zigzag with their sign alternating between plus and minus (Eq. (13))
Close modal

Figure 8 shows the maximum absolute PK force values acting on the top and middle dislocation. These maximum points are evaluated numerically, and each point shows the absolute maximum force irrespective of the x-location of this maximum point on the force curve. To get a curve for a middle dislocation at least three dislocation are required in a multipole. As seen in this figure, all the curves seem to plateau to some limiting value because the dislocation interaction has an inverse relationship to the separation distance between dislocations in the multipole configuration, i.e., the distances between the top dislocation and other dislocations will increase as n increases. Note in this figure that only when n = 2 or n = 3 that the absolute value of the glide force is the same whether one considers a same-sign multipole configuration (all vertically aligned) or a vertical zigzag configuration with alternating dislocation signs. For any other n, the glide forces are not exactly the same. This figure shows that the middle dislocation could be twice as stable (factor of 2) as the top dislocation with this factor depending on n.

Fig. 8
Absolute maximum force for each value of n
Fig. 8
Absolute maximum force for each value of n
Close modal

Simulation Methodology

To investigate the dislocation interactions mentioned in the Theory section in a three-dimensional finite-sized space, a number of simulations are performed by placing two, three, four, and five FR sources in different parallel slip planes (see Fig. 9 as an example) and applying constant strain rate using the code developed by Refs. [614]. A main part of the simulations is to calculate the PK force experienced, at any time-step, by the different dislocation segments in the computational box (Eq. (1)) and then applying a mobility law:
vg=MgFg
(16)
where Mg is the glide mobility of the dislocations and Fg is the glide component of force F in Eq. (1). The study aims to understand how the initial multipole configuration affects the elastoplastic responses in the presence of external loading on the crystal. The shear modulus for the selected material (aluminum) is taken at 26.32 GPa and Poisson's ratio is taken as 0.33. The dimension of the representative volume element (RVE) for these simulations is 60,000b × 60,000b × 40,000b. The Burgers vector of the Frank–Read sources is taken as [0 1 0]. The shear load is applied in the yz-direction, and the applied strain rate is 10 s−1. Dislocations lie parallel to the x-axis with a separation distance d. The length of each dislocation is 4000b.
Fig. 9
Same-sign pentapole configuration: (a) five Frank–Read sources inside an RVE and (b) front view of the RVE
Fig. 9
Same-sign pentapole configuration: (a) five Frank–Read sources inside an RVE and (b) front view of the RVE
Close modal

The first group of simulations is conducted by changing the separation distance between active slip planes for the same-signed dislocations in the multipole (dipole, tripole, quadrupole, and pentapole) configurations. Initially, all FR sources are placed on different slip planes as they make 0deg angle with each other with respect to the vertical axis. This is a stable position for such multipoles, as mentioned in the Stability of Multipoles section.

Then in the second group of simulations, the dislocation sign was alternated between neighboring dislocations while the dislocations themselves are arranged in a vertical zigzag pattern. This is done for different multipoles (dipole, triple, quadrupole, and pentapole). To reverse the dislocation sign one can take opposite Burgers vectors for the neighboring dislocations, i.e., [0 1 0] and [01¯0]. Here also, the separation between the dislocations on the different slip planes is varied and the result of this is obtained and studied. In other words, for the zigzag pattern each dislocation is making a 45deg angle with respect to the vertical axis with the neighboring dislocations. See Fig. 3(b) and the quadrupole and pentapole figures in the  Appendix.

At the start of each simulation, i.e., at time zero, the glide force Fy experienced by any dislocation, or dislocation segment, in the above multipole configuration can be obtained using Eq. (1). For this simulation setup, Fy = σyzb since e^=[100]. Of course after the bowing of the FR sources commence, each dislocation segment can experience both glide components of the PK force, i.e., Fx and Fy.

Results and Discussion

Stress–strain diagrams are generated for each 3D DDD simulation. Although these diagrams were generated for different dipole, tripole, quadrupole, and pentapole configurations, they are only shown here for the pentapole configurations for brevity and as exemplary figures of the results of the DDD simulations. Figures 10 and 11 show the stress–strain responses of the RVE for different slip plane separations (normalized d) for pentapole configurations with the same and oppositely alternating dislocation signs, respectively. All the strain–stress diagrams are compared with a single FR source (labeled as “One FR”) to capture any occurring hardening or softening phenomena associated with the different multipole configurations. The proportional limit (PL) (stress value at the end point of the elastic region) and the flow stress (FS) (averaged stress values over strain after any initial drop in stress beyond the elastic regime) are recorded for each simulation. Note that in Fig. 10, some stress–strain diagrams show a clear drop in stress value right after the elastic regime. This sudden drop is similar to the Yield Point Phenomenon (YPP) observed in low carbon steels, although the initial locking of dislocations here is due to the other dislocations in the multipole, unlike YPP in steels which is due to the solute carbon atoms.

Fig. 10
Stress–strain diagram for the material under different pentapole separations when all the dislocations have the same sign
Fig. 10
Stress–strain diagram for the material under different pentapole separations when all the dislocations have the same sign
Close modal
Fig. 11
Stress–strain diagram for the material under different pentapole separations when the dislocation signs alternate between neighboring dislocations/sources
Fig. 11
Stress–strain diagram for the material under different pentapole separations when the dislocation signs alternate between neighboring dislocations/sources
Close modal

Figures 1215 show the proportional limit and flow stress for different multipoles as the separation distance d between any two slip planes is varied. The minimum d/b considered is 50. All the diagrams show both hardening and softening effects from the elastic interaction between two or more FR sources having the same sign.

Fig. 12
Comparison of proportional limit and flow stress of a dipole with a single Frank–Read source when both dislocations have the same sign
Fig. 12
Comparison of proportional limit and flow stress of a dipole with a single Frank–Read source when both dislocations have the same sign
Close modal
Fig. 13
Comparison of proportional limit and flow stress of a tripole with a single Frank–Read source when the dislocations have the same sign
Fig. 13
Comparison of proportional limit and flow stress of a tripole with a single Frank–Read source when the dislocations have the same sign
Close modal
Fig. 14
Comparison of proportional limit and flow stress of a quadrupole with a single Frank–Read source when the dislocations have the same sign
Fig. 14
Comparison of proportional limit and flow stress of a quadrupole with a single Frank–Read source when the dislocations have the same sign
Close modal
Fig. 15
Comparison of proportional limit and flow stress of a pentapole with a single Frank–Read source when the dislocations have the same sign
Fig. 15
Comparison of proportional limit and flow stress of a pentapole with a single Frank–Read source when the dislocations have the same sign
Close modal

This result shows good agreement with the information in the Theory/Stability of Multipoles sections regarding small source separations. For small separations, the sources bow together, trying to hold each other in equilibrium as they bow (Fig. 16). This raises the proportional limit (i.e., the critical bowing of source(s)) and thus the flow stress above a one source. However, as the distance between the sources increases more and more, the interaction effect of the sources on one another become less and less during bowing. This results in the proportional limit and flow stress eventually dropping to the same level as that of one source. For this kind of interaction, two parameters play an essential role in conjunction with one another: one is the separation distance between any two slip planes, and the other is the PK force due to the self-stress field created by the dislocations, i.e., the line tension of a dislocation. This force, with perturbations in the self-stress of the dislocations, can cause the same-signed dislocation sources to either take longer to critically bow (i.e., a hardening or strengthening effect) or bow critically quickly, i.e., at lowered strain value (i.e., a softening effect as seen in a stress–strain diagram). The reason for this softening effect at very small d values, is that the interaction force is very high and any perturbations in the bowing between the sources can at one point flip a stable position between any two nearby dislocations to an unstable position pushing the dislocations away from each other instead of pulling them to stay together. That will accelerate the production of strain at the same applied load value which is the exact definition of softening.

Fig. 16
A snapshot of a bowing quadrupole with four FR sources. The outer sources are breaking away first from the pack.
Fig. 16
A snapshot of a bowing quadrupole with four FR sources. The outer sources are breaking away first from the pack.
Close modal

Another result from the simulations is that for any number n ≥ 3, the DDD simulations show that the outer FR sources bow out critically first before the middle ones. Figure 16 shows a snapshot of a quadruple were the outer sources are starting to take off, i.e., break away from the pack, before the middle ones. This is in line with the stability discussions above regarding infinite-length dislocations that the outer dislocations are the least stable. This figure also shows how all the dislocations in a same-sign multipole initially bend/bow together while trying to maintain their vertical originally stable configuration. This continuous attempt to align the dislocations in a vertical configuration produces resistance to applied loading and hence a hardening effect that reflects itself on the PL results in Figs. 1215.

Figures 1720 show different results for the oppositely-signed multipole configurations (in a zigzag pattern) compared with the same-signed multipoles having the same number n (Figs. 1215). Here also, the minimum d/b considered is 50. For smaller separation distances in the oppositely-signed multipole configurations, both the proportional limit and the flow stress are significantly higher than for a single FR source. It can be as high as a factor of 3. For these multipole configurations, the entire slip plane acts as a basin of attraction for all the Frank–Read sources back to the equilibrium points, as shown in Figs. 4 and 5 (see the  Appendix for higher-order multipoles). Thus one should not expect any softening effect from this configuration because this continuing stability of the configuration and continuous attraction to the equilibrium position hinders plastic flow. Figures 1720 also show a good agreement with the study hypothesis: as the separation distance increases, the stress–strain response mimics that of a single Frank–Read source response. This is true in the limit of high d values.

Fig. 17
Comparison of proportional limit and flow stress of a dipole with a single Frank–Read source when both dislocations have opposite signs
Fig. 17
Comparison of proportional limit and flow stress of a dipole with a single Frank–Read source when both dislocations have opposite signs
Close modal
Fig. 18
Comparison of proportional limit and flow stress of a tripole with a single Frank–Read source when the dislocations have alternating opposite signs
Fig. 18
Comparison of proportional limit and flow stress of a tripole with a single Frank–Read source when the dislocations have alternating opposite signs
Close modal
Fig. 19
Comparison of proportional limit and flow stress of a quadrupole with a single Frank–Read source when the dislocations have alternating opposite signs
Fig. 19
Comparison of proportional limit and flow stress of a quadrupole with a single Frank–Read source when the dislocations have alternating opposite signs
Close modal
Fig. 20
Comparison of proportional limit and flow stress of a pentapole with a single Frank–Read source when the dislocations have alternating opposite signs
Fig. 20
Comparison of proportional limit and flow stress of a pentapole with a single Frank–Read source when the dislocations have alternating opposite signs
Close modal

It is important to note that in such simulations the ensuing dislocation microstructures can be complex as they evolve with strain. This complexity is a function of the source vertical separation (d/b) and the number of sources (n). To give one example of this complexity, consider Fig. 21. In this figure, a pentapole is shown at one strain value and as can be seen from the differently colored (or grey-shaded in the print version) dislocations/sources, their interaction is complex. This snapshot shows, for example, the formation of multiple dynamic dipoles at that instant of time. All such dynamics dipoles represent temporary pinning of dislocation motion/glide and thus cause an increase to the applied stress value needed to maintain plastic flow.

Fig. 21
A snapshot in time of a pentapole (of the zigzag type) showing the interaction between the different active sources. Here, d/b = 200. The arrows are pointing toward dynamic dipoles.
Fig. 21
A snapshot in time of a pentapole (of the zigzag type) showing the interaction between the different active sources. Here, d/b = 200. The arrows are pointing toward dynamic dipoles.
Close modal

The peak (maximum) and minimum values for the PL and FS from Figs. 1215, independent of separation distance d, are plotted in Fig. 22. Figure 22 shows how added dislocations in the multipole contribute to dislocation interactions that results in increased PL/FS with n. Given the prior discussion on the stabilizing effect of dislocations in a multipole, this result is not surprising. The curves in Fig. 22 seem to be reaching a plateau similar to what is seen in Fig. 6.

Fig. 22
Comparison of proportional limit and flow stress of multipoles with a single FR source (last legend) when the dislocations all have the same sign. All curves are extrapolated to n = 1 (i.e., single FR source).
Fig. 22
Comparison of proportional limit and flow stress of multipoles with a single FR source (last legend) when the dislocations all have the same sign. All curves are extrapolated to n = 1 (i.e., single FR source).
Close modal

There are a few things to note about this figure. First, the dislocation/source interactions are contributing to higher FS and PL than a single FR source (last legend in the figure). Indeed, at the highest n (i.e., n = 5), this increase from a single FR source is about 55% or a factor of 1.52 higher to critically bow a pentapole than to critically bow a single FR. A higher factor is possible if n is higher. Note that this increase in PL is due to dislocation interaction between sources and not due to line tension (which is driving the base value for a single FR source). Hence, dislocation interactions are shown here as clearly significant for the elastoplasticity of a single crystal.

Another thing to note about Fig. 22 is that if one divides the increase in PL value from n = 2 to n = 5, the factor here will be 1.33 (16/12). However, from Fig. 8 for the top dislocation (the least stable), the ratio of the breakaway force for n = 5 compared with n = 2 is 1.76 (0.44/0.25). These values indicate the effect of dislocation interaction going from n = 2 to n = 5 and not the effect of line tension. Moreover, the reason the ratio from Fig. 8 is higher is that it considers infinite dislocations. Therefore, it appears that the theoretical analysis done in the Stability of Multipoles section is relevant to the analysis of finite-length sources arranged similarly as multipoles.

The peak (maximum) and minimum values for the PL and FS from Figs. 1720, independent of separation distance d, are plotted in Fig. 23. Figure 23 shows how added dislocations in the multipole contribute to dislocation interactions boosting the PL/FS over a single FR source. However, unlike Fig. 22 for dislocations of the same sign, Fig. 23 for dislocations of alternating sign shows an oscillating behavior. These oscillations are similar to what is seen in Fig. 7 and should ultimately reach a limiting value with higher n. Notice that the PL and FS maximum values in Fig. 23 are higher than those in Fig. 22. The reason for this was discussed earlier. Note also that the minimum PL/FS values in this figure occur for large source separation d.

Fig. 23
Comparison of proportional limit and flow stress of multipoles with a single FR source (last legend) when the dislocations are zigzag-arranged and alternating in sign
Fig. 23
Comparison of proportional limit and flow stress of multipoles with a single FR source (last legend) when the dislocations are zigzag-arranged and alternating in sign
Close modal

An important outcome of these studies is that there were significant changes in the elastoplastic properties of a crystal not from a change in the dislocation density of that crystal but rather from changes in the configuration and relative position of the dislocations in the crystal. This is an important outcome of this work as dislocation density is a major parameter used in continuum plasticity models.

Limitations

The current study is limited to the hardening/softening of a crystalline material only due to the dislocation interactions in a multipole. Here dislocations are kept equally apart from each other in both same and opposite type multipole configurations. Moreover, the other aspects of material hardening processes (i.e., solid solution hardening and precipitation hardening) are not considered because the objective of this study is to understand the dislocation interactions in a multipole and its hardening or softening effect on the crystalline materials. In the analytical study, dislocations are considered to be infinitely long, so the line tension is neglected. Though the DDD simulations are done with finite dislocation sources considering line tension, the analytical prediction still reflects the results.

Again the unequal separation between dislocations in the multipole may somewhat alter the trend of the current results. Simulations are done for a certain value of mobility constant, and the effect of dislocation mobility is not studied in this research.

Conclusions

In this article, the authors presented a study on dislocation multipoles, and presented analytical stability diagrams/equations and 3D DDD simulation outcomes. The simulations incorporated line tension effects whereas the analytical models did not since they assume infinite length of dislocations. Although the simulations used finite-length dislocation sources, the analytical models were still helpful in understanding the simulations. The simulations showed that both hardening and softening behaviors in the material can occur. One of the main findings of this study is that zigzagged multipoles with alternating dislocation sign only cause hardening. On the contrary, same-sign multipoles can exhibit either hardening or softening behavior depending on the separation distance between slip planes. At a separation distance of 4500b, the sources act the same as a single FR source, i.e., the interaction between them diminishes. Another important finding is that the elastoplastic properties of a crystal are dependent not just on dislocation density but also on the configuration and relative position of the dislocations in the crystal. Yet another finding is that dislocation interactions significantly contribute to the proportional limit and flow stress of a crystal and not just the line tension of the dislocations. In future studies, one could consider the initial source length as a parameter in the multipole configurations.

Acknowledgment

This paper describes objective technical results and analysis. Any subjective views or opinions that might be expressed in the paper do not necessarily represent the views of the U.S. Department of Energy or the United States Government. Sandia National Laboratories is a multi-mission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC., a wholly owned subsidiary of Honeywell International, Inc., for the U.S. Department of Energy's National Nuclear Security Administration under contract DE-NA0003525.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The authors attest that all data for this study are included in the paper.

Appendix

Quadrupole.

Same-sign dislocations: For top dislocation
Fxκ=(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)24][(xd)2+4]2+(xd)[(xd)29][(xd)2+9]2
Same-sign dislocations: For middle dislocation (third from bottom). The number of the middle dislocation in a multipole, counting from 1 at the bottom dislocation, is given by: (n − 1)/2 + 1 if n is odd and n/2 + 1 if n is even
Fxκ=(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)24][(xd)2+4]2
Opposite sign dislocations: For top dislocation
Fxκ=(xd)[(xd)21][(xd)2+1]2+(xd1)[(xd1)24][(xd1)2+4]2(xd)[(xd)29][(xd)2+9]2
Opposite sign dislocations: For middle dislocation (third from bottom)
Fxκ=(xd1)[(xd1)21][(xd1)2+1]2(xd1)[(xd1)21][(xd1)2+1]2+(xd)[(xd)24][(xd)2+4]2

Pentapole.

Same-sign dislocations: For top dislocation
Fxκ=(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)24][(xd)2+4]2+(xd)[(xd)29][(xd)2+9]2+(xd)[(xd)216][(xd)2+16]2
Same-sign dislocations: For middle dislocation (third dislocation from bottom)
Fxκ=(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)21][(xd)2+1]2+(xd)[(xd)24][(xd)2+4]2+(xd)[(xd)24][(xd)2+4]2
Opposite sign dislocations: For top dislocation
Fxκ=(xd1)[(xd1)21][(xd1)2+1]2+(xd)[(xd)24][(xd)2+4]2(xd1)[(xd1)29][(xd1)2+9]2+(xd)[(xd)216][(xd)2+16]2

Opposite burgers vector: For middle dislocation (third dislocation from bottom)

Fxκ=(xd)[(xd)24][(xd)2+4]2(xd1)[(xd1)21][(xd1)2+1]2(xd1)[(xd1)21][(xd1)2+1]2+(xd)[(xd)24][(xd)2+4]2
Fig. 24
Quadrupole configurations: (a) same-sign dislocations and (b) opposite sign dislocations
Fig. 24
Quadrupole configurations: (a) same-sign dislocations and (b) opposite sign dislocations
Close modal
Fig. 25
Normalized glide force for the quadrupole configurations
Fig. 25
Normalized glide force for the quadrupole configurations
Close modal
Fig. 26
Pentapole configurations: (a) same-sign dislocations and (b) opposite sign dislocations
Fig. 26
Pentapole configurations: (a) same-sign dislocations and (b) opposite sign dislocations
Close modal
Fig. 27
Normalized glide force for the pentapole configurations
Fig. 27
Normalized glide force for the pentapole configurations
Close modal

References

1.
Gilman
,
J. J.
,
1964
, “
Influence of Dislocation Dipoles on Physical Properties
,”
R. Soc. Chem.
,
38
, pp.
123
137
.
2.
Stokes
,
R. J.
, and
Olsen
,
K. H.
,
1963
, “
Dislocation Interactions and Dipole Formation
,”
Philos. Mag.
,
8
(
90
), pp.
957
965
.
3.
Kroupa
,
F.
,
1966
, “
Dislocation Dipoles and Dislocation Loops
,”
Le Journal de Physique Colloques
,
27
(
C3
), pp.
C3-154
C3-167
.
4.
Neumann
,
P. D.
,
1971
, “
The Interactions Between Dislocations and Dislocation Dipoles
,”
Acta Metall.
,
19
(
11
), pp.
1233
1241
.
5.
Kubin
,
L. P.
,
Canova
,
G.
,
Condat
,
M.
,
Devincre
,
B.
,
Pontikis
,
V.
, and
Brechet
,
Y.
,
1992
, “
Dislocation Microstructures and Plastic Flow: A 3D Simulation
,”
Solid State Phenom.
,
23–24
, pp.
455
472
.
6.
Zbib
,
H. M.
,
Rhee
,
M.
, and
Hirth
,
J. P.
,
1998
, “
On Plastic Deformation and the Dynamics of 3D Dislocations
,”
Int. J. Mech. Sci.
,
40
(
2–3
), pp.
113
127
.
7.
De La Rubia
,
T. D.
,
Zbib
,
H. M.
,
Khraishi
,
T. A.
,
Wirth
,
B. D.
,
Victoria
,
M.
, and
Caturla
,
M. J.
,
2000
, “
Multiscale Modelling of Plastic Flow Localization in Irradiated Materials
,”
Nature
,
406
(
6798
), pp.
871
874
.
8.
Khraishi
,
T. A.
,
Zbib
,
H. M.
,
de La Rubia
,
T. D.
, and
Victoria
,
M.
,
2002
, “
Localized Deformation and Hardening in Irradiated Metals: Three-Dimensional Discrete Dislocation Dynamics Simulations
,”
Metall. Mater. Trans. B
,
33
(
2
), pp.
285
296
.
9.
Yan
,
L.
,
Khraishi
,
T. A.
,
Shen
,
Y.-L.
, and
Horstemeyer
,
M. F.
,
2004
, “
A Distributed-Dislocation Method for Treating Free-Surface Image Stresses in 3D Dislocation Dynamics Simulations
,”
Modell. Simul. Mater. Sci. Eng.
,
12
(
4
), pp.
S289
S301
.
10.
Jing
,
P.
,
Khraishi
,
T.
,
Young
,
J. A.
, and
Wirth
,
B. D.
,
2005
, “
Multi-Scale Simulations of the Effects of Irradiation-Induced Voids and Helium Bubbles on the Mechanical Properties of Aluminum
,”
Philos. Mag.
,
85
(
4–7
), pp.
757
767
.
11.
Khraishi
,
T. A.
, and
Zbib
,
H. M.
,
2002
, “
Dislocation Dynamics Simulations of the Interaction Between a Short Rigid Fiber and a Glide Circular Dislocation Pile-up
,”
Comput. Mater. Sci.
,
24
(
3
), pp.
294
306
.
12.
Siddique
,
A. B.
, and
Khraishi
,
T.
,
2020
, “
Numerical Methodology for Treating Static and Dynamic Dislocation Problems Near a Free Surface
,”
J. Phys. Commun.
,
4
(
5
), p.
055005
.
13.
Siddique
,
A. B.
, and
Khraishi
,
T.
,
2021
, “
A Mesh-Independent Brute-Force Approach for Traction-Free Corrections in Dislocation Problems
,”
Model. Numer. Simul. Mater. Sci.
,
11
(
1
), pp.
1
18
.
14.
Siddique
,
A. B.
, and
Khraishi
,
T.
,
2021
, “
The Treatment of Singularities Associated With a Dislocation Segment With Applications
,”
Int. J. Theor. Appl. Multiscale Mech.
(in press).
15.
Rhee
,
M.
,
Zbib
,
H. M.
,
Hirth
,
J. P.
,
Huang
,
H.
, and
Rubia
,
T. D.
,
1998
, “
Models for Long-/Short-Range Interactions and Cross Slip in 3D Dislocation Simulation of BCC Single Crystals
,”
Modelling Simul. Mater. Sci. Eng.
,
6
(
4
), pp.
467
492
.
16.
Frank
,
F. C.
, and
Read
,
W. T.
,
1950
, “
Multiplication Processes for Slow Moving Dislocations
,”
Phys. Rev.
,
79
(
4
), pp.
722
273
.
17.
Hirth
,
J. P.
, and
Lothe
,
J.
,
1982
,
Theory of Dislocations
,
John Wiley & Sons
,
New York
.
18.
Hull
,
D.
, and
Bacon
,
D. J.
,
2011
,
Introduction to Dislocations
,
Butterworth-Heinemann
,
New York
.
19.
Weertman
,
J.
, and
Weertman
,
J. R.
,
1992
,
Elementary Dislocation Theory
,
Oxford University Press
,
Oxford
.
20.
Cottrell
,
A. H.
,
1953
,
Dislocations and Plastic Flow in Crystals
,
Oxford University Press
,
Oxford
, p.
48
.
21.
Dieter
,
G. E.
, and
Bacon
,
D. J.
,
1976
,
Mechanical Metallurgy
, 3rd ed.,
McGraw-Hill, New York
. p.
167
.