Vous êtes sur la page 1sur 13

STRUCTURAL KINETIC ASPECTS OF THE PHYSICS OF EVOLUTION OF PLASTIC DEFORMATION V. V. Rybin UDC 539.4.

035

EVOLUTION OF PLASTIC DEFORMATION Classical dislocation physics of rigidity and plasticity has a quite narrow range of validity on the scale of deformations. It describes the mechanical behavior of crystalline solids only near the limit of fluidity (Fig. I), and sharply loses its prognostic abilities with increasing extent of plastic deformation ~. Starting with g0 ~ 0.1-0.2,* the regular dislocation theories make it possible to provide only general qualitative estimates. The reasons are manifested in two fundamental premises on which they are based. Firstly, the plastic deformation process is usually treated as a trivial consequence of the motion of a set of, within the first approximation, individual noninteracting dislocations, which are displaced through the crystal in the field of stresses ~ext created by the distribution of external loads (external stresses). The velocity tensor of plastic deformations s is expressed in terms of the velocity V and the density of mobile individual dislocations p by an equation of the form

=--~

~ p=l

(pV)p(rib + bn)p,

(1)

where n and b are the normal to the glide plane and the dislocation Burgers vector, and p is a subscript running over the glide system. Secondly, it is assumed that the structure remains qualitatively unchanged (i.e., the original) during the deformation. If its evolution is treated as well, this can be done quantitatively due to the increase in dislocation density, the formation of heterogeneities in their distributions in glide planes, and the formation of loops, dipoles, and multipole configurations. These assumptions are verified only for small deformations. Precisely therefore the classical dislocation theories describe well the value of the fluidity limit, the features of brittle destruction, the fine and crude geometry of gliding band, construction of the simplest dislocation formations, deformation hardening, and other effects characteristic of initial flow phases [1-4]. With increasing s, however, they correspond less and less to the real situation. The dislocation density increases with s, the interdislocation distance decreases, and the interdislocation interaction forces fin~ = (i/2~)Gb2~, become large, where G is the shear modulus. For small E they can be neglected in comparison with the forces fext = boext acted on the dislocation by the external sources. For this reason the dislocation ensemble is a set of noninteracting individual dislocations. *Everywhere below the topic of discussion is the values of actual logarithmic deformation.

a2 7

Y;
~o

!
-

Fig. I.
' "

Deformation curve of nickel.

C@ntral Scientific-Research Institute of Construction Materials. Translated from Izvestiya Vysshikh Uchebnykh Zavedenii, Fizika, No. 3, pp. 7-22, March, 1991. 186 0038-5697/91/3403-0186512.50 9 1991 Plenum Publishing Corporation

f,nt

Fig. 2. Schematic representation of the deformation dependence of the forces acting on dislocations.
Deformation

Gradually fint approaches fext, and for some critical deformation they become equal (Fig. 2). Starting at this moment, a strong interaction is generated in the dislocation ensemble, collective effects are manifested, and the nature of dislocation displacement becomes dependent not only on the external stress o ext, but also on the change in configuration of the surrounding dislocations. The evolution of collective modes of motion in the ensemble of strongly interacting dislocations leads to generation of specific inhomogeneities in the dislocation density distribution, whose formation and displacement over the crystal obeys rigorously defined regularities. They have acquired the name mesodefects [5]. Mesodefects include hundreds and thousands of dislocations, extending to distances larger than interdislocation distances, but smaller than the size of a typical macroscopic defect, to which one usually assigns a grain of a polycrystalline aggregate. Mesodefect propagation over the crystal is accompanied by intense rotations and shears. Plastic deformation transforms from micro- to mesolevel, changing its kinetics substantially and visibly. Structural evolution is still more complicated. It changes not only quantitatively, but also qualitatively. This is primarily related to the appearance of mesodefects. The evolution of mesodefects is essentially also the evolution of structures of deformation origin. To the extent that the density of mesodefects is enhanced, so do their size, intensity, type, and characteristic fine structure change. In a plastically deformed crystal there occur still more complicated and thorough variables, so that the structure can differ radically from the original one following its large deformation. Plastic deformation in phases of mesodefect formation is called evolution [5]. Mesodefects play an active, not passive, role in the process of plastic deformation. Their formation leads to a realization of new localization channels of plastic deformation, valid at a higher scale level. The regular evolution of mesostructures in phases of evolution of plastic deformation is nothing else than a method of its realization. The study of this evolution implies the study of mechanisms of evolution of plastic deformation. The necessity of generating mesodefects can also be seen in the action of more general thermodynamic principles. Plastic deformation of crystals is an open^thermodynamic system, in which the mechanical energy propagates with velocity Sext .. (~ + ~el). It dissipates into work of plastic deformation, creation of elastic distortions and crystalline structure defects, as well as dissipation by partial transformation into heat. The following obvious relation holds between these components [6]:
A A A A
(2)

ucxt"

" 8-----Uint " " kel-~-F-~-D'


A

where ~int is the tensor of internal stresses, Sel is the velocity tensor of elastic deformations, and F and D are the accumulative velocities of latent energy and dissipation. ^ estimates show that the quantities $int .. Sel and ~ are substantially smaller than
^ A

Simple

~ext .. ~. Hence ~ext .. ~ is approximately equal to D, and for extension of deformations with constant velocity it is necessary to guarantee a constant rate of dissipation. One cannot achieve satisfaction of this condition solely due to homogeneous dislocation flow: with the accumulation of plastic deformation there is a disproportionate fast increase in stoppers and obstacles, hindering their mobility through the crystal. The dislocation flow is gradually depleted, and the level of internal stresses is enhanced. This is continued until cracks are generated, and brittle destruction of the sample occurs. For this not to occur, and for the energy supplied to the sample not to start being transformed primarily into elastic distortions, but continue to dissipate to the extent

187

of depletion of the previous channels by plastic deformation, new modes of plasticity must be found for the system. They become large-scale shears and changes in the internal regions of the crystalline sample. Mesodefects are formed for their realization. Consequently, the mesostructure of interest to us is a structure of dissipative type. It creates conditions for the continuation of plastic deformation at high E values. Different indications of dissipative structures, characteristics of phases of evolution of plastic deformation, as well as the kinetics of their formation and their subsequent evolution were investigated for a wide class of metals and alloys [5]. Data are available on metals with bcc, fcc, and hcp lattices, obtained for different schemes of loading and conditions of deformation. On the whole they make it possible to formulate the principles of the structurally kinetic approach to the description of evolution of plastic deformation. EVOLUTION OF DISLOCATION STRUCTURES IN PHASES DIRECTLY PRECEDING FORMATION OF MESODEFECTS The physics of evolution of plastic deformation includes the basic statements of classical dislocation physics of hardness and plasticity. By their means one describes processes occuring inside mesodefects.* Other concepts are introduced to describe plastic deformation at the following structural and scale levels. Besides the concepts of collective effects and mesodefects, these are concepts of joint dislocations, broken dislocation boundaries, fragmentations, rotational modes of plasticity [5, 7], as well as structural deformation levels [8, 9]. Understanding their meaning is simplest by treating, for specific examples, the evolution of deformation structures from their moment of generation to the moment directly preceding the nucleation of mesodefects. We select a well-annealed, dislocationless region of the crystal, such as a grain, and follow what occurs with it with increasing e. Intragrain Structures. At the initial phase of plastic deformation the dislocation structure of internal bulk grains can be characterized in terms of the dislocation density p, and when a cell structure is formed also by the density of dislocations Pb located at the cell boundaries, the thickness of cell boundarires A, the cell sizes d, their disorientations Their variation with increasing @, and distribution functions over these characteristics. E is illustrated in Figs. 3, 4 [i0]. *The bulk of the crystal, confined by mesodefects.
#,)%~ cm

4
\
a

iO ~o

!O~}

,
1 q --' ' !

I0
6 c

Fig. 3. Cell structure parameters as a function of the value of plastic deformation (polycrystalline nickel, stretching; T = 300~

188

'm ~0,0

tgo[ 4,o
.o

Fig. 4. Variation of distribution function over cell sizes (areas) with increasing extent of deformation E(Ni): i ) 0.3, 2) 0.7, 3) 1.1.
o,r o,2

qa

0A

S, pm 2

The dislocation density increases, achieves a maximum of order i0 II - 1012 cm -2, and then starts falling off. Following deformations exceeding 3-4 units, the dislocation density decreases to values of the order of 10 s cm -2 , so that homogeneously oriented regions of strongly deformed crystals become again practically dislocationless (Fig. 3a). The cell thickness and mean size drop monotonically with increasing s, and then stabilize for large E at levels approxiamtely equal to 0.i and 0.2 pm, respectively (Fig. 3b, c). A saturation effect is typical for these structures. A sharp variation in p, 9b, A, and d occurs in the s interval from 0 to approximately 0.4. For increasing deformation beyond 0.4 they change weakly. The same can be verified by analyzing the distribution function. As an example, we show in Fig. 4 the cell distribution over size for nickel with a mean grain size D = 160 pm (single-axis stretching at room temperature with velocity i0-~-i0 -3 sec -I) [i0]. The distribution narrows with increasing E, and the cells themselves reduce to fragments on the average. The most likely average cell size approaches asymptotically d = 0.2 Dm, which is in agreement with Fig. 3c. A sharp variation in the distribution function occurs in the same interval 0 < s N 0.4. < The disorientation angle O of neighboring cells does not exceed 0.2 ~ . Cells are disoriented in a disorderly fashion even at distances of dozens of d, and are disoriented by the same small angles of order 0.1-0.2 ~ , as are neighboring cells. With increasing deformation the characteristics mentioned of cell structures remain practically unchanged. Up to a deformation s = 1.2 the disoriented cells remain distributed chaotically, but for some cells the value reaches 0.5 ~ [5]. Assuming that the plastic deformation process is directly related to evolution of the defect structure and directly reflected in it, the experimental data provided imply that for E > 0.4 the mechanism of plastic deformation by means of quasihomogeneous motion and multiple dislocations (including those leading to formation of cell structures) is exhausted. Since, nevertheless, plastic deformation is extended, though at the same velocity as earlier, it must emerge at other structural levels and be realized due to formation of larger scale defect formations, whose evolution must correlate with increasing ~. Before turning to concrete mesodefects and to explanation of their nature, we dwell one one aspect of structural variations in a plastically deformed polycrystal. Variations at Disorientation Boundaries. A dislocation structure varies not only in the grain bulk, but also at the intergrain boundaries. Lattice dislocations stick to them and nucleate, realizing plastic deformation of adjacent grains. The dislocation distribution at a boundary with a normal N, separating s- and s'-grains, is conveniently described by the difference tensor AB N, equal to [ii]
A 2 M s

sol

p=~ \ s m ~ ]p

.J

~p,

(3)

where p~ is the linear density of lattice dislocations of the glide system at the boundary of s grains, while ~pS = arc cos ( n A q s -p.

(n, b)~) , disabled

An equation describing the variation of the tensor AB N under the assumption of invariance of boundary orientation is obtained from (3) by straightforward transformations. It is
A 2 M

= N
a=l

(-p=l

(4)

189

where p is the bulk dislocation density of the system ( n, b ) s, and V is their velocity. Introducing into the treatment the plastic distortion tensor of the s-th grain, equal by definition to
A ~! p=l

we obtain a more compact description of the evolution equation: d ^ A --{ABs q - N x [~]g} = 0, dt

(6)

where [~]N i s t h e jump normal N. I n t e g r a t i n g

in t h e p l a s t i c d i s t o r t i o n t e n s o r a t t h e i n t e r g r a i n boundary w i t h (6) p r o v i d e s t h e g e n e r a l i z e d B e l l o w - B e e l b y e q u a t i o n
A A A

ABN=BoN-- N X [ ~ ] N ,
where B0N is the dislocation density tensor at the original boundary in the Undeformed crystal. Boundaries as Sources of Long-Range Stresses. We partition the plastic distortion

(7)

tensor into symmetric ~ and antisymmetric ~ parts. The first is the plastic deformation tensor, and the second will be called the plastic deflection tensor with an invariant lattice. ^ Correspondingly, this partition is represented in the form of two parts and the tensor ABN:
A A A A A A A

AB~ = AB~ q- ABe; AB~ = -- N X [s]~; AB~ = -- N X [~]n, N.


The tensor ABNm describes the distribution at the boundary N

(8)

of the dislocations re-

lated to the plastic deflection of a grain, and ABN E is related to its shear plastic deformation. All dislocations related to ABN ~ are distributed self-consistently and do not create long-range internal stresses. As to dislocations described by the tensor ABNg , the situation is more complicated. Part of them also forms self-consistent configurations, and does not lead to generation of planar sources of long-range stresses. Another part consists of the planar, unbalanced dislocation distributions. To the extent of the deformation they transform the boundaries into intense sources of internal stresses. For an infinitely extended boundary the corresponding stress field equals [Ii]:
M

I - - ~ (~_B ~)~ x ( ~ + -~.~)~ + ~- (~a~ ~)~ ( ~ -F


Here v is the Poisson coefficient, ~ = n X N ,

(t).

(9)

n = ~ X N, and (ABNE)pS is the contribution

of the dislocation glide system (n, b)p s to the tensor ABNg; and t is the coordinate, measured along the normal to the boundary. The first term in the square brackets of the first part of Eq. (9) refers to the stress field due to the edges, and the second - to noncompensated screw dislocations, respectively. We estimate the stress ON int. For this we use the scalar notations

o~ ~ G

int

[H~,

(lO)

where ~ is a geometric factor, which can be taken to be 0.5 within the first approximation. We put
s. = ~o ~ \% /

so

F~,

(11)

where F s is the Schmid factor, ~s is the shear stress in the s-th grain, o is the stretching strain in both grains, and g0, To, n are parameters of the velocity dependence. For small grain disorientations the Schmid factors differ insignificantly. case we have a simple expression for the jump in deformation [~] = ~--~----~(I--~)"--~----- n~ ( ~ ) , In this

(12)

190

where AF is the jump in the Schmid factor for transition through the disorientation boundary, and E and F are the mean values of the deformation and of f. We select two grains, separating the boundaries of a slope with disorientation angle ON, deformed due to a single glide of edge dislocations. We introduce the notations cos ~ = i/b (b I) , cosx = (n I) . In our model cosx = -sin~N, so that F = 0.5 sin2~N. Hence the jump in the Schmid factor equals F = 0.5 [ s i n ( 2 ~ + 2@) - sin2~], while

AffF =_ sin (2,~+ 20) - - sin2Q? F sin 2~ G l i d i n g u s u a l l y o c c u r s i n p l a n e s o r i e n t e d a t an a n g l e n e a r 45 ~ w i t h r e s p e c t t o t h e stretching axis. T h e r e f o r e , f o r f u r t h e r e s t i m a t e s we t a k e ? = 45 ~ . U s i n g , t h e n , ( 1 0 ) , ( 1 2 ) , and ( 1 3 ) , we o b t a i n int 1 2 ~N - - ~ 2 ~ n G ~ s i n Ox,

(13)

(14)

i.e., the long-range internal stresses induced by the disorientation boundaries in plastically deformed crystals increase linearly with s, and are directly proportional to the square of the sine of the disorientation angle. The sharp O-dependence leads to the result that weakly disoriented boundaries do not become sources of any noticeable long-range stresses up to the limiting deformations, while at the same time the large-angle boundaries at relatively low deformations start generating stresses capable of affecting substantially the geometry and dislocation gliding intensity in adjacent regions. As an example consider cell boundaries with disorientation O = 0.2 ~ and a large-angle boundary of intergrain type with 0 = 20 ~ . We put 6 = 0.5 and n = 4. For the cell boundaries we then have o intN = l'2"10-s GE, while for the large-angle boundary o int = 0.I Ge. The cell structures remain practically unstressed, while near the intergrain boundaries the internal stresses reach the limits of fluidity (-~G/500) already following a deformation s = 0.5%. Disorientation Variation at Boundaries. Compensated dislocation configurations generate variations in disorientation. It follows from the preceding discussion that contributions to them are provided by incompatibilities in both plastic deflections with an invariant lattice, and purely shear plastic deformations. The increase in the disorientation vector k@ can also be represented in the form of two terms
'-O,v = nO?c

nO\.,
A
=

(15)

where

[12]

Z~
~0~,
9

A
9

'[~i\, ~Oi.. '

- - N X [ q N , ~ . "'

(16)

Here E is the trivalent

antisymmetric

Levi-Civita

tensor.

The f i r s t c o n t r i b u t i o n is quite of the boundary attitude, and i s t h e [13]. The d e f o r m a t i o n c o n t r i b u t i o n of the boundary, and, consequently, limits of a single boundary.

trivial. It is independent of the plane orientations basis of contemporary theories of deformation textures i s n o t so o b v i o u s . I t d e p e n d s on t h e a t t i t u d e planes its value changes from facet to facet even within the

The jump i n p l a s t i c d e f o r m a t i o n [~]N d e p e n d s on t h e d i s o r i e n t a t i o n of neighboring grains. U s i n g t h e same a r g u m e n t s a s i n t h e p r e c e d i n g s e c t i o n , one e a s i l y o b t a i n s an e q u a t i o n f o r estimating the quantities kO ~, s i m i l a r t o Eq. ( 1 4 ) : AO~ _~a vnesinf@~, ( 17 )

where a N is a geometric parameter, different f r o m 6, b u t a g a i n c l o s e t o 0 . 5 . It follows from (17) that weakly disoriented cell structures practically do not change their disorientation during the process of plastic deformation, which is observed experimentally. On the contrary, large-angle boundaries can substantially affect their disorientation. Approximately the same contribution, such as incompatibility of deformations, is provided to k@ N by incompatibility deflections.

191

Equation (17) was obtained for the case of approximately scalewise homogeneous dislocation fluxes exceeding the grain size, when the incompatibilities in plastic deformation are small. However, the maximum jump in plastic deformation occurs at the boundaries, where deformation and nondeformation grains are joined. In that case there is no dependence on disorientation, and

AO "=-- A O m a x
where ~ is a geometric factor close to unity.

"v

aS,

(18)

GENERATION AND EVOLUTION OF MESODEFECTS Disclination of an Orientational Discrepancy. The fact that the magnitude and direction of the supplementary disorientation vector depend on the plane orientations of the adjacent boundary implies that disorientation discrepancies are generated on boundary fractures and joints during plastic deformation. The simplest scheme of such a process is illustrated in Fig. 5, where boundary fractures are shown in which facets with normals 7Vl and N 2 are joined. The difference deformation [c] is realized by glides in the s-th grain of lattice dislocations with a Burgers vector b in planes at identical distance h from each other. Following deformation [c] = pbX at the intersection of shear planes with boundaries, dislocations are stored in m = 0Xh (X is their mean free path). Their model can be represented in the form of two sets of dislocations: gliding and sedentary with Burgers vectors b s i n ~ N and b cos~N, respectively (Fig. 5b), sin~N = ~iV The distances between the dislocation groups equal h/cOS~N, whence it follows that the value of supplementary disorientation is AO~ = @kb) # cos %v, (19)

where k is the normal unit vector to the plane of the figure. Contributions to the A O found are provided by the rotational A O ~ and deformational AO~. components of the supplementary disorientation. In the given case they are AO~=--I 2 h O ~ = 1 # (pbk) (cos 2 ~N - - sin 2 ?.v). 2 (20) #(obk);

The contributions h e } are identical for all facets at the given boundary, while the contributions AO~, depend on N as ( n N ) 2 Therefore, the difference effect is determined only by the deformation contributions. The discrepancy of the disorientation, concentrated at the fracture, equals

s = l__k (?bk) sin 2 ~N = --~ k [~l sin~ 9N. 1 2

(21)

hy-i- k j

-,...

-I
c

Fig. 5. Scheme illustrating the formation process of a joint disclination at a fracture of a large-angle grain boundary: a) original situation; b) dislocation representation of defect, c) dislocation-disclination representation.

192

a t

' 25
F i g . 6. Fundamental s t a g e s o f t h e e v o l u t i o n p r o c e s s o f p l a s t i c def o r m a t i o n [ 24] : a) g l i d e d i s l o c a t i o n under t h e a c t i o n o f e x t e r n a l s t r e s s e s ; b) f o r m a t i o n o f s t i c k d i s c t i n a t i o n s ( d e n o t e d by shaded t r i a n g l e s ) and o f n o n - s e l f - c o n s i s t e n t p l a n a r d i s t r i b u t i o n s o f d i s l o c a t i o n s ; c) a c t i o n o f accommodation g l i d i n g ( t h i c k l i n e s ) ; d) f o r m at i on o f broken b o u n d a r i e s - i n i t i a l phase o f f r a g m e n t a t i o n ; e) fragmented grain. Thus, we r e a c h t h e c o n c l u s i o n t h a t t h e n o n i d e n t i t y o f p l a s t i c d e f o r m a t i o n s o f n e i g h b o r i n g g r a i n s g e n e r a t e s s p e c i f i c l i n e a r m e s o d e f e c t s o f r o t a t i o n a l t y p e , which became known as s t i c k d i s c l i n a t i o n s [5, 14]. In t h e g e n e r a l c a s e , t h e y e n c l o s e p l a n a r m e s o d e f e c t s o f t h e Somilyano d i s l o c a t i o n t y p e . In F i g . 5c t h e y a r e r e p r e s e n t e d i n t h e form o f c o n f i g u r a tions of equidistant glide planes. The problem o f c o n s t r u c t i o n o f s t i c k d i s c l i n a t i o n s was c o n s i d e r e d in d e t a i l in [ 1 4 ] . Also p r o v i d e d were e x p r e s s i o n s f o r r o t a t i o n a l s t i c k d i s c l i n a t i o n v e c t o r s , l o c a t e d in a j o i n t o f k g r a i n s ( o r b o u n d a r i e s ) f o r a r b i t r a r y v a l u e s o f t h e p l a s t i c d e f o r a m t i o n t e n s o r jumps o f t h e s e g r a i n s :
~ j=l g

a = -- Z N j X [Z]NjN, where t h e s u b s c r i p t j c o u n t s j o i n t
was investigated in [15].

(22) field due t o s t i c k d i s c l i n a t i o n s

boundaries.

The s t r e s s

Broken Dislocation Boundaries. Stick disclinations located in joints and on boundary fractures form in the bulk of the deformed crystal a three-dimensional grid of linear sources of intense internal stresses (Fig. 6a, b, c). The rotation vector of stick disclinations is proportional to the accumulated deformation. For disclinations located at fractures we obtain, generalizing (12) and (21), an equation expressing ~ in terms of e, the disorientation angle, and the fracture angle:

c2 = ~ }nsin2Osin29; ~ m a x = ~ s i n ~ 9
2 2

(23)

At some threshold deformation ~0 the intensity of stick disclinations is so high that the stress field generated by them excites collective shapes of motion in the dislocation ensemble located nearby. The result is nucleation at the stick disclinations of broken dislocation boundaries. There occurs relaxation of short-range stress by stick disclinations, since part of the disorientation discrepancy emerges from the joint into the grain volume (Fig. 6d). A broken boundary is usually interpreted as a dislocation realization of partial disclination, and within the first approximation - simply as partial disclination [17, 18]. The basic difference in that case is observed only in the construction of the region directly adjacent to the geometrical line of the corresponding linear defect. At the fractured boundary this is a cylindrical region of a diameter of order (5-10) b/a, filled chaotically by distributed dislocations, while the kernel of partial disclination is represented more compactly and completely. The real situation is more complicated, and the fragmented boundary can be interpreted as a dislocation model of partial disclination in one case only: when 'the direction of the fracture line ~ is parallel to the rotation vector of partial disclination Q or the boundary disorientation vector O . In all remaining cases the fragmented dislocation boundary does not coincide with the dislocation model of partial disclination, and is a true defect of rotational type.

193

There are two principal differences. Firstly, at the fragmented boundary one cannot indicate uniquely positions of rotation axes. The boundary can be represented as a surface defect, forming a set of identical rotations 0, each of which is completed around an axis passing through the 0-point and covering part of the surface relating to this 0-point (the Wigner-Seitz cell) [19]. F o r partial disclination there always exists a unique axis of rotation. Secondly, for disclinations having a torsional component the elastic displacements near disclination lines diverge upon moving away from the center of rotation [17, 18]. At the fragmented boundary, as for a real substantial defect, this divergence may not take place, and all points along the fragmentation line must be equivalent. If one attempts to describe the fragmented dilocation boundary in terms of disclinations, it seems to be a disclination chain - a defect consisting of subsequently linked disclination segments, supplemented at the linking point by closed dislocations [20, 21]. The stress tensor field of a fragmented dislocation boundary was calculated in [15]. The components of this tensor are ~
022--

2=(I--~)

Inr-}-~--F~-

0~3 (In r + ~
+--

x~

+1

= ~ ( ~ . + ~22); ~1~ = 2=(1


~) r ~ ' -

~13= ~2:~

__ 09___ t n r + r 2 L 4~ -= -4~
-

2 4~

( xx~ -~= ~ r ~ --{Inr+-~ \ r ~ ' 2/

(24)

where a rectangular coordinate system with basis (e,,e~, e~) has e~=NX~, e2=N, e3={.

been introduced, with

If the vector ~ is parallel to the fracture line (~ = a e8 ), the field o coincides with the field of wedge dislocations [17, 18]. At the same time the stress field of the fragmented boundary, corresponding to the components ~z and ~2, differs from it for the stress field of partial disclination. As seen from (24), at the fragmented boundary there exists no stress divergence along ~ typical of rectilinear torsional disclination. By its structure the stress field of the fragmented boundary is the field of an effective edge dislocation with a Burgers vector ~eff= ~2X~el-- ~lX~d2" The predictions on generation of mesodefects in plastically deformed crystals, such as stick disclinations and fragmented dislocation boundaries, are fully verified experimentally. By methods of scanning electron microscopy it was demonstrated that they indeed nucleate on fractures and joint large-angle boundaries, and then progress gradually into the depth of the body of the grain. In that case several characteristic configurations are encountered [16]: i) single rectilinear or smoothly bent fragmented boundaries; 2) branched fragmented dislocation boundaries; 3) dipoles consisting of pairs of parallel fragmented boundariers, each of which having disorientations identical in magnitude, but opposite in sign; 4) loop-shaped closed configurations consisting of fragmented boundaries. Fragmented dislocated boundaries have a number of clearly distinct features. These dense dislocation formations extend over several microns, which are sharply distinct on the background of loose, orderly, and much smaller extended boundaries of cell structure. The largest differences are manifested in the value of disorientation angles. Numerous measurements have shown that disorientations at fragmented boundaries exceed typical intercell disorientations by approximately an order of magnitude. For example, under conditions of deformation at room temperature the characteristic values of @ at fragmented boundaries in iron, molybdenum, titanium, nickel, and other metals and alloys are 1-2 ~ [5]. Knowing the initial disorientations at fragmented boundaries @min, and using expression (23), one can find the threshold deformation value c 0 at which nucleation starts:
8O

~nsin2@ sinZ?

194

Lb" i0- ~,
7~ cm- i o

i
i o

Fig. 7. Total length L b of fragmented boundaries per unit area of ground Ni as a function of deformation g.
o,8 ~,2 c

o,4

Taking the same values as earlier for the parameters appearing in this equation (p = 1/2; n = 4; @ = 20~ @min = 2~ one obtains E 0 = 0.2. Thus, theory predicts that, starting with deformations of order 20%, one must expect the appearance of the first mesodefects of the type of fragmented boundaries in a structure of deformed samples. The first mesodefects occur at the phase of deformation in which ordinary characteristics of dislocation structures, such as p, Pb, A, andd, display a tendency to saturation. As was shown, this corresponds to g = 0.3-0.4. In [i0] it was attempted to determine the quantity g0 accurately. With this purpose we constructed for nickel the g-dependence of the total length L b of broken dislocation boundaries manifested per unit area of the transverse cross section (Fig. 7). The approximation at g = 0 gives g0 = 0.23, which is in good agreement with the theoretical estimate provided. FRAGMENTATION Fragmented dislocation boundaries propagate with a velocity proportional to the velocity of plastic deformation (Fig. 7), branching and intersection with each other, and gradually fragment the crystal, i.e., they partition it into strongly disoriented microsamples - fragments (see Fig. 6). Fragment sizes decrease with increasing s, while the disorientations are enhanced. The rate of volume implication of a plastically deformed crystal in the fragmentation process is determined by three factors. Firstly, the density of nucleation centers of fragmented dislocation boundaries - mesodefects, inducing fragmentation. It depends on the defect structure of the crystal, such as the grain sizes, and extent of perfection of largeangle boundaries. Besides, it is proportional to the fractional volume of the nonfragmented crystal (i - N), where n is the fractional volume covered by the fragmentation. Secondly, the quantity of interest to us dq/dg is determined by the degree of preparedness of the nucleation centers mentioned of the fragmented boundaries to emergence of similar mesodefects. Therefore, dq/dg is proportional to the intensity of stick disclinations ~ thus formed, and, consequently, also to (s - E0). Thirdly, dN/dg is proportional to the propagation rate of fragmented dislocation boundaries dLb/dg, which, according to Fig. 7, is a constant quantity. The considerations provided make it possible to write down an equation for the fragmentation rate:
dq/de ----z(l--n) (e--e0),

(25)

where < is a structural-geometric parameter, which can be found from experiment or by turning to a more detailed structural model. Integration of (25) provides an expression for the dependence of q on the extent of deformation ~(~)=l--exp[--~(~--%)~] 9 (26)

Figure 8 shows experimental points for the g-dependence of the fractional fragmented volume for armco iron, deformed by drawing at room temperature [22]. Also given is the theoretical dependence q(g) in the form (26). For K = 0.55 the experimental points are in good agreement with the theoretical curve, implying that natural assumptions have been used in deriving Eq. (25). With increasing g fragmentation does not simply cover the whole larger volume of the deformed crystal. There occurs a change in the fragmentation structure itself, and the boundaries become rectilinear. In external shape they approach boundaries of intergrain type. The internal volumes of fragments are free of dislocations. The density of the latter, both inside fragments and in the crystal bulk, drops, and following large deformations,

195

2.O

o, qo85 ojo+ o,222 o,29, :,~.s~o~27,~, ~m olz

Fig. 8

Fig. 9

Fig. 8. Fractional volume ~, occupied by the fragmented structure, as a function of deformation e, armco iron drawn at room temperature [22]: experimental points (A) and theoretical curve for K=0.55. Fig. 9. Transverse size distributions of fragments in a fragmented structure of armco iron, created under conditions of drawing at T = 20~ for E = 0.91 (A); 1.38 (.); 2.3 (o); 4.6 (D). of the order of 3-4 units, they emerge out of the level of values typical of the annealed crystal. The fragment shape changes too. They are extended or compressed in the directions of boundaries of the principal deformation axes. For extension the fragments become cigarshaped, and for compression or flattening they acquire the shape of flattened films. The transverse sizes of fragments change much less. Figure 9 shows a series of distribution curves of transverse sizes of fragments for the same drawn armco ion.* The character of variations of these curves is exactly the same as observed for cell structures (see Fig. 4). With increasing deformation the distribution function narrows down and becomes displaced toward smaller sizes. The variations of dispersion and of the most likely d value from have a noticeable saturation effect. The d value tends to dmi n = 0.15 ~m, and displaced to 3.6-10 -3 . The saturation approaches the region of deformation of order 1-2. The facts provided show that the Polanyi-Taylor principle does not apply to a fragmented structure. At the same time the macroscopic sizes of the sample during a longitudinal plastic deformation vary by dozens of times, and the corresponding sizes of fragments remain practically unchanged. As in the case of cell structures, this can imply inclusion of larger-scale modes of plasticity. Indeed, observations show that with the formation of quite perfect fragmented structures, intense rotational shear instabilities and faults appear at the background. They have shapes extending over hundreds of microns of rigorous rectilinear sections, oriented along the principal deformation axes of thin "knife" boundaries [5]. Large-scale blocks of fragmented crystals, contained between these boundaries, perform with respect to each other avalanche-type large growing plastic shears and rotations. Detailed crystal-geometric analysis of fragmented structures at different phases of deformation reveals one more mode of evolutionary development. It seems that with increasing g there is an increase in fragment disorientations. Since for any E value the disorientation distribution in the fragmented structure obeys a quite complicated law [5, 23], the explanation of the character of its evolution is possible only by statistical analysis. Figure i0 shows the disorientation distribution in a fragmented structure of hydroextrudable molybdenum at T = 250~ In the original state it was a monocrystalline sample, *In the study of fragment distribution functions over sizes we have enjoyed the collaboration of V. A. Potapov. 196

f6 ']

nJt|H h~ h ~ifli~!,;::!ll

Fig. i0. Disorientation distribution in a fragmented structure of hydroextrudablemolybdenumfor e = 0.67 (a); 1.89 (b); 2.56 (c). with orientation axis <011> along the direction of hydroextrusion. It is seen that with increasing s the distribution is displaced toward large angles. This is even more noticeable in analyzing the distribution of the true disorientation angles Otrue , which for large E much exceed the minimal crystallographic equivalent disorientation angles. It has been shown that for a given deformation the maximum values of the true disorientation angles obey an equation of the type
~ e = ~ (~ - - %), (27)

w h e r e a i s a p a r a m e t e r c l o s e t o u n i t [5, 2 3 ] . I t i s i n good a g r e e m e n t w i t h t h e t h e o r e t i c a l e s t i m a t e ( 1 8 ) , and i m p l i e s t h a t i n a f r a g m e n t e d s t r u c t u r e plastic deformation is accompanied by c o n t i n u o u s r e o r i e n t a t i o n s of the crystalline lattice of fragments. And, e v e n t h o u g h the controlling micro-mechanism is in this case motion of a fragment of lattice dislocations along the body, the process of fragment reorientation is substantially specific, distinct from the trivial consequence of emergence of lattice dislocations, realized by lattice deformations, on its boundary.

Indeed, in the latter case, along with crystalline lattice reorientation it would also undergo a change in shape, which, according to the Polanyi-Taylor principle, would follow the change in shape of the macrosample. Precisely this situation occurs in a grain structure. This grain is not only reoriented, but also changes its shape. For example, during extension its transverse sizes are stretched and decrease by the value of local narrowing. As was shown above, nothing similar occurs in fragments. Their transverse sizes, despite the increasing deformation and the increase in disorientation which is proportional to it, practically do not change. This makes it possible to introduce to the treatment the concept of rotational modes of plasticity [5, 23], which are realized during the phases of development of plastic deformation and participate actively in the evolution of structural fragmentation. LITERATURE CITED I. 2. 3. 4. 5. 6. 7. 8. 9. i0. ii. 12. J. Friedel, Dislocations, Pergamon Press (1964). J. P. Hirth and J. Lothe, Theory of Dislocations, 2nd ed., Wiley, New York (1982). R. W. K. Honeycombe, The Plastic Deformation of Metals, 2nd ed., E. Arnold, Baltimore

(1984).
A. N. Orlov, Introduction to the Theory of Crystalline Defects [in Russian], Vysshaya Shkola, Moscow (1983). V. V. Rybin, Large Plastic Deformations and Destruction of Metals [in Russian], Metallurgiya, Moscow (1986). V. V. Rybin and A. A. Zisman, Fiz. Met. Metalloved., 69, No. 4, 5-14 (1990). V. V. Rybin, Thesis, Kiev (1979). V. E. Panin, V. A. Likhachev, and Yu. V. Grinyaev, Structural Deformation Levels in Solids [in Russian], Nauka, Novosibirsk (1985). V. E. Panin, Yu. V. Grinyaev, T. F. Elsukova, and A. G. Ivanchin, Izv. Vyssh. Uchebn. Zaved., Fiz., No. 6, 5-27 (1982). A. S. Rubtsov and V. V. Rybin, Fiz. Met. Metalloved., 44, No. 3, 611-622 (1977). V. V. Rybin and N. Yu. Zolotarevskii, Fiz. Met. Metalloved., 57, No. 2, 380"390 (1984). V. V. Rybin, Theory and Computer Modeling of Defect Structures in Crystals [in Russian], Akad. Nauk SSSR, Sverdlovsk (1966), pp. 55-64.

197

13. 14. 15. 16. 17. 18. 19. 20. 21.

Ya. D. Vishnyakov, A. A. Babareko, S. A. Vladimirov, and I. V. Egiz, Theory of Texture Formation in Metals and Alloys [in Russian], Nauka, Moscow (1979). V. V. Rybin, A. A. Zisman, and N. Yu. Zolotarevskii, Fiz. Tverd. Tela (Leningrad), 27, No. i, 181-186 (1985). V. V. Rybin, N. Yu. Zolotarevskii, and I. M. Zhukovskii, Fiz. Met. Metalloved., 69, No. i, 5-26 (1990). A. N. Vergazov, V. A. Likhachev, and V. V. Rybin, Fiz. Met. Metalloved., 42, No. i, 144-154 (1976). R. DeWitt, Continuum Theory of Dislocations [Russian translation], Mir, Moscow (1977), p. 208. V. I. Vladimirov and A. E. Romanov, Dislocations in Crystals [in Russian], Nauka, Leningrad (1986), p. 223. W. Bollmann, Crystal Defects and Crystalline Interfaces, Springer-Verlag, Berlin (1970). V. V. Rybin and I. M. Zhukovskii, Fiz. Tverd. Tela (Leningrad), 19, No. 8, 1474-1480

(1977).
I. M. Zhukovskii, N. Yu. Zolotarevskii, and V. V. Rybin, Dislocations; Experimental Studies and Theoretical Description [in Russian], Fiz. Tekh. Inst., Leningrad (1982), pp. 104-117. A. N. Vergazov, V. V. Rybin, Yu. Ya. Meshkov, et al., Fiz. Met. Metalloved., 69, No. 6, 171-177 (1990). A. N. Vergazov and V. V. Rybin, A Method of Crystalline Analysis of the Structure of Metals and Alloys in the Practice of Electron Microscopy [in Russian], Leningr. Dom Nauchno-Tekh., Propagandy, Leningrad (1984). N. Yu. Zolotarevskii, V. V. Rybin, and I. M. Zhukovskii, Fiz. Mec. Metalloved., 67, No. 2, 221-232 (1989).

22. 23.

24.

DISLOCATION STRUCTURE AND DEFORMATION HARDENING OF bcc METALS S. A. Firstov and G. F. Sarzhan UDC 669.01

The relationship between mechanical properties and laws governing dislocation substructure development during plastic deformation and other treatments of single-phase materials has received much attention from materials scientists over the past three decades. Consider the example of single crystals: It has been clearly shown how the step-by-step development of deformation' hardening relates to changes in the deformation mechanism and the corresponding structure formation [1-5 et al.]. Turning to polycrystalline materials, one usually finds them characterized by parabolic hardening [2, 3, et al.] in which the coefficient and index of deformation hardening are determined by the structure and nature of the material. Detailed studies of the evolution of dislocation substructures [6-22 et al.] over the past 10-15 years have helped establish the characteristic stages in the evolution of defect structures and the corresponding change in the deformation hardening mechanism. In particular, it has been demonstrated that, at large plastic deformations, strength is determined not by dislocation distributions of different densities and morphologies, but by the socalled fragmented (disoriented cell) structures. In this article we analyze a number of experimental facts that help systematize regularities in structure formation, and we discuss the stages of structure formation during the parabolic stage of hardening, ahd the unique deformation mechanism operating during the final (linear) stage. We also discuss possible new avenues of research, including theory development. An important problem is the theoretical description of the evolution of the dimensions of structural elements during large plastic deformations.

Institute of Problems in Materials Science, Academy of Sciences of the Ukrainian SSR. Translated from Izvestiya Vysshikh Uchebnykh Zavedenii, Fizika, No. 3, pp. 23-34, March, 1991.

198

0038-5697/91/3403-0198512.50 9 1991 Plenum Publishing Corporation

Vous aimerez peut-être aussi