0% found this document useful (0 votes)
49 views37 pages

Fracture Mechanics

Fracture mechanics

Uploaded by

MudassarHashmi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
49 views37 pages

Fracture Mechanics

Fracture mechanics

Uploaded by

MudassarHashmi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 37

Engineering Fracture Mechanics Vol. 51, No. 6, pp.

997-1033, 1995
Copyright ~L 1995 ElsevierScienceLtd
Pergamon 0013-7944(94)00323-8 Printed in Great Britain. All rights reserved
0013-7944/95 $9.50+ 0.00

FRACTAL FRACTURE MECHANICS--A REVIEW


GENADY P. CHEREPANOV
Department of Mechanical Engineering, Florida International University, Miami, FL 33199, U.S.A.

ALEXANDER S. BALANKIN
Materials Science Department, Mexican Institute of Technology (ITESM), Mexico City 52926, Mexico
and
VERA S. IVANOVA
Physics of Metals Department, Baikov Institute of Metallurgy, Russian Academy of Sciences,
Moscow 117334, Russia

A~tract--The revolution in geometry, which has recently created the notion of the fractional dimension
of real world bodies and has formed fractal geometry, has substantially influenced fracture science. An
experimentally measured crack in real materials has appeared to be substantially a fractal, that is, a
geometrical object of fractional dimension. In the present paper, an attempt has been undertaken to
provide an account of some achievements in this area, herein called "Fractal Fracture Mechanics".

1. INTRODUCTION
THE CLASSICALmethods of solid mechanics assume that the homeomorphism of deformation
sometimes does not provide an adequate description of the theological behavior and fracture of
real materials. Irreversible macro deformation and the fracture of solids are predetermined by
material behavior in nanoscale.
The fact that continuum approximation is often unsatisfactory for a real material is now
beyond doubt. In man-made structures, a variety of nano-, micro-, and macro-defects appears at
the production stage that may evolve during the structure's service life. Numerous geophysical and
fractographic studies [1-4] indicate the non-Euclidean character of the geometry and the defor-
mation of the fracture of solids. Moreover, it has been empirically established [2, 3] that defects
emerging during deformation of even initially homogeneous continuous media gain an hierarchical
block structure, whose specific spatial scales, L~, obey the relation, X = L,+~/Li= const,
i = 0, l, 2 . . . . . As a consequence, assumption of the homeomorphism of deformations (considered
as the mapping of a deformed medium in the Euclidean coordinate-time space) is generally not
met. The most sophisticated fracture theory based on the use of invariant integrals [5, 6], which
allows one to take into account the topological features of fracture [7], is free of this drawback.
However, this theory does not take into account the fractal nature of the geometry of the
deformation and fracture of solids. The phenomenon of"fractality" most strikingly manifests itself
in the following aspects:
(l) the statistical self-similarity of crack configurations, whose metric dimension does not
coincide with the topological dimension, dr, and usually has the fractional value,
dr < dn < dr+j, where dn is the Hausdorff-Besicovitch metric dimension [l];
(2) the multifractal character of the spatial distribution of defects, pores, and cracks in a
deformed medium [1-3, 8-10], and the power laws of their size distributions I11, t2]; and
(3) the scaling invariance of the pore space and cracked surfaces in solids [1-3, 8-12].
During the last decade, the foundations of a novel theory were laid down that may be termed
fractal solid mechanics [1-3, 8-27]. Like the theory based on invariant integrals [5, 6], fractal
mechanics allow a sufficiently rigorous mathematical formulation that meets the requirements
adopted in classical solid mechanics (in contrast to the numerous heuristic models [28-30] that rely
upon intuitive synergetic ideas[31-33] applied to the problems of the mechanics of deformed
EVM 5n/6-1 997
998 G.P. CHEREPANOV et al.

media). It has been shown that the fractal character of the geometry of the deformation and fracture
of solids is due to the fundamental effect of transverse deformations [16, 20]. Furthermore, the
topological features of a deformed medium are determined by the transverse deformations (the
effective coefficient of transverse deformations coincides with Poisson's coefficient only by the limit
of infinitesimally small elastic strains), as well as by the mechanics governing the deformation and
fracture on nano-, micro-, macro-levels [14-23].
Similar to the elasticity theory and the mechanics of deformed solids [34, 35], the effect of
transverse deformations within the scope of phenomenological fractal mechanics is considered as
an experimentally established fact that is used in constructing the fundamental equations of the
theories of elastic and plastic deformation or fracture [36].
The irreversible deformation and fracture of solids may be related to a class of processes in
which the macroscopic effects are predetermined by microscopic level behavior [3, 20, 36]. For this
reason, a reliable prediction of the response of a solid to an external force should be based upon
a clear understanding of the mechanics of the processes in nanoscale. It is evident that only the
description of these processes within a single system, taking into account the interrelation of
different processes in nano-, micro-, and macroscales, would provide the development of an
adequate theory of the deformation and fracture of solids from the first principles.
From this point of view, nanofracture mechanicst results [37-41] constructed on a very simple
and general fundamental concept in nanoscale, may be extended to micro- and macroscale, if one
uses the experimental facts of statistical self-similarity and scaling invariance of cracking in solid
materials. As a consequence, a deformed solid may be considered as a thermodynamically open,
non-equilibrium system in which the self-organization of dissipative structures takes place. For this
reason, the development of strength theory from the first principles, allowing an adequate treatment
of the response of a deformed solid to an external physico-mechanical impact, is possible on the
basis of the quantum-statistical approach [3, 21, 36, 42]. The self-similarity of dissipative structures
on different scaling levels provides the opportunity for transition from nano- to micro- and
macro-levels. In this case, the plastic deformation and fracture of solids can be considered as the
kinetic phase transitions related to the self-organization of corresponding dissipative structures,
thereby ensuring effective dissipation of external mechanical energy [1-3, 29, 42, 43].
Creation of a statistical theory of the response of a material to an external physico-chemical
action is the supreme problem that would provide a prediction of the behavior of materials and
structures without resorting to special experimental tests. Of special interest are the questions
concerning the influence of scaling factors in penetration mechanics [3, 44-47] due to the highly
non-equilibrium processes of irreversible deformation and fracture.

2. M U L T I F R A C T A L ANALYSIS IN MATERIALS SCIENCE AND


SOLID M E C H A N I C S
Although the objects called fractals with fractal dimensions have been known in mathematics
for a long time (the Cantor set is a classical example of this type, see Fig. 1 from [12]), the concepts
of fractal geometry began to penetrate in physics, materials science, and mechanics only after B.
B. Mandelbrot had his famous book published[i]. In recent years, we have observed the
exponential growth of publications devoted to the fractal analysis of different natural objects and
phenomena, from polymer molecules and Brownian motion to galaxy accumulations and geody-
namicai and astrophysical phenomena [1-3]. The concept of "fractality" of a physical vacuum has
also been introduced in physics [20].
It should be noted that, unlike regular mathematical fractals such as Cantor sets, Koch figures,
or Sierpinski carpets (Figs 1-3), natural fractals, e.g. colloid aggregates, aerogels, clouds, polymer
molecues, porous media, rocks, cracks, and fractures exhibit only a statistical self-similarity, which,
in addition, takes place in a limited range, L, of spatial scales, where L0 < L < L,, (Figs 2-4). Most
of the objects mentioned above can be represented by a series of fractals of different dimensions

tThis name was suggested by John Gilman for the authors' efforts. Also, John Hutchinson suggested the term,
"microfracture mechanics", for them. The original name given by the author was "quantum fracture mechanics".
Fractal fracture mechanics 999

iIIIIiIi
i I I I
I I I I I I I I
iI II II II II iI II II
IiU Illl Illl IIli IIII Illl Illl Illl
Fig. 1. Cantor sets (fractal dimension, dI = In 2/ln 3).

weighted with different weights. Such multifractals are characterized by the spectrum of Renyi
dimensions [12]:

In ~_~ P~(E)
imo j, q~-I

d~=- lim~ I ~ lnE , q=l (1)

where Pi(E) is the probability that a given structure point belongs to the ith element of a structure
coverage by d-dimensional cubes of face, E. For uniform Euclidean objects, the identity of
dq = dr = d holds, whereas for regular fractals, we have dT< d,~= dn <~d [1]. In the general case, the
family of dimensions, dq, possesses the property of monotony;
dq >/dq +,. (2)

Multifractal structures are characterized by the spectrum of the singularity indexes [12],
In Pi(~)
di - In c ' (3)

which represents a kind of "local dimension" of the structure near the center of the ith cell covered
by d-dimensional cubes of volume, Ea. It is evident that the value ~i(¢) depends on the adopted
covering scale, ~. Let N,(~, 6#) be the number of cells with the scale, E, in which the singularity

n=l

n=2

n=3

Fig. 2. The Koch curve (d/ = In 4/In 3).


1000 G . P . C H E R E P A N O V et al.

(a) (b)

Fig. 3. The Sierpinski triangle as a geometric fractal (a) and that corresponding to one of the regular
geometric scales (b).

index, ai(¢) takes a value in the range (a, a + b~). For multifractals, N,(o~, 6~) may be represented
as [1]

N,(o~, c5c¢)= W,(o~)6c¢[1 + 0(1)1, ~ 0 , (4)

where

pc, (~) = p (~, E)E ~=~ (5)

is the distribution density, and p (~, ~) is a non-negative, slowly varying function that may depend
on ~. In the latter case, the value of In p(~, c)/ln(1/c) tends to zero at ~--*0. The exponent f(c¢) in
eq. (5) will be a non-negative, E-independent function satisfying the relationship [12].

d d%
dq [(q - 1) d,] = o%+ ~-q [q --f'(~q)] = aq. (6)

The maximum of f ( a ) at point, ~0, where f'(~0)= 0, corresponds to the fractal dimension,
do = a, =f(~0).
Let us consider, as an example, a continuous binomial measure, B{dx}, defined on the
binary-rational half-intervals [i/2 n, (i + 1)/2 ", 0 ~< i ~< 2" - l], by the following relationships:

B{[0, 1/2)} = 1 - p , B{[1/4,2/4)} = (l - p ) p ,

B{[l/2, l)} = p , B{[2/4, 3/4)} = (1 - p ) p ,

B{[3/4, l)} =p2, B{[0, 1/4)} = (1 _ p ) 2 etc., (7)

where p is an operator whose value lies in the range, 1/2 ~<p < 1. The measure, B {dx }, continuously
proceeds from the binary-rational intervals to all measurable Borrell sets. The measure, B {dx }, is
characterized by the distribution function, B ( x ) = B { [ O , x ) } , with 0~<x ~< I. The following
recursion relationships obviously exist:

where C ( x ) = 1 - B(x). It may be seen that B(x) is a continuous, increasing nondifferentiable


function on the unit segment (Fig. 5).
Fractal fracture mechanics 1001

(a) (b)

0~,~ 0 'O'q
~,)_0 0 ( 0
0 O0 ~0 °°° 0
toO$(


(c) (d)

Fig. 4. Fractal geometry of the fracture of a solid: (a) a crack front showing fractal properties in brittle
fracture; (b) a crack surface showing fractal properties in ductile fracture; (c) the crack path showing
statistical self-similar in quasibrittle fracture; (d) the scaling law observed in a crack size distribution by
quasibrittle fracture.

0.8

0.6

t~
0.4

i
0.2

0.0 0.2 0.4 0.6 0.8 l.O

Fig. 5. Singular binomial distribution function, B(x): (I) p = 0.3; (2) p = 0.7; (3) p = 0.85, according to [2].
1~2 G.P. CHEREPANOVetal.

(a) (b)

1.01
i321 1
2
3 A 1

Fig. 6. (a) Generalized dimensions, dq; and (b) function,.f(~) for the binomial m e a s u r e (I) at p = 0.85;
(2) p = 0.7; and (3) p = 0.6, according to [2].

If we use the binary-rational points, ~, = 1/2", for the coverage, then


p,(E,,) = p ' ( l - p ) " - ' , (9)
where the number of different points with the same i is equal to c,i = n!/i!(n 1)!. In this case,

~pq(E) = ~ C'.[p'(1 - p ) " ']~ = [pq + (1 -p)q]", (10)


i i=0

and the generalized Renyi dimensions are defined by the relationships,

nin[p"+(l-p)"] f lnD( 1~+


- q )(Il n-p)q]
2 ' q #1
dq=lim,,~ ( l ~ q - ~ U c , ~ = plnp+(1-p)ln(l-p) (11)
ln2 , q= 1

The plots of dq vs q for several p values are shown in Fig. 6(a). The functions corresponding to
f(~) are plotted in Fig. 6(b).
The most informative are the first three Renyi dimensions, which have a clear physical sense:
do = dn = d/is the fractai (metric) dimension, d~ = d~ is the informative dimension, and d2 = d, is
the correlation dimension. By definition, it follows that
M(Q
In Z p~(E)
d~ = - l i m In M(E) d, = lim e=
,s0 lnc ' ,40 InE (12)

,. I(E) Mc,I
d/= --um--, I(E) = - ~ pi(~)lnpi(c)
~0 In i=l

Here, M(E) is the minimum number of d-dimensional cubes of face, ¢, required to cover all the
structure elements; I(E) is the Shannon entropy that is the quantity of information needed to
specify, with the accuracy, E, the state of the system (multifractal structure), and C(r) is directly
proportional to r-"', where ~,.---d- d, is the correlation function of the multifractal,
1
C(r) = ~ p(ri)p(r i + r). (13)
Sr "7
The density, p, is equal to unity in the occupied cube of the coverage and zero if the cube is empty.
As follows from eq. (2), the inequality, d, ~< d~ ~<dr, is always valid.
Fractal fracture mechanics 1003

In practice, fractai analysis in fracture mechanics is used to study the crack surface and
pore space, the spatial distributions of pores, inclusion and cracks, the size distribution of
defects, etc.
One of the most widely used methods of determining statistically self-similar objects is the
study of the low-angle scattering of X-rays or neutrons. This technique allows one to study the
structures with lengths from 0.5 nm to 1 #m [1, 12]. The intensity of the low-angle scattering by
statistically self-similar structures as a function of the transferred momentum, k, is proportional,
within the framework of self-similarity law, to k', where the exponent x, defines the fractai
dimension that relates the mass of the object to its size. In the case of scattering by three-dimen-
sional objects which have fractal surfaces, we have x = 6 - ?, while for porous materials, x = 7 - V,
where ? is the index specifying the size distribution of pores;

N ( r ) ~ r ~, (14)

where 7 = ds+ 1.
Another method, the adsorption probe technique, is based on the comparison of adsorption
isotherms for different substances in the monolayer region [1, 12]. However, the range of molecular
sizes of possible absorbates is quite narrow (from 0.2 to 1 nm), whereas the seemingly prospective
use of highly-molecular probes is connected with the necessity to take into account the change in
the conformation of polymer chains in the adsorption process. It is interesting, in this connection,
to try to determine the fractal dimension of clusters from the distribution curves. Fractals are
typically characterized by the power-law size distribution of the number of particles, N(r) [24]. The
angle of the slope of straight lines, In N - In r, specifies the fractal dimension connected with the
d r value for the crack surface.
Figures 7(a) and (b) show the size distribution of the number of pores for high-temperature
creep in AISI 304 steel. One can see from the data in Fig. 7(b) that the dimension of distribution
increases with time.

(b)

(a)

6 2
oo

12
Z

t o ,2 10
In r
2 \\\2

4 8
"'"7- 12
r (MKM)

Fig. 7. Size distribution of the number of cracks N(r) for high-temperature creep of 304 Grade steel [49]:
(a) the plot of N(r) vs r; and (b) the plot of In N(r) vs In r. 1., = 0.27, t = 139.3 h; 2. ~ = 0.20, t = 76 h;
3. E = 0.13, t = 33.3 h (~ is creep deformation during time, t). The values of dr derived from the slope of
the In N(r) curve are equal to: (1) d1= 2.62 + 0.1; (2) d1= 2.55 + 0.05; and (3) dt = 0.87 + 0.1.
1004 G.P. CHEREPANOV et al.

lg P

Fig. 8. A typical log S-log P diagram.

As for metallographic methods of determining the fractal dimension of the crack surface, they
are also quite complex. In practice, three basic methods reviewed in [48] compete with one another:
(1) Slit Island Analysis (SIA),
(2) Methods of Vertical Sections (MVS),
(3) Fourier Analysis of the Distribution Profiles (FAP).
The first method is based on measuring of the dependence of the area, S, on the perimeter, P, when
plotted on a full logarithm scale (Fig. 8) for fractal islands obtained by cutting the crack surface
by parallel planes. To do this, the samples with the crack surface are mounted into conducting
bakelite and then ground until many small "islands" become visible on the crack surface. Individual
islands have been photographed with high magnification. The photographs are used to measure
the area and perimeter of each island (for example, by means of image analysis technique). From
the slope of the line (Fig. 8) drawn through these data points by least square fitting, one can
calculate the fractal dimension for each crack surface using the equation, dr= (slope + l). This
method requires rather sophisticated procedures in practical use.
The method of vertical solutions consists of investigating the relation between the length of
the profile of the crack surface and the scale of the measurement. To enhance the statistical
significance of the results obtained, the measurement must be conducted for several different
orientations of the surface profile obtained by polishing the vertical sections of the crack surface
on which a composite or plastic is deposited. It has been established that the dependence of the
profile length on the measurement scale, E, obeys the asymptotic law,
L(E) = Loc ('~"-dr). (15)

One of the variants of the VSM is to measure the ratio, RL (c) = L(E)/L', where L ' is the length
of the profile projection on the plane parallel to the crack surface. This ratio characterizes the
surface roughness.
RL(~ ) = Cc-~'("-dr), (16)

where L0 and C are constants. Thus, in the VSM, the fractional part of ds. is evaluated from the
slope of straight lines with the coordinates, log L - log c, or log Rt - log c. Figure 9 shows [49] the
dependence of the surface roughness for the specimen made of AISI 4340 steel (tempered at 700°C)
on the measurement scale, c. From data in Fig. 10, it can be seen that the fractal dimension of
the crack surface, dy, depends on the tempering temperature of steel, that is, on the initial
microstructure [49]. The value of dI correlates with the variations of roughness (RL), shape (Rs),
and local configuration (RF) of the crack surface (Fig. 10). Here Rs denotes the ratio of the
Fractal fracture mechanics 1005

!"5~e

i.4 I- ~*

i.2 - o'~

1.1 -- - ~

1.0 I I I
0.5 l 5 10 50 100 200
• (MKM)

Fig. 9. The profile roughness parameter, RL, vs the measuring unit of the crack surface for a specimen
of AISI 4340 Grade steel tempered at 700°C. (The correlation coefficient, ~t, is equal to 0.991.) The data
are taken from [49].

(a)
zoo

1.80

1.611

1.40
(b)
!.12

i.lo-

1.08

1.06
(c)
0"55/

a~ ,/
I I I I I I
Ioo 2O0 4OO 5OO 6OO 7OO
Ta (°C)

Fig. 10. Some functions of tempering temperature for a specimen of AISI 4340 Grade steel tempered
during 1.5 h, namely: (a) the profile (RL) and surface (Rs) roughness parameters; (b) the fractal dimension
measured by several methods; and (c) the local shape parameter (Rr) of a crack surface. The data are
taken from [49].
1006 G . P . C H E R E P A N O V et al.

2.6

2.5

2.4-

2.3

1.6

1,5-

1.4

1.3 1 I I I I I
100 200 300 400 500 600 700
T a (°C)

Fig. 11. The modified fractal dimensions, D and D~ of a surface profile in terms of tempering temperature
for specimens of AISI 4340 Grade steel tempered during 1.5 h. (The data are taken from [49].)

measured area to its projection, and RF is the ratio of the mean height of the profile peaks to the
mean distance between peaks. Evidently, the relation between Rs and RL can be written in the form,
4
Rs = - (RL-- 1) + 1. (17)
7[

Figure 11 shows [49] the dependencies of the fractal dimension for the modified profiles and
the crack surface of AISI 4340 steel on the tempering temperature, T,,. Those were derived from
the curves, Rs(~ ) and RF(E), for the modified profile surface fractal dimensions, D7 and D/~,
respectively. It may be seen that the fractal dimension of the crack surface depends on the
roughness parameters used in the study.
The determination of the fractal dimension of statistically self-similar crack surfaces by the
FAP method is based on the representation of the spectrum of the profile power (the sum of the
amplitudes squared) by the following equation:
SG) ~ ~B, (18)
where k is the wave vector, and B is the exponent connected with the fractal dimension by the
equation:
B = 6 - 2d,.. (19)
A typical Fourier spectrum of the profile is shown in Fig. 12.
The data collected in the review [48] evidence the existence of the relation between the fractal
dimension of the crack surface and some mechanical properties. The data in Fig. 13 show the
correlation between dy measured by the VSM and FAP methods and the fracture energy under
impact loading for the specimens of martensite-aging 300 steel. Another example of the correlation
between di and some fracture characteristics of a specimen provides the data showing the relation
between the fractal dimension and the parameters specifying crack growth, such as the fracture
toughness, Kic, and the threshold value of the stress intensity factor range, AK,~ (Fig. 14).
Nevertheless, some authors [23] indicate a limited character of the correlations obtained and
express doubts that the fractal dimension of a crack surface may serve as the parameter that
uniquely defines the mechanical properties of materials. This point of view seems true because the
fracture process is the result of the cooperative interaction of micro- and macrofracture processes.
Fractal fracture mechanics 1007

{/3

In k

Fig. 12. A typical Fourier spectrum of a crack surface profile.

This means we need a relation between the fractal dimension of a crack surface and the parameters
controlling the fracture instability in nano-, micro-, and macro scales. Such a relation has been
experimentally found in [4]. It has a universal character for the alloys based on the same metal
and may be written in the form of the following dimensionless parameter, A, o f similarity:

Ao + M = p , , / e , , , (20)

where

E( K~dX)2 (Kzs a.~)2


e,, = (1 + v ) ( l - 2v)' e** - IV, (21)

150 -

100-
3•,• NA
331605

340

t-j

50-

0.0 0.1 0.2 0.3


al

Fig. 13. The impact fracture energy, E, vs the fractal dimension of a crack surface, dr without any
correlation (numbers indicate test temperature).
1008 G . P . C H E R E P A N O V et al.

_j~... 15
7 •
2

g 3

<1 !o

5 I I I I
1.06 1.10 1.14 1.18 1.22

Fig. 14. Correlation between the fractal dimension, d/, and the stress intensity factor amplitude, AK~h,
for some steels of dual microstructure: (1) needle-shaped martensite; (2) net martensite; (3) granular
martensite: (4) rolling steel; (5) different types of heat-treatment.

Here, ay is the yielding stress; W, is the critical stress of the initiation of a microcrack; K1s is the
stress intensity factor characterizing microcrack initiation; E is Young's modulus; v is Poisson's
ratio, K ~ ~x is the maximum stress intensity factor characterizing the mode, I; microcrack, M, is
the coefficient that is equal to zero for quasibrittle fracture and unity for ductile fracture; and D
is the fractal dimension of the crack-tip zone. The dependency of D on P for steels with E = const
and v - const and various ay is shown in Fig. 15. The fractal dimension of the crack-tip zone is
a quantitative measure of crack surface structure. The power law of the size distribution of defects,
the multifractal character of such a distribution, and the statistical self-similarity of cracks and
crack surfaces may be considered as some consequences of the fractal dynamics of a deformed
medium whose phase trajectories represent the stochastic attractor. Thus, a deformation cannot
be a homeomorphism.
For dynamic systems, the conditions favoring the occurrence of trajectories of the strange
attractor are: (i) a global compression of the phase volume, and (ii) the local instability of the phase
trajectories. In real deformed media, the first condition is evidently satisfied since irreversible
deformation and, hence, fracture is necessarily accompanied by dissipation of the elastic strain
energy.

---.-.. . . . . . . . . . . . . .
"• I 000 '
".. ..
\o~ ! / M-0
: "'../
/ ....
- M = 1 ~ ' ~

p..rain p.c. "'~"'~"" • " ,~ '~°'~° o~,~ ° p.~ax

I I "" I ~ , I
105 106 tO7 108
lgP,, (MPa3m)

Fig. 15. The fractal dimension, D, of a crack process zone vs some mechanical properties characterizing
the crack instability for a quasibrittle (M = 0) and ductile (M = 1) mode, 1, crack in a steel with v = 0.28
and E = 2.1.106MPa. The data are taken from [4].
Fractal fracture mechanics 1009

The dimensions that characterize the fractal geometry of a deformed medium are related with
those of the phase trajectories. The latter are connected with Lyapunov stability indexes [33]. For
example, for a three-dimensional dissipative system, the trajectories of the strange-attractor type
are realized, if one of Lyapunov indexes is negative (2~ < 0), the second is equal to zero (22 = 0),
and the third is positive (23 > 0). In this case, the Lyapunov dimension of the strange attractor may
be expressed by the following parameter [33]:

dL = 2 + 23/1211 > 2. (22)

Figure 16 shows the results of the numerical calculations of dL for the phase trajectories of
the dynamic system whose behavior is determined by the mapping.

x=mx +y--xz; y=--x; z=q[z--l(x)x2],

I(x)= {lo',X >x 0<~O (23)

at m = 1.5 and g = 0.2. We see from Fig. 16 that the dimension, d, represents a linear function
of I / l n , in the range 0.01 < E < 0.30. The quantitative least-square fit estimate of the coefficients
of the linear function and the approximation of the relevant straight lines by point, c, showed the
limits of that:

ds= dl = dc = 2.29 _ 0.013,


within the limits of the accuracy of the method. The value of dL(--2.33) estimated from the
calculated Lyapunov indexes (2~ = 0.062, 22 = 0.000, and 23 = 0.187) is close to dr.
The extension of eq. (22) is the relationship between the dimension, dr, of the phase trajectories
of the compressing mapping and characteristic number, A ÷ and A -, of the mapping [33],

lnA +
d,- = (d - 1) + ln(1/A
-----~ ' (24)

where A defines the degree of compression of the element of the phase volume in the compression
direction and A ÷ characterizes the phase volume expansion in the orthogonal direction. For
conservative systems, A +A = 1, whereas for dissipative ones, A + A - < 1. In either case, the

1,'% , / " / 5 /
2.0
°/

/;Pf
///~"
1.5

!.3
0.0
/4' I
0.1 0.2
I
0.3
I
0.4
I
0.5
[In Eq] "l

Fig. 16. The d~ as a function of lnE -~ (at E = 0 , d1=2.306+0.015, d~=2.300+0.013, and


d, = 2.277 + 0.0017).
1010 G. P. CHEREPANOV et al.

characteristic numbers, A -+, can be expressed in terms of stability indexes, 2i, even if they depend
on the step of the mapping [33].

A-+= lim 1272f """ a+l '/', (25)


t~ ~ zrv

which may be rewritten as:

A + - = l"i m {.l / m, k=
~ 12~]}

The right side of this equation is self-averaged because of the law of great numbers, which gives

A + = exp(lnl2-+1). (26)

If A +A- = e - ' < 1, we obtain from eqs (24) and (26),


h
d r = dr + f + ~ , (27)

where
('
h = (ln A +) = Jdx,, +(x,)p(x,),
and h is the K entropy and p (Xi) is the steady-state distribution function on the stochastic attractor.
In the case of small dissipation, when ~7'~ h, the fractal dimension of the attractor is equal to
4"(dT+ 1)-~/h.
Remarkable is the manifestation of fractal mechanics in geomaterials. The earthquake
frequency law established by Guttenberg and Richter as far back as 1954 can be expressed in the
form [2],

lg n ( M ) = a - b M = al - bt lg E, (28)

where n is the number of earthquakes of the magnitude, M ( M t < M < M2), for a long enough time
interval recorded over a long enough area. The constant, a, determines the averaged value of
seismic activity in a given region, b is the fractal exponent, which is approximately the same for
each region, and E is the earthquake energy. Equation (28) is valid in a wide range of magnitudes,
from disastrous earthquakes to the acoustic emission revealed in microcracking [2].
As is easily seen, eq. (28) is equivalent to the condition of the similarity of the n (M) function,
namely, n ~ M b, which means that an increase in m by the factor, 2, causes an increase in "mass"
by the factor, 2b;
n(2M) ~ 2bn(M). (29)

The fractional value of b provides evidence of the fractal character of the earthquake
frequency. For many seismic regions, the "energetic" fractal exponent, b, is equal to about 0.9.
Guttenberg and Richter did not explore the spatial limitations of eq. (28), assuming the
uniform seismicity distribution in each investigated region. However, later studies [2], based on
computer analysis of seismic catalogs, have shown the existence of the relation between the number
of earthquakes in a certain range of magnitudes and the size of the seismic area for earthquakes
with M > 5. A modified frequency law has the form:

ig n ( M , L ) ~ a2 - B z ( M ) + C lg L, (30)
where a2 is a constant for a given region ( - 3 < a < - 5); b z is the "energetic" fractal dimension;
L determines the size of the seismoactive region; and C is the "spatial" factor dimension ranging
for different regions from 1 to 2. If earthquake foci were distributed uniformly throughout the
seismic area for any L values, we would have C = 2. The departure of C from 2 means that the
seismic process is fractal with respect both to energy and area. Moreover, the study of the
Fractal fracture mechanics 1011

Table 1. T h e f ( ~ ) for various seismic regions


Energetic ./(~ )
region Class ~ .f(~0 ) ~min f(~min ) ~. . . . f(~max)
Pamirs 7 2.3 2.0 1.2 0 4.3 0
Tien Shan 8 2.3 2.0 1.3 0.1 4.3 0
Caucasus 7 2.5 2.0 1. I 0 4.4 0.2
8 2.5 2.0 1.0 0.3 4.2 0.6
California magnitudes 2.3 2.0 1.0 0 4.1 0.2
3+6.7

earthquake distributions in time in fixed regions showed that they also obey the similarity law in
the form:

lg x ( z ) = a3+ b3 lg r, (31)

where x(z) is the fraction of time intervals in which an earthquake occurs as a function of the length
of the interval, z. According to estimates for different regions of New Hebrides [2], the value of
a3 varies in a range of 0.126-255. Thus, all the major seismicity features obey the similarity (scaling)
law. This observation allowed T. Chelidze to suggest the generalized fractal seismicity law,

lg n(M, r, L) = A B M + C lg L + E lg r, (32)

where n (M, r, L) is the number of earthquakes of the magnitude, M ( M t ~ M .~ M2), that occurred
for the period, r, in the region of the characteristic size, L.
A detailed examination of three seismoactive regions was reported in [2], see Table 1. The maps
of these regions were divided into the nearly square cells of the area, E2, each, the minimum ~ value
being as small as 3 km. The frequencies of earthquakes in each cell for a selected period of time
were used as sampling estimates of earthquake probability. Figure 17 shows the Renyi dimensions
for different regions obtained in [2]. The corresponding distribution functions, f(~), are plotted in
Fig. 18. In addition, Table ! contains the characteristic values o f f ( ~ ) ; this function reaches a
maximum at ~ = ~0-
We can see that the values of the fractal dimension, d t = do, for all the regions studied are equal
to do = 2, because empty cells virtually do not occur in the coverages under consideration. At first
sight, this fact indicates that the distribution is uniform. But a finer multifractal analysis has
revealed a non-trivial spectrum of singularity indexes that reflect the self-similar structure of seismic
fields. This example demonstrates wider opportunities of multifractal analysis based on the
relationships, eqs (1)-(8), compared with the more commonly used fractal analysis exemplified
above.
The topological features of crack surfaces and fields of defects in real solids can be studied
using methods of percolation theory in solid mechanics [3, 50]. Percolation clusters as typical
multifractals are widely used in simulating various stochastic structures [12, 50]. The topological

t I I I I I
-60 -40 -20 0 20 40

Fig. 17. Generalized dimensions, D(q), for some earthquakes in the Pamirs Tien Shan region [2].
1012 G . P . C H E R E P A N O V et al.

I
! 2 3 4

Fig. 18. Some multifractal characteristics of earthquakes: (I) the 7th and 8th energy class earthquakes
in the Pamirs-Tien Shan region; (2) the 7th and 8th energy class earthquakes in the Caucasus region; and
(3) the same class earthquakes in the California region. The data are taken from [2].

properties of percolation clusters are characterized by a wide spectrum of dimensions linked by the
critical indexes that determine the behavior of the system near the percolation threshold [50].

3. T H E O R Y OF ELASTICITY A N D T H E H I G H ELASTICITY
OF M U L T I F R A C T A L S
In the majority of works devoted to multifractal elasticity, the elastic behavior of self-similar
structures is investigated by methods of numerical simulation [51-54]. Two limiting cases, with
isotropic and central elastic forces, are usually considered in these works [53]. It has been shown [54]
that the problems of the elasticity of polymer networks refers to two different universal classes.
At the same time, according to numerical experiments [51, 52], the elastic stochastic percolation
plane lattices (d = 2) of the size, L less than 5~,, where {, is the correlation radius, are characterized
by the negative Poisson's ratio, v, whereas for the lattices with L > 5~,., we have v > 0. The limiting
values of v at L/~,--+ ~ are universal (Table 2), i.e. independent of the relation between the local

Table 2. Comparison of some fractal properties


Numerical simulation of a two-
dimensional random network of size, L x
L, near the percolation threshold (E, is
the correlation length)
Aerogel SiO 2
Properties L I ~,.---, ~ L/~,---,0 (experimental data)
Connectedness of a Elasticity of the Bonds determining Fractal cluster
random network network is elasticity of the
determined by network are multi-
dangling bonds duplicated
Fractal dimension of Dimension of the R a n d o m walk d1, measured by
an elastic frame, dr geodesic line dimension small-angle neutron
scattering and
D r = 1.1 _ 0.02 D , = 0.67 molecular adsorption
2.3+0.1
Poisson's ratio, v 0.08 _ 0.04 - 1/3 0.12 + 0.08
Analytical calculation
Poisson's ratio 0.01 + 0.01 -0.33 0.15+0.05
v = ~./(d- l)- l
Fractal fracture mechanics 1013

elasticity parameters (used in the numerical calculations of the elastic properties of stochastic
percolation lattices).
Another approach to the elastic behavior of fractals bears on the classical assumption of the
entropic nature of elasticity of elastic materials with a self-similar structure [55]. The classical theory
of high elasticity polymers is based on the assumption that polymer chains follow Gaussian
statistics [56, 57]. When the structures following Gaussian statistics are deformed, the dependence
of entropy on the elongations, 2~ = L~/li, in the d space has the form [56]:

AS = cT 2~-d , (33)
i I

where T is temperature and c is a constant defined by the topological features of the structure.
According to [19], only fractals with the information dimension, di = 2 do obey Gaussian statistics;
this fact explains the discrepancy and experimental data [58].
Besides the above mentioned approaches, an analogy between the elastic behavior of
self-similar structures and that of a random spring, whose size exceeds a certain length scale, has
been noted [55]. Based on the analogy, an heuristic model of elastic fractal deformation was
suggested in [59]. According to this model, deformation of a fractal subjected to external force,
F, occurs only within the scales exceeding the characteristic LF, depending on F. Thus, a new
characteristic length arises in the elastically deformed fractal under the external force, F [19]. This
model was used in [20, 26] to develop the theory of elasticity for statistically self-similar structures.
It is well known that the theory of the elasticity of solids is based on two experimental facts,
namely, Hooke's law, which claims that the relative deformation, Eii= 2i - 1, is proportional to the
applied stress, aii, and Poisson's effect, according to which the deformation, E~i,is accompanied by
the transverse deformation, Ejj= x / ~ - 1 (provided that the corresponding stresses are absent, i.e.
at aij = O, j # i, [34]).
To construct the theory of multifractal elasticity, Hooke's and Poisson's laws have been
replaced by the following two statements [20, 26]:
(1) Deformation of an elastically isotropic multifractal structure subjected to an external
force, F, causes the appearance of the only characteristic length, LF, which should satisfy
the equation,

F- ~LvOU T-~=OS Li JOU OS)


~-~ - T-~F., (34)
where the first summand is the energy term and the second the entropy term.
(2) Under elastic deformation, the multifractal self-similarity is retained, namely, the variation
in the density, p, with 2v is similar to that occurring in the "geometric" change of the
structure size. In both cases, the variations are described by the power law,

p(2~) ~ 2~', ~ = a - d~. (35)


In the case of unidirectional deformation, the change in the size of a multifractal along this
direction, which is equal to 21 = 2,, is accompanied by a simultaneous change in the multifractal
transverse sizes measured by the transverse elongations, 2j = 2±, where j = 2, 3 . . . . . d. From eq.
(35), it follows that

2i = 2i-" or 2± = 2~-", (36)


where

= 1-(d- 1)v. (36 ~)


The coefficient, v, of the transverse deformation of a multifractal structure is thus determined
by its fractal (metric) dimension [20],
In 2± d~
v= In2~ d-1 I. (37)

EFM 51/(~'-J
1014 Ci. P. C H E R E P A N O V et al.

Equations (36) and (37) are also valid in the case of the multidirectional deformation of a
multifractal (in the d space). For example, for biaxial and three-axial deformations of a multifractal
structure in a three-dimensional space (d = 3), we have
2~=(2122) I/(j+') and ),1~=(,~1/],2J,3) I/(1-20, (38)

respectively. Note that the coefficient of transverse deformations, v, coincides with Poisson's ratio,
v0, only in the limiting case of infinitesimally small deformations. According to the postulated
self-similarity in elastic deformation of a multifractal, we have v = const, and therefore, v0 is a
function of ).~.
From Table 2, one can see that the values of 2~, calculated with the use of eq. (37), agree well
both with the results of the numerical simulation of the elastic behavior of the plane percolation
lattices [51, 52] and with the results of the experimental measurement of v0 for aerogel SiO2 [1 I].
When calculating the value of v for percolation lattices, it was taken into account that in the limiting
case of L >> ¢,, the lattice elasticity is determined by non-duplicated bonds, and therefore, the
dimension of the elastic bonds in the limit, L / ¢ , ~ , coincides with the geodesic dimension,
DR = 1.1 + 0.02. In the opposite case, when L ,~ ¢, (but L >> 1), bonds are multi-duplicated and at,.
coincides with the dimension of random walk, Dw = 0.67. To calculate v for aerogel SiO2, the value
of d,- was taken from the data obtained by porometry and low-angle neutron scattering [1 I].
According to eq. (33), the changes in entropy and internal energy due to the reversible
deformation of a multifractal may be represented as:
d
AS=-c~ ~ (2dI-d);AU=-c22~; a,=d-d,, (39)
i=l

where c~ and c2 are some constants. Hence, substituting eqs (36) and (39) into eq. (34) and taking
into account that F = 0 and d~(d - C ) = d , ( d - d,) are valid at 2~ = 2v = 1 [19], we obtain the
following equation:
F=C(d,2f, '-d,[d~-(d-1)]2?",t~,-'a-')~-'-d,.(d-d,.),~-' ', (40)

where C is a constant. Using the stress, a~, instead of Fl and eq. (37), we obtain (in the
approximation d,. "~ d~ = df at d = 3) the following equation [20]:

E {~.11+4"--1 -- 2V(2i- 1 - 2 : - 1)}, (41)


a11-- l + 6 V + 4 V 3

where E = d a ~ / d 2 ~ at 2~ = 1 is Young's modulus. Similarly, using eqs (35)-(39), the functions, a~


of 2~ in the case of an arbitrary deformation of a d:-dimensional fractal in the d space, can be
analogously obtained. In the limiting case of an infinitesimally small deformation, when
[eil= 12i- 11 ~ I, there functions can be represented in the classical form of Hooke's law for
d-dimensional elastically isotropic solids that yield the following equations between the second-
order elasticity moduli of self-similar structures:
2 + 2(d - 2)(d - d~) 2
Cil = G; B = # +- Z
(d - dr) (d - 1) a
d-1
G =/~ = ~ E. (42)

Here, G, C~j, and B are shear, longitudinal, and bulk elasticity moduli, respectively, and 2 and/~
are Lame's constants. Substituting eq. (37), when d = 2 and 3, into eq. (42) provides the expressions
that are identical to the relationships between the second-order elasticity parameters for the two-
and three-dimensional elasticity of isotropic solids. Note that the Cauchy equation, 2 =/~, holds
for the structures whose fractal dimension is expressed through the topological dimension of the
embedding space by the equation,
:+,(2)
d,-= d + l ' V-d2 1 ' (43)
Fractal fracture mechanics 1015

that coincides with the equation for the fractal dimension of clusters formed when an aggregation
of particles in the d space is limited by diffusion. This equation was derived in [60] in the averaged
field approximation.
Experimental data on the deformation of elastic materials possessing, according to [56, 57],
the multifractal structure in the range, 2 < d r < 3, are usually considered in terms of the entropy
theory of high elasticity [57] based on the assumption of incompressibility (v = 0.5) and Gaussian
statistics of polymer chains. In accordance with the classical theory of entropy high elasticity, the
equation for F1 (2,) has the form,
E
F, = -~ (;., - ;t i-2). (44)

However, eq. (44) is in poor agreement with experimental data (see [15] and Fig. 19). As a
consequence, phenomenological or empirical functions, F1 (2~), are used in practice to fit experimen-
tal data [56, 58]. In either case, the required accuracy of the approximation is provided by fitting
parameters that, in essence, do not have a clear physical sense and cannot be found by independent
experiments.
On the other hand, using eqs (40)-(42), we come to the following equation [20, 26]:
E
F, - 1 + 6v + 4v 3 {21 +zv_ 2v2 - ' - 2~.(,+ ,.) _ (1 - 2v)2 (zv}, (45)

which agrees with experimental data without involving fitting parameters (Fig. 19).
It is noteworthy that even in the limiting case of incompressible elastic materials (at dr = d = 3),
we obtain from eq. (45),

E
F = ~-~ (2~ - 2 (zs), (46)

which differs from the classical eq. (44). The latter cannot be derived from the theory of multifractal
elasticity constructed above because two main assumptions, on which the classical theory of
entropy elasticity [15] is based, are not compatible with each other. Indeed, according to [15],

oo//'2

I I 1~31 I t I
2 3 4 5 6 7 8
k

Fig. 19. Comparison of Fl(2) calculated by means of eq. (12) at v = 0.48 and E -- 0.192 MPa (curve 1)
with the experimental data for the extension of a rubber specimen at v = 0.2 and E = 0.2 MPa (curve 2).
Curve 3 was calculated using the formula, F~ = E/3(2 - 2? 2) at E = 0.2 MPa. (The data are taken from
[58].)
1016 G.P. CHEREPANOV et al.

Gaussian statistics are valid only for the structures with the information dimension, d l = 2, whereas,
the incompressibility condition, according to eq. (37), holds only in the case, d~ = d = 3. Because
the condition, 0 < d, ~< d/, is always valid [12], from eq. (40), it follows that at dt = 2 and d,- = 3,
the condition, F = 0, does not hold at 2i = 2~ = 1.
For a variety of actual materials with a multifractal structure, such as composites, porous
materials, and geophysical media, the approximate equation, d,.---dl ~-dI, is not applicable [1-4].
In this case, using eqs (40) and (37) in the region of nonlinear elasticity, one can find the
information, correlation, and fractal dimensions of the multifractal structure from the experimental
measurements of the second-order and the third-order elastic moduli. Note that the results obtained
in such a way allow comparison with the results of determining d,, dl, and d~ by independent
methods considered in [1,4, 12].

4. M E C H A N I C S OF FRACTAL CRACKS IN S O L I D S
As mentioned above, cracks formed in the fractures of actual structural materials and rocks
do not resemble those idealized geometric models of cracks with smooth edges, which are usually
considered in traditional fracture theories. As a rule, the crack front and the crack surface have
a quite irregular structure that is characterized by the roughness of different sizes. With the
exception of some specific cases, a crack surface and front visually look like a chaotic structure
that does not display any kind of order. Nevertheless, it was experimentally established that they
possess the statistical self-similarity in a wide range of spatial scales, L0 < R < L,,, where L0 and
L m a r e the lower and upper bounds of the self-similarity [I-4]. In the general case, the Hau-
dorff-Besicovitch dimensions of the front, du, and surface, DE, of a crack are different. The
structure and geometry of cracks is substantially determined by the fracture mechanism under a
given loading.

4.1. Plane problems


To begin with, for simplicity, consider the propagation of cracks in a two-dimensional (d = 2)
lattice that, up to fracture, undergoes elastic deformations solely. Let a crack propagate in the
lattice under a load directed along the x axis so that, ( y ) ~ 0 along the crack trajectory, as shown
in Fig. 20.
It is well known that, for a crack with smooth edges, when dH = dr = 1, the energy of elastic
strain released during crack growth will be concentrated at the crack tip and will dissipate there,
being expended on the formation of a new surface of the crack. When a fractal crack with a
multifractal hierarchical structure propagates in an elastic lattice, the energy dissipation process
is more complex and determined by the structure of singularities of the elastic field in the
neighborhood of the crack front. Generally, a variety of models can be suggested to describe this
process for particular fracture mechanisms. However, in the case of the brittle or quasibrittle
fracture of an elastic lattice, it would be natural to suppose the propagation of a crack tip as being
accompanied by the cascade process of the elastic energy transfer from the greater scales to the
smaller ones, up to the lower boundary of self-similarity, L 0, where energy dissipates, being
expended on the formation of a new surface of the crack.
Because the cascade process occurs with the energy conservation, the energy of elastic
deformation, AU,, (released by the moment when the characteristic size of the crack along x
increases by AR) will be compensated for by the increment in the surface energy, AUs, of a newly
formed crack surface:
AU,, = AUs. (47)
Because
0-2
AU,....-fERAR and A U s ~ yDHRnu-IAR, (48)

where R is the half length of a crack, a is the value of applied stress, E is Young's modulus, and
y is the specific surface energy, from eq. (47) it follows that:
0-2R2 - n...~ DnEy. (49)
Fractal fracture mechanics 1017

l ot

..tdxt.. ~
X

l tlr

Fig. 20. A fractal crack reproduced from [20].

Hence, we have:

(50)

where r/I is a dimensionless factor. F o r a s m o o t h crack, when D , = 1, eq. (50) is reduced to the
Griffith equation.
The stress intensity factor in this case is equal to [24]:

K;- D~ 2 (ot=d-D,=2-DH), (51)

which, in the limiting case o f a s m o o t h crack, when DH = dr = 1, goes over into the k n o w n
equation:

= (52)

E q u a t i o n (51) m a y be written as:

KI = K;o(rcR ) aT - t)n a (53)

where Kt0 is the coefficient o f stress intensity for a s m o o t h crack. Thus, we can see that for a fractal
crack with dimensions satisfying the inequality, 1 < D n = D r < 2, a factor appears in the expression
o f Kt, which, within the limits o f self-similarity, L0 < R < L,, m a y be written in the form:
0,I
f(R) = , (54)
1018 G . P . C H E R E P A N O V et al.

where D r is the fractal dimension of a crack front. This means that, as D I increases (which is visually
equivalent to the increase in irregularity of the front), the stress intensity factor decreases. In this
case, In K; linearly depends on the difference, dr - dr, that may serve as a measure of the crack
surface roughness (see Section 2). The latter conclusion agrees well with the results of experimental
studies on steels of different composition and initial structure.
To find the order of the asymptotic distribution of stresses and displacements near the tip of
a fractal crack, one can employ the method of invariant or path-independent integrals [6]. The
potential strain energy of a lattice with a fractal crack under the loads, T °, applied to the surface,
Er, of the crack, is equal to:

U= fs W(~°")dS = fz T°U°dE' (55)


T

where S is the lattice domain, W is the specific elastic energy, E,,,


0 are the strains, u0 are the
displacements, and ~ = a°.nj.
When the crack surface increases by AZ, the displacements, strains, and stresses gain the
increments, Au;, AE;j, and Aao, respectively. Following [6], we find:

AU = ~1 fa z T~;Au;dE, (56)

where AU is the increment of potential energy due to the crack surface increment, AY,. When the
size of a crack along the direction, x, increases by a certain quantity, r, the length of the crack
increases by AS ~ ro. The value of AU averaged over all possible realizations of the crack surface
increment can be represented by the integral:
1 ~e+,
(AU) ~2JR a2;(x)[u+(x +r)-- U?(x + r)] dx, (57)

where a2; and u + are the mean values of stresses and displacement near the crack tip.
Provided that the asymptotics, a;j ~ R -~ and u + ~ R ~-~, are valid, ( A U ) can be estimated
as:

(AU) ~ r 2-2e,

which provides

(AU) r2-2~ o, (58)


(As)
where ( A U ) / ( A S ) is the mean increment of the potential energy over the mean increment of the
crack length, which is equal to the energy flux to the crack tip. Taking into account the condition:
AU
F = = const, (59)
AS

adopted in the theory of brittle fracture[6, 36], one obtains from eq. (58) the equation,
2 - 2fl = Dn, and therefore, on the basis of eq. (57), one has:

tro = Kix-~/2f.tO ,, ui=--~Ktx°u/2gi(O), (60)

where
R
X=Loo, ~=2--DH.

In the case of a smooth crack, when Dn = dr = 1, the classical asymptotics follow from eq.
(60). As seen, in the neighborhood of a fractal crack tip (1 < DH = Dr ~< 2), stresses decrease more
slowly than for a smooth crack.
Fractal fracture mechanics 1019

4.2. Three-dimensional problems


The results obtained can be extended to the case of cracks in three-dimensional bodies. We
consider the following two types of fractal cracks: (!) cracks with a fractal surface approximately
disk-shaped with disk radius, R, for which the fractal stress intensity factor has the form:

R
gt,,=rl2ffx (3-D1)/2, where x = ~ 0 , and 2~<D/~<3, (61)

where ~/2 is a dimensional factor. (2) Cracks with a smooth two-dimensional surface (Dn = dr = 2),
but with a fractai front, the fractal dimension, ds, of which, within self-similarity (L0 < R < Lm),
varies in the range, i < dye< 2. For these cracks, the equation for K~. may be written as:

K1 = ~13ax ~2- dI ~/2 (62)

where r/3 is a dimensionless factor. The preceding results refer to the case of the propagation of
a "brittle" fractal crack in a linearly elastic body.

4.3. Power-law hardening material


It is well known [6] that the stress and strain distribution near the crack tip of an
incompressible power-law hardening material has a power-law asymptotics. When a mode, L crack
having a smooth front (dH = 1) and a smooth surface ( D . = 2) propagates in such a material, the
stresses and strains in which are related by the power law, a ~ c", the near-crack-tip asymptotics
of stresses, aij, have the following form [6]:
a
ai~ = K I R - " f j ( O ) , where n - 1+ a . (63)

Here, Kt is the stress intensity factor for a power-law material, f j are some functions, and R is the
distance from the crack tip.
Comparison of eq. (63) with eqs (60)-(62) shows that an increase in the surface dimension,
DI, or in the front dimension, di, of a crack is equivalent to the decrease in the exponent, a, in
the hardening power-law.

4.4. Cascade process o f the elastic energy transfer


When the fractal crack tip grows by AL., in the macroscale, L.,, the elastic energy,

A U = Go AL.,, (64)

is released. According to eq. (60), the energy, Go, is density proportional to:

G 0 " ~K- ~~ L- ., ,- ~ , (~=d-O.). (65)

The growth of a fractal crack by the amount, ALto, in the macroscale is connected with the growth
of the fractal crack in the smaller scales, L., by the power law, L. = L.,/x", where x = L,,/L.+, is
the self-similarity parameter. (It is assumed here that the fractal is considered within the bounds
of the self-similarity law, i.e. L0 < L < L..) If x = const in the scale, L., each fragment of the crack
increase by AL. = A L e x p. Therefore, if all elastic energy is expended on the formation of a new
surface, we obtain: "

A U = N.G.AL,,, (66)

where N. is the number of fragments of a crack with the scale, L., and G. is the density of released
energy in the n th scaling level. The cascade process of the energy transfer is described by the chain
of equations that express the energy conservation law in the transition from the n th to the (n + 1)
scaling level.
A U = GoLm = N1Gi ALl . . . . . N.G.AL.. (67)
1020 G . P . CHEREPANOV et al.

From eq. (67), one can obtain the equation for the function, G ( L , , ) = G", i.e.

(68)

the solution of which is:

G ( L , ) ..~ L~, p = D/- (d- I). (69)


Therefore,

Go = G(Lm) "~ L p. (70)


By comparing eqs (69) and (70) with eq. (65), we obtain:

D t = Dn, (71)
i.e. the self-similarity dimension coincides with Hausdorff-Besicovitch dimension of the crack [24].
When the transition from one level to another is accompanied by a dissipation of elastic
energy, for instance, due to plastic deformation, eq. (67) is no longer valid. However, if the
dissipation in the cascade elastic energy transfer from the greater scales to smaller ones is
self-similar (similar to the process of splitting vortices in turbulent flows), the energy conservation
law taking dissipation into account may be written in the form:
(1 - q ) N , G , , A L , = N,+ i Gu+ I AL,+ i, (72)
where ~/ is the portion of energy expended for dissipation in the transition from the nth to the
(n + 1)th level. In this case,
ln(l - q )
Dr-= Dn l n ~ ~>Du, (73)

so that the dissipation of energy causes an increase in the fractal crack dimension, in accord with
the empirically established facts [1-4, 49].
In the special case when the dissipation is the only reason for the fractal character of the ductile
crack, the fractal dimension of the crack is given by the following equation:
ln(l --r/)
Dr= 2 in x (74)

Since Ds~< 3, from eq. (74) we obtain:


r/~< 1 + x - t . (75)

At higher r/values, a crack growth is impossible if the energy absorption is too high, namely if
> 1 - (l/x).

5. FRACTURE SIMULATION IN LATTICES


In computer simulation of fracture of plane (d = 2) triangular lattices [61, 62], the elastic lattice
deformation was described by the Lam6 equations:

(2 + #, O~(~ 0juj) + # ( ~ 0})u~ = 0, (76,

where u; is the displacement component and c~ is the partial derivative with respect to the ith
Cartesian coordinate. Propagation of a crack after the break of the first bond was assumed to be
governed by the following rule: the probability of the breakage of a bond between the neighboring
ith and jth atoms is proportional to the ruth power of the stress,
Pq ~" (ffij/ffc) m, ffij ~ tr,. (77)
Fractal fracture mechanics 1021

Table 3. Comparison of fractal dimensions calculated using analytical equation, d t = I + v"', with those computed
numerically based on a two-dimensional network with 2 = g(v = 2/3) and d = 2
Effective fractal dimensions obtained from the crack growth
models
A typical crack
The exponent, m, in (broken bonds) Results of the numerical simulation for
the probability of a generated by means different boundary conditions at v = 2/3
bond failure of the stochastic Analytical value
(P, ~ ~7') fracture model Dilatation strain, ~l Shear strain, ts d 1 = 1 + v"

1 Scattered fracture of 2,00 2.00 2.00


an elastic lattice
0.5 Scattered fracture of 1.9 _+ 0.1 1.9 + 0.1 1.82
an elastic lattice
1.0 Multiple fracture of 1.66 + 0.05 1.65 + 0.05 1.67
an elastic lattice
2.0 Growth of a fractal 1.45 + 0.05 1.40 + 0.05 1.44
crack
Propagation of a 1.00 1.00 1.00
smooth crack

In this case, the fractal dimension of a self-similar crack configuration does not depend on the
loading mode (defined by the corresponding boundary condition, see Table 3), and is a function
of m and Poisson's ratio, v (Fig. 21).
From the data presented in Fig. 21(a), it follows that ds(v ) at m = 1 is a linear function of
the dimension, d, of the field of non-uniform deformations [20] namely:

as= (d - l)(l + v).

This equation is identical to that of the fractai dimension, d~, in terms of the coefficient of transverse
deformation for self-similar structures.
Due to shear stiffness, non-uniform density fluctuations in solids are necessarily accompanied
by the appearance of shear stresses. Besides, the minimum scale of thermodynamically stable
density fluctuations appreciably exceeds the interatomic spacing. Hence, the spatial distribution of
non-uniform fluctuations usually exhibits the scaling invariance. The latter fluctuations are
spontaneous quantum or thermal ones, as well as those induced by external forces or by fast
changes in the boundary conditions, as in the crack propagation model based on eqs (76) and (77).
This explains the self-similarity of a spatial crack distribution in thermofluctuation fracture, as well
as the multifractal character of crack configuration.
Let us consider in more detail the stochastic model of the fracture of an elastic lattice described
by eqs (74) and (75). To derive ds as an analytical function of v and rn, we consider unidirectional
deformation, ~~, ((E ~) = 0), of the d-dimensional lattice subject to the stress, a~ (ds, according to
the results of numerical simulation, does not depend on the loading mode).
Due to Poisson's effect, the transverse stresses, a~ = va~,, arise. The break of the bonds during
the propagation of a crack formed in accordance with the numerical model of [61, 62] is
accompanied by the redistribution of stresses on intact bonds and by the relaxation of stresses in
the region of broken bonds (according to the results of computer simulations). The stress relaxation
results in the dissipation of energy of elastic deformations on a newly formed rough surface of a
crack. To determine dl, we may express the fractal dimension of the phase trajectory containing
the strange attractor in terms of the characteristic numbers, A ÷ and A - , of the compressing
mapping [33] as follows:

I ln + q
4 = (d - 1) 1 - !im ln(1/A -)J' (78)

Taking into account [61, 62] that the breakage of bonds from the ( a ± ) occurs, on the average
(d - l)v m times more seldom than from the (a~) stress (where the sign ( ) denotes the averaging
1022 G.P. CHEREPANOVetal.

(a)

1.8

• 2 •
1.6

1.4

1.2

0.0 0.2 0.4 0.6 0.8


v
(b)

(c)

1.8

1.6
2 \ I
• 0.0 0.5 1~ 1.5
1.4

1.2

2 4 6 8
m

Fig. 21. The fractal dimension, dj, of a crack path, d / o n v; (a) and m; (b) and (c) in terms of: (a) v, (b)
and (c) m. See the data in Table 3 taken from [7].

over realizations), the fractal dimension of a configuration (d = 2) and a surface (d = 3) of a crack


formed due to fracture of two- and three-dimensional bodies, can be expressed in the form:

df=l+v m and d:=2(l+vm), (79)

respectively. F r o m Fig. 21(b) and the data presented in Tables 3 and 4, it follows that the
calculation based on eq. (79) at d = 2 is in excellent agreement with the results of the above
mentioned numerical simulation.
At m = 1, eq. (79) is identical to that for the fractal dimension of the free fracture surface of
d-dimensional solid derived in [20]. The dynamics of free fracture is controlled by the cascade
process of the elastic strain energy transfer from the large scales, L,, to the smaller ones, with the
parameter of self-similarity, x = L/Li+~, being constant. Elastic strain energy, Ue, stored in a body
can be completely expended on the formation of a new fracture surface with the total area,
~, = U,/~, only if the surface is smooth and its Hausdorff-Besicovitch dimension is equal to
dn = d - 1. F r o m eq. (79), it follows that this is possible only for the free fracture of bodies with
v0 = 0. In the free fracture of bodies with v > 0, a portion r / o f the stored potential energy of elastic
deformations is inevitably lost. If a body experiences only elastic deformations until it is fractured,
the energy loss, UR= r/U,., is determined by the relaxation of elastic stresses on the rough
Fractal fracture mechanics 1023

Table 4. Mechanism of the fracture of solids simulated for various values of m in the stochastic model of fracture
of an elastic lattice, the topologically-equivalent classes of percolation, and the kinetic model of the fracture of solids
The class of percolation
models that is The w in the kinetic
The m in the probability Simulated mechanism of topologically equivalent model of the fracture
of bond failure (Pi ~ ~") the fracture of a solid to the stochastic model of solids
•~ 1 Ductile fracture Anisotropically- < 3/8
correlated percolation
0.5 Quasiductile fracture (a Anisotropically- 3/8 < w < 1/2
crack with a fractal correlated percolation
surface)
1.0 Free fracture Chaotic percolation > 1/2
2.0 Quasibrittle fracture (a Anisotropic percolation > 1/2
crack with a fractal
front)
oo Brittle fracture Anisotropic percolation
(Gritfith's crack)

(self-similar) crack surface. In this case, the fraction of elastic deformation energy lost in free
fracture is given by:

r/ = 1 - x -~, (80)

where fl = ds - (d - 1) = (d - 1)v0, and the self-similarity parameter, x, is also defined by Poisson's


ratio [14, 20].
2(1 - v)
x 1 - 2v (81)

Equation (81) agrees well with the empirically established range of x for various materials [2]. The
case of a Gritiith's crack with a fractal surface is somewhat different [22]. For the definition of
fractal, it follows, that the number, M, of broken bonds in the main crack formed during fracture
of a body is proportional to:

M ~ R~, (82)

where R is a characteristic size of the body. On the other hand, using the asymptotics in the
stochastic fracture model [61, 62], we obtain:

m ~ N P r ( R ) ,,~ R 3 R - ° s ~ ' , ~ = d - ds. (83)

Equating the exponent, dr, to that in eq. (83), we find out that the dimension of a crack formed
in the model [2, 3] coincides with the dimension of the fractal Griffith's crack, provided that m = 2.
If the only mechanism of energy dissipation is the relaxation of elastic strains on the self-similar
fractal crack surface, the following equations are valid [7, 14]:

dr= 2(1 + v2), r / = 1 - x -a, /~ = 2v2; Kt-~- ox(I-2v2)/2. (84)

Here, ds is the crack surface dimension, q is the portion of lost energy, and x is the small distance
from a crack tip. If a / > a,., the classical Griffith's crack with a smooth surface develops, i.e. d = 2,
r / = 0 , and K I = av/-x.
Fracture of real bodies is preceded, as a rule, by plastic deformations [4] ensuring a relaxation
of elastic stresses and an accumulation of damage in the region of irreversible deformations [15].
Therefore, the relative change of the volume of a body in irreversible deformation can be
represented by the sum, 6v = 6, + 6, +_ 6s, where 6, = (1 - 2v0) a,./E is the elastic constitutent, 6, is
the deformation, and 6, is the relative increase of the body volume caused by the formation of
nano-, micro-, and macro-damage.
On the other hand, according to [15], the relative change of the volume is equal to:

O'ef
t
,~v = (l - 2vow) ~ - , (85)
1024 G.P. CHEREPANOV et al.

where E = d o d d c is the tangent modulus, and v~ is the effective value of the coefficient of
transverse deformations. Hence, when ac~/E = ~/E [15], the expression for the effective coefficient
vc~ of an irreversibly deformed solid will be:
v0 + 0.56 6~ _ 6r
ve~--- where 6 = - - (86)
1 +~ 6v
Here, the sign " + " corresponds to compression, and the sign " - " corresponds to extension.
Although the parameters, v and vc~, are some functions of the process of irreversible deformation,
each fracture mechanism is probably characterized by a certain sign and a certain absolute value
of 6(v), and therefore, by dr(v) given by the equation:
a t = [1 -4- vc~(6*)]. (87)
Using eq. (79), one can imitate the topological feature of various fracture mechanisms and models
in terms of the stochastic model [61, 62] (see Table 3).
Topological fracture characteristics can also be imitated by the percolation models [7, 14],
which consider the formation of a main crack as a percolation process of crack clustering on the
prefracture level. One of the fundamental properties of the percolation process is that the
percolation threshold variance (which determines the ultimate strength, a,) depends universally
upon N:
6(a 2) ~ N -l/kdt, 0.9 ~<k ~< 1.3,
where N is the number of simulated nodes (cracks) N [2]. Taking into account the hierarchical
structure of cracks in solids, it can be shown that

6(a")=x/6(a~)a, (N, 1)~2kJl)-~= (L~-sp)L"-'~2,) I (88)

where L~ is the specimen size.


The most interesting consequence of eq. (88) is that the oscillations of the strength depend on
L~p, which reflects the hierarchy of the structural fracture levels [14].
It is well known that the mechanism and kinetics of nano-, micro-, and macro-fracture may
be different t3, 4, 15]. It means that, in the general case, dr will be a function of the scale, L~, which
determines the multifractal character of crack surfaces and manifests itself in the existence of
non-commensurate scaling phase transitions in fracture [14, 44]. According to the D theorem [21],
ds(L,) ~ dr(L,+ ~), i.e. cracks are "smoother" and the fraction of dissipated energy is lower on the
larger scales. This conclusion is confirmed by the experimental data [1-4].
According to [48], the observed manifestation of scaling effects may be connected with (1) the
dependence of the strength characteristics (yield strength, ultimate strength, endurance, etc.) on the
specimen size, provided their geometrical resemblance is retained; (2) the dependence of the
strength characteristics on a specimen size, and (3) the change in the behavior of solids with
increasing or decreasing specimen size.
The classification of possible scaling effects is given in [48]. It should be stressed that the
existence of scaling phase transitions in the deformation and fracture of solids makes the possibility
of adequately simulating the reliability of large-scale structures and geodynamical objects on the
laboratory models quite problematic [44].

6. NANOFRACTURE MECHANICS AND FRACTAL KINETICS


OF FATIGUE CRACKS
It is an important implication of fractal fracture mechanics that the latter suggests a way for
the prediction of fracture and fatigue phenomena in a scale using the corresponding data for
another scale. So, nanofracture mechanics [37-41] providing a strict approach to an irreversible
deformation and fracture in nanoscale may be used to derive the mechanics of fracture and
plasticity in micro- and macro-scales from nanofracture mechanics by means of the power-law
self-similarity and scaling invariance of fractal cracking. We should not overestimate the
possibilities of this approach because of the many dangers of the empirical and statistical nature
Fractal fracture mechanics 1025

of the predictions of fractal fracture mechanics. (For example, prior to an experiment, we do not
know whether a specific case under study enters into the range between the lower and upper bounds
of self-similarity, or not.)
However, we cannot neglect this general fractal approach in view of its simplicity. Besides,
the information obtained in this way is sometimes unique, especially for a new material that has
not as yet been tested. Therefore, we provide here some basic information [3741] on fracture in
nanoscale (that is, roughly in the range from l nm to l # m realized inside the grain of a
polycrystalline material). The brittle vs ductile behavior of a crystalline grain in nanoscale may be
specified by a single parameter, q, called the brittleness number [39].

--
x-;, - - ~ 1
oao
- (89)
2~/G~,/(1 - Vo) 7

Here, K~ is the critical value of the stress intensity factor corresponding to the emission of the
first pair of dislocations from a mode I crack tip (that is, the initiation of plastic deformation),
7 is the true surface energy corresponding to the crack growth without any dislocations emitted,
G is the shear modulus, v0 is Poisson ratio, a0 is the interatomic spacing, T0 is the Schmidt constant
for a given crystal depending on a slip plane, temperature, and purity of the crystal, and ~ is
a number depending on a crystal lattice type (for a cubic lattice ~, = 2.236). The value of K~, is
equal to:

K~i = ohx/Groao/(1 - Vo). (90)

Here, 72 is a number depending on a crystal lattice type (for a cubic lattice, ~2 = 5.149).
If r / > 1, a crack tip cannot serve as a source of dislocations, so the fracture will be ideally
brittle. If r/ < 1, some edge dislocations are beirg generated from the crack tip under loading, and
therefore, a slow subcritical growth of the crack occurs. Ifq < 1, but r / ~ 1, the number of emerging
dislocations is comparatively small and fracture is quasibrittle. If

z0L < O%aoG/(1 - Vo), (91)

the crystal specimen displays the property of superplasticity, because the crack growth is impossible
in this case. Here, L is the characteristic width of the specimen, and ~% is a number depending on
a crystal lattice type (for a cubic lattice, ~3 = 0.085).
According to nanofracture mechanics, there is no incubation period for initiating fatigue
nanocracks in metal grains. The growth of nanocracks begins under extremely low loads in cyclic
loading (K~i is the threshold stress intensity factor). The first step on the d l / d N - KI diagram of
fatigue nanocrack growth equals one interatomic spacing per cycle, that is the minimum possible
nanocrack growth velocity, d l / d N (here l is the nanocrack length, and N is the number of cycles).
The second step is equal to 2 a0 per cycle, the third to 3 a0 per cycle, and so on. The minimum fatigue
crack growth velocities were observed while cracks were growing in thin intercrystalline layers [4].
The R-curve of a nanocrack specifying nanocrack growth in terms of the stress intensity factor
(during loading) is shown in Fig. 22, see [40]. We use different scales along the horizontal axis (for
< 0 and ¢ > 0), so that ~ = 0 corresponds to the initiation of a stable brittle nanocrack growth:
For ~ > 0: ~ = - 1 +Al/(a0n~)

V/2(A/--aoni) aoG.i 8.125T


- , r. o= 0.083 ~o(1 v° ), ni=-- (92)
/"n0 -- ~0 "CO

The characteristic points in the nanocrack R-curve are: the maximum point, ~,---0.3
(Kt, = 4.685); the minimum points, ~p = 2.3 (K/o = 0.448); the point corresponding to the initiation
of a brittle crack, ~ = 0 (K/, = 5.149) and the asymptotics, ~ 0 (R'/~0.702).
1026 G . P . C H E R E P A N O V et al.

t~7

I
I
I
I
I

i KIc
I ' °

-1 0 ~ ! 2 ~D 3

Fig. 22. Theoretical R-curve for a cubic lattice in nanofracture mechanics [40].

At 0 < ~ < 4,. and 4 > 40, the nanocrack growth is stable, whereas in the region, 4o > 4 > 4,.,
equilibrium is unstable and only dynamical propagation is possible. Thus, we have the following
criteria of fracture mechanics in nanoscale:
(1) criterion for initiating a brittle nanocrack:

Kt >/K/i = 12.78[G7/(1 - v0)]u2;


(2) a stable growth of a brittle nanocrack before the onset of instability:

Al = 4,.r,/x//2 = 0.146G7/z20(1 - v0);


(3) nanofracture toughness:

Kt,. = 13.35[G),/(1 - v0)]u2;


(4) the minimum fracture toughness specifying the arrest condition:

K,o = 1.25[Gy/(I - v0)]'/z;


and
(5) the nanocrack growth before the nanocrack arrest:

AI = (40 - 4<)" r,,olv/2 = Gr l ~ ( 1 - Vo).

The numerical factors in all these equations correspond to a cubic lattice; they were determined
using an approximate approach and will be improved in the future.
In a topological sense, ductile and brittle fracture differ in that, in brittle fracture, the crack
front is fractal (with the dimension, d:, in the range, 1 < d: <~2), while in ductile fracture, the crack
surface is fractal (with the dimension, d:, in the range, 2 < dr <~3). This difference is determined
by kinetics of the cluster formation of a crack from defects of different size in ductile and brittle
fatigue fracture [27].
The kinetics of clustering defects in fatigue fracture of solids obey the Smoluchowski equation
[27]:
dCj
dt = 2 ,~.=, k ( i , j - i)C~Cj_, - Cj ,=, k ( i , j ) C , (93)
Fractal fracture mechanics 1027

with the kernel of the form, K ( i , j ) ~ (i,j)". Here, Cj is the concentration of clusters involving j
elementary defects. The parameter, w, is determined by the fractal dimension of the ensemble of
defects [27]:
2+0
w = D ~i _ (94)
2,/,,'
where dn is the Hausdorff-Besicovitch dimension of elementary defects and 0 is the exponent in
the equation of anomalous diffusion on the fractal, D ~ r-°.
The Hausdorff-Besicovitch dimension of dislocation pileups in the slip plane does not exceed
the topological dimension of the slip plane surface, dr = 2, i.e. dn ~< 2. In this case, w > 0.5 because
0 is always greater than (or equal to) zero. For example, D = 4/3 and w = 3/4 for the correlation
structures [3].
According to [3], at w > 0.5, the asymptotic solution of eq. (93) with the multiplicative kernel,
K ~ (i,j)", describes the growth of a single "surviving" cluster that may be, in this case, identified
with the main crack. The rate of growth of the main crack, depending on its length, R, is then given
by the equation [27]:
dR d,
~R:, z=l+(2w-l)d,=l+(2+0)d-~H-d,, (95)
d~-
which can be reduced to the following form:
dR
d t ~ K] ~ R"". (96)

Here, a = 0.5 (2 - dr) is the exponent that determines the stress intensity factor for the fractal crack
front, and d, is the correlation dimension of the cluster. Equating the exponents in eqs (95) and
(96) provides:

n__2F=+o
L2 - ~ r \ ~ j
d,-I 1
- ~j, (97)

whence, at d I ~ d, < 2, we find that n > 2. The latter inequality agrees with the results of the
numerical calculations [4].
If dr= d n = 211 - (1 + O/n - 2)],

n=2 1+~), m=n


(o) 1 + ~ (98)

where m is the exponent in the Weibull distribution [4]. Equation (98) is in agreement with the
experimental data of [4, 49].
In the general case, at d r = d,, the dimension of the crack front formed in brittle or quasibrittle
fracture is equal to:
n-I
4 .= dR 2 + 0 + (0.5n -- 1)dn" (99)

At the same time, the dimension of the crack surface in brittle fracture is equal to the topological
dimension, i.e. d s = dSr, whereas in quasibrittle fracture, d s = d i + 1 > 2.
The dimension of dislocations responsible for the rotational instability in ductile fracture of
solids[3,27] is greater than the topological dimension of the surface, i.e. d n > d r = 2. If,
additionally, the stronger inequality,
dn > 2 + 0, (100)
is satisfied, which makes valid the condition, w < 0.5 (i.e. DF > 2), then the solution of eq. (93)
describes the simultaneous growth of a large number of similar clusters (cracks) merging into a
main crack by means of the percolation mechanism.
The dimension of a crack surface by ductile fracture is determined by the fraction of energy
dissipated on the rough crack surface.
1028 G . P . C H E R E P A N O V et al.

7. QUANTUM MECHANICS AND THE FRACTURING OF SOLIDS


The interatomic potential, U(r), is characterized by the following common properties:
(i) the minimum points exists, which corresponds to the equilibrium interatomic distance, a0,
at zero absolute temperature;
(2) the position, r.,, exists where the interatomic force (proportional to OU/Or) reaches a
maximum, i.e. ~2U/c~r2[...... = O.
The energy spectrum for such a potential represents a set of the energy levels, E,. In the main
state, an atom, owing to zero oscillations, possesses a finite energy:

Eo • (101)
a o ~/ m

Here, h is Planck's constant and m is the atomic mass.


The amplitude of zero atomic oscillations is equal to:

(Ar) = aomC,/[mn U(a0)] '/2,

where
~h
C - (102)
ma o

is the maximum speed of the atomic motion in the potential well of U(r) at this energy level. Due
to the uncertainty principle, an atomic energy level, En, has a finite width, &n; in particular, at the
main energy level, &0 = ~r2h2/2mao • Note that &0 is equal to the kinetic energy of the atom, for
which the De Broglie wavelength is equal to ,~8 = 2ao (it corresponds to the speed of atomic motion,
7rh / mao ).
Because of the asymmetrical shape of the potential, U(r), the excited energy levels, En, with
n = 1, 2 . . . . are far from the degeneration, so that the difference in the energies of the neighbor
levels, Aen = E~+~-E~, decreases with growing n. Each energy level is characterized by its stable
atomic configuration and specific interatomic spacing, an. As a rule, an + ~> a,. If AE0 < &0 in a
solid, the solid will be a quantum crystal [3]. For most condensed matter substances, AE0 >> &0, but
the higher energy levels excited, there may be A~, < &n [36].
The interatomic potential is commonly considered by constructing the classical models of
fracture. As a matter of fact, the states of a condensed matter, namely crystalline, quasicrystalline,
amorphous, and liquid, are characterized by a strong correlation in the relative motion of atoms
at the distance, L0, much greater than a0, which implies the shear stiffness or resistance of a
condensed system. Therefore, the response of a deformed medium is determined by the dynamics
o f the collective excitated motion of atoms resulting in dislocations, disclinations, etc.
In order to correctly determine the excitation spectrum (or defects) of a deformed solid, it is
necessary to solve a set of non-stationary equations of stochastic mechanics. The equations
represent the generalization of Schrodinger's equation for a many-particle system allowing for the
vortex motion of particles. The potential, Uij(r) of the system is formed by an ensemble of atoms
that defines the "structural memory" of a deformed solid.
The fundamental properties of the collective motion of excitated atoms can be established
prior to the complete solution of these equations. Probably, it would be sufficient to use the
approximation of the potential relief in the following form:

Uu(r) = U°jf(r), (103)


w h e r e f ( r ) is a periodic function with the period, 2a0, that is defined on the scale of the order of,
L 0, and

~U(r), when [r~- rj[ ~<a0 (104)


U° = [ Uolri- rj[ ~, when ]ri - j l > ao
Here, U(r) is a pit-shaped potential that describes the interaction within the first coordination shell
of the radius, a0. The short-range potential, U(r), whereas, Uolr~- ril -~ describes the long-range
Fractal fracture mechanics 1029

interatomic correlation ensuring the medium shear stiffness specified by the shear modulus, G. The
following properties can be shown in this case [15-30]•
(I) Only a restricted set of atomic configurations can be realized with a significant probability
in the elements of a condensed matter having the volume of the order of, L30.
(2) There may exist excitated motions of both the potential (rot u = 0) and solenoidal
(div fi = 0) type. The latter determines the existence of rotational deformation modes.
(3) There exists the hierarchy of the characteristic spatial scales (of collective excitations in
a deformed medium) described by the ratios,
Ln + i rot rot u
x = L, Igrad div ul' (n = 0, 1, 2 . . . . ), (105)

where u is the material velocity vector• Such an hierarchy manifests itself in the
self-organization of structural levels of deformation and fracture of solids that are not
connected with the initial structure of the material•
(4) Spatial distribution of collective excitations in a deformed medium is self-similar; the
fractal dimension of the field of such excitations is equal to:

dr= 2(1 + veer) (0 < dr< 3) (106)


where ve~ is the effective Poisson's ratio.
(5) The momentum relaxation time, v~"), for individual atoms of the main level (n = 0) and
for collective excitations of the n th level is much less than the time of energy relaxation,
i.e. zl") ~> r~") (z~°) ~ ao/c,, z~°) ~ L°/c,, Lo/ao ~ cx/~/c, >> 1). As a consequence, inelastic
deformation is accompanied by accumulation of excessive energy in some autolocalized
structural excitations.
(6) The limiting value of potential energy of elastic strains that may be accumulated in the
autolocalized excitation of the nth level is proportional to (L,)°i "J, where D(f ) = 2(1 - v~)
is the dimension of collective excitations of the n th level defined by the effective
"structural" value v~ • It is shown that D(")> .f
4/3.
(7) The intensity of entropy export from the highly non-equilibrium regions where an
excessive energy is autolocalized (controlling the kinetics of self-organization of dissipa-
tive structures), is determined by the fractal dimension of the surface of highly
non-equilibrium regions, dr= 211 - v~].
(8) According to the S theorem [31], the self-organization of dissipative structures in open
systems is accompanied by a decrease in entropy and in entropy rate production, which
are normalized to a constant value of the mean kinetic energy• Analog of the S theorem
for the processes of self-organization of dissipative structures in a deformed medium may
be represented in the form of the D theorem [21]•

D(fl)>~D ('+t), d(f)<~ df~"+1) (107)


(9) An analog of the brittleness number in nanofracture mechanics [39] and the Reynolds
number in hydrodynamics is the parameter [36].

/-/0 10
B,, = - , (108)
c~ ¢0
where H0 is the atomization energy (that is equal, for pure elements, to evaporate heat),
and 6 is the velocity of transverse elastic waves.
In eq. (108), the values Of/o = 2aoAs and ~0 = 2aoc]AB/Ho, where AB = 6/c,,, are the
characteristic scales of the atomic momentum relaxation and the radius of elastic
correlations. Thus, B,, is an analog of the Ginzburg-Landau number in the theory of
superconductivity, whereas An is an analog of De-Bur parameter [3] [the value of C,, is
given by eq. (102)].
For ductile materials, B,.r > 1/x/~, while for brittle materials, B,, < l/x/~, which is
confirmed by the data in Tables 5 and 6 (see [36])•

E F M 5116--K
1030 G . P . C H E R E P A N O V et al.

Table 5. Physical properties of ductile materials at t/,~ 1, and B,., > 0.707 according to [3]
Material AI Ti Fe Co Cu Pb
structure fcc hcp bcc foe fcc fcc
a,A 2.86 2.95 2.48 2.51 2.56 3.5
p, 103 kg/m ~ 2.734 4.5 7.87 8.83 9.02 I 1.36
6's , m / s 3235 3100 3228 2553 2388 1100
c. m/s 6794 6038 5751 5414 4833 2420
Ca, m / s 26 15 12 11 13 2.75
ck , m/s 290 212 195 168 172 55
cT, m/s 1400 1130 1060 955 910 365
H o/c: 1750 1620 1287 1348 1000 388
P~., GPa pc,,, c~ 0.48 0.39 0.54 0.53 0.6 0.076
pn~/c~c~ 1.18 1.3 1.2 1.87 1.26 1.10
experiment 0.41 1.0-2.0 0.94-1.4 0.8 0.08
Ho, 103 c, ct 9.5 7.2 6.1 5.3 4.4 0.88
kJ/kG * 11.9 9.8 7.4 7.3 5.3 0.94
8.2 10.6 7.0 5.3 0.85
s,, = no/c? 1.14 1.I 0.71 1.12 0.93 0.78
q 0.02 0.2 0.08 0.03
* Heat of formation.
t Activation energy of fracture.

(10) The response of a solid to a force is different for materials with acr > 1/~/~ and
B,.r < l/x//2 [2]. In the case of a dynamic loading of ductile materials, the shock wave is
formed when the mass transfer velocity, u, exceeds c,. In this case, the mass transfer
velocity in the elastic precursor is equal to c,, and the Hugoniot elasticity limits are:
Pn = pc~ct, (109)
where p is the material density and c~ is the velocity of longitudinal elastic waves. For
brittle materials (B,, < l/w/2 ), we have, according to [36]:
Pn = pckct, where ck = ~ c ~ . (110)
(11) If lu[ > c,, a crystal lattice loses its stability and a crystal goes over to the non-equilibrium
state [36]. As a result, if
Ck < U <Ccr , where c , . , = x / ~ k c % (111)
a brittle material suffers multiple fracture [45], whereas a ductile crystalline material is
deformed in a non-stationary deformation regime [36, 42]. The transition to a stationary
deformation mode occurs when the velocity of the deformation wave exceeds the limiting
velocity of the crack propagation that is equal to:
c,r = x/~,c% = Ho/ct. (112)
The region of the deformation mode has an upper bound, ct.
(12) If [ul >~ c~ and B,., > 1/x/~, a solid material turns to a liquid phase in the front of a shock
wave by the tunneling fusion effect (if u < ct, the material melts only in an unloading
wave). If [ul > el, materials (B,r < l/w/2) undergo dissociation fracture [45]. The results
of calculations of physical parameters of ductile and brittle materials using eqs (101), and
(108)-(112) are in agreement with experimental data (Tables 4 and 5).

Table 6. Physical properties of brittle materials at B, r < 0.707 and r / > 1 taken from [3]
Material Si Mo W SiC B IC AI203
p, 103, kg/m 3 2.33 10.28 19.3 3.215 2.5 3.99
c,, m/s 5510 3355 2904 7906 9857 6401
Cl, m/s 9140 6418 5237 12,516 14,365 10,847
c, = c , / c l , m/s 380 160 150 204 782 700
q , m/s H o / c I , m/s 1840 1040 882 1600 3350 2637
experiment 1800 900 1600 3200 2500
p,., GPa p c , , cl 7.9 10.5 15 8.2 28 30.3
experiment 7.6 10. l 12 8.3 18 2I
B,., 0.52 0.6 0.55 0.32 0.6 0.7
q 1.3 1.7 1.6
Fractal fracture mechanics 1031

4I

3- . , 8

I I I
0 ! 2 3

Re-~

Fig. 23. The H,./G in terms of B,rrI for the materials numbered as follows: 1. Nb; 2. V; 3. AI; 4. Ti; 5.
Cu; 6. Ag; 7. Pb; 8. Mo; and 9. W. The dashed line shows the results of the theoretical calculation. (The
experimental data are taken from [64].)

(13) The following equation was derived in [14]:

Hv = (1 + 2B,7~)G, (113)
where G. is the yielding stress, and H,. is the Vickers hardness. The results o f the
calculation using eq. (113) agree satisfactorily with some experimental data (Fig. 23).
(14) The generalized equations for the mass, m o m e n t u m , and energy (heat) can be written in
the form:

Ot ~ = D~Au, (114)

where 3~/Ot ~ is the fractional derivative with respect to time and fl is determined as
follows:

~, -- 1 + ~ (")
Lye,. (! 15)
(D~ is the specific diffusion or thermal conductivity.) Equation (115) can be rewritten for
the shear modulus in the integral form:

G(t)=

where k(t) is a certain m e m o r y function.


fo' k(t - z ) u ( z ) d z , (116)

In the case if k(t - r ) is a ~-function ( M a r k o v process), we obtain fl = 2, that is,


dM= d = 2. In the case if k(t - ~) is equal to unity at z < t and zero at z > t, we obtain
fl = 0, which corresponds to d s = I. In the general case, k(t - z) ~ (t - r) d~- ~, that is, the
fractal dimension o f the field in a deformed media is determined by the " m e m o r y " of
a medium. The invariant is the product:

c~,~)D~, = c o n s t , n =0;l ..... (ll7)


where Et'~) is the limiting density o f energy that can be accumulated in a volume o f the
order o f L 3.
F o r the creep processes in a solid, the following equation holds [35]:
6 R(t) 2G
, a~= ~<-- (118)
d"= 2 + tn ~ 3S'
1032 G . P . CHEREPANOV et al.

so that
4-co
1~</~- ~<2, (119)
2+(o
where B, G and R are the bulk, shear and relaxation moduli.

8. CONCLUSIONS
A deformed solid, as a whole, often represents a closed system for which the corresponding
fundamental thermodynamical principles are valid, as well as those of Prigogine, Clausius-Dugem,
etc. However, strictly speaking, the continuum approximation never holds because of the
non-uniform density fluctuations, so that shear exists even in the equilibrium state. Generally, the
space occupied by a deformed medium does not possess the homogeneity property. At the same
time, for deformed media, the property of scaling invariance should be valid. It is the scaling
invariance that provides the possibility of calculations of the macroscopic parameters of deformed
media in microscale [65].

Acknowledgement--The study was partially supported by AFOSR Grant 59107.

REFERENCES
[1] B. B. Mandelbrot, The Fractal Geometry of Nature, Freeman, New York (1984).
[21 M. A. Sadovskii and V. F. Pisarenko, Seismic Process in Bloc Medium (in Russian), Nauka, Moscow (1991).
[3] A. S. Balankin, Synergetics of Deformed Solid (in Russian), MO SSSR, Moscow (1991).
[4] V. S. Ivanova, Strength and Fracture of Metallic Materials (in Russian), Nauka, Moscow (1992).
[5] G. P. Cherepanov, Mechanics of Brittle Fracture, McGraw Hill, New York (1979).
[6] G. P. Cherepanov, On propagation of cracks in continuum, Prikl. Mat. Mech. 31(3), 476-488 (1967).
[7] A. S. Balankin, Fractal mechanics of deformed medium and topology of solid fracture, Dokl. Ross. Akad., Nauka,
322(5), 869-874 (1992).
[8] Z. Q. Mu and C. W. Lung, Studies on the fractal dimension and fracture toughness of steel, J. Phys. D, 21(5), 848 854
(1988).
[9] J. J. Mecholsky, D. E. Passoja and K. S. Feinberg-Rigel, Quantitative analysis of brittle fracture surfaces using fractal
geometry, J. Amer. Ceram. Soc. 72(1), 60q55 (1989).
[10] R. W. Cahn, Fractal dimension and fracture, Nature 388, 201 202 (1989).
[11] B. M. Smirnov, Physics of Fractal Clusters (in Russian). Nauka, Moscow (1991).
[12] J. Feder, Fractals. Plenum Press, New York (1991).
[13] A. S. Rosenfield, Fractal mechanics, Set. Met. 21(11), 1359-1361 (1987).
[14] A. S. Balankin, Self-organization and dissipative structures in a deformable body, SOL,. Tech. Phys. Lett. 16(4), 248-251
(1990).
[15] A. S. Balankin and V. S. lvanova, Micro-, meso- and macrokinetics of self-similar crack growth, Soy. Tech. Phys. Lett.
17(1), 32-36 (1991).
[16] A. S. Balankin, Fractal dynamics of deformed media, Soy. Tech. Phys. Lett. 17(6), 84-90 (1991).
[17] A. S. Balankin and G. N. Yanevich, Ergodynamics of impact crater formation and impact simulation principles, Soy.
Tech. Phys. Lett. 17(7), 4-9 (1991).
[18] A. S. Balankin, Fractal fracture dynamics, Soy. Tech. Phys. Lett. 17(11), 9-13 (1991).
[19] A. S. Balankin, Theory of fractals elasticity, Sot,. Tech. Phys. Lett, 17(17), 68-72 (1991).
[20] A. S. Balankin, Elastic properties of fractals, effect of transverse deformation and dynamics of free solid fracture, Dokl.
Akad, Nauk. SSSR 319(5), 1098-1101 (1991).
[21] A. S. Balankin, Quantum-statistical approach in synergetics of deformed medium, Soy. Tech. Phys. Lett. 17(14), 9(~100
(1991).
[22] A. S. Balankin and A. L. Bugrimov, Fractal dimension of cracks formed in brittle fracture of model lattices and solids,
Soy. Tech. Phys. Left. 17(17), 63~57 (1991); Fractal geometry of cracks and fatigue fracture laws for solids, Soy. Tech.
Phys. Lett. 18(14), 101 105 (1992).
[23] L. E. Richards and B. D. Dempsey, Fractal characterization of fractured surface in Ti-4.5 AI-5.0 Mo-l.5 Cr alloy
(coronas), Scripta Met. 22(5), 687-689 (1988).
[24] A. B. Mosolov, Cracks with fractal surface, Dokl. Akad. Nauk SSSR 319(4), 840 844 (1991).
[25] M. Matsushita, Fractal viewpoint of fracture and accretion, J. Phys. Soc. Japan 54(3), 857-860 (1985).
[26] A. S. Balankin, Theory of elasticity of fractals and models of nonlinear elasticity, high elasticity and fracture of
materials with multifractal structure, Dokl. Ross. Akad. Nauk 325(3), 463-469 (1992).
[27] A. S. Balankin, V. S. Ivanova and V. P. Breusov, Effects in kinetics of metal fracture and spontaneously changing
of fractal dimension of dissipative structures in ductile-brittle transition, Dokl. Ross. Akad. Nauk 322(6), 1080-1085
(1992).
[28] V. E. Panin, V. A. Likhachev and Yu. V. Grinyaev, Structural Levels of Solid Deformation (in Russian), Nauka,
Moscow (1985).
[29] V. S. lvanova, Strain energy density criterion for characterization of damage in metallic alloys, Mechanics and Physics
of Energy Density (Edited by G. C. Sih and E. E. Gdoutous), pp. 75-87. Kluver Academic, Netherlands (1991).
[30] V. E. Panin, Structural Levels o f Plastic Deformation (in Russian). Nauka, Novosibirsk (1990).
Fractal fracture mechanics 1033

[31] H. Haken, Information and Self-Organization: A Macroscopic Approach to Complex Systems, Springer, Berlin (1988).
[32] G. Nicolis and I. Prigogine, Exploring Complexity, Freeman, New York (1989).
[33] G. M. Zaskavskii and R. Z. Sagdeev, Introduction to Non-Linear Physics (in Russian), Nauka, Moscow (1988).
[34] L. D. Landau and E. M. Lifshits, Theory o f Elasticity (in Russian), Nauka, Moscow (1987).
[35] Yu. N. Rabotnov, Mechanics of Deformed Solids (in Russian), Nauka, Moscow (1988).
[36] A. S. Balankin, Synergetics and mechanics of a deformed body, Soy. Tech. Phys. Lett. 15(1 I), 878-880 (1989).
[37] G. P. Cherepanov and V. Syturin, Screw dislocation emission and fracturing in nanofracture physics, Int. J. Solid~"
Structures (submitted).
[38] G. P. Cherepanov, Some novel approaches in the mechanics of composites, Composite Materials and Structures (Edited
by C. W. Bert, V. Birman and D. Hui), pp. 1-12. ASME Press (1993).
[39] G. P. Cherepanov, Initiation of microcracks and dislocations, Soviet appl. Mech. (American Edition), 23(12),
1165-1176 (1988).
[40] G. P. Cherepanov, Microcrack growth under monotonic loading, Soviet appl. Mech. (American Edition), 24(4),
396-409 (1988).
[41] G. P. Cherepanov, Closure of microcracks during unloading and nucleation of reverse dislocations, Soviet appl. Mech.
(American Edition), 24(7), 635-648 (1989).
[42] A. S. Balankin, Kinetic (fluctuational) nature of the hydrodynamic regime of high-velocity deformation of solids, Sov.
Tech. Phys. Lett. 14(7), 539-540 (1988).
[43] G. P. Cherepanov and A. G. Cherepanov, Nanocrack nucleation by condensation of pores, Physico-Khimicheskaya
Mechanika Materialov, Soviet Mat. Sci. (American Edition), 23(6), 17-21 (1987).
[44] A. S. Balankin, A. A. Lyubomudrov and I. T. Sevryukov, Scaling effects in the kinetics of impact fracture and explosion
of solids and the problem of modeling of highly non-equilibrium systems, Sot,. Tech. Phys. Lett. 34(12), 1431-1433
(1989).
[45] A. S. Balankin, Penetration of a plastic-object into obstacles made of brittle material, Sot'. Tech. Phys. Lett. 14(7),
534-536 (1988).
[46] A. S. Balankin, Physics of cumulative jet penetration into porous media, Fizika Goreniya i Vzryva 31(4), 130 140 (1989).
[47] A. S. Balankin, A. A. Lyubomudrov and I. T. Sevryukov, Kinetic Theory of Cumulative Armor Breakdown (in Russian),
MO SSSR, Moscow (1989).
[48] V. S. Ivanova and G. V. Vstovskii, Mechanical properties of metals and alloys from synergetical viewpoint. In ltogi
Nauki i Techniki. Metallovedenie i Termicheskaya Obrabotka, Vol. 24, pp. 43 83. V1NITI, Moscow (1990).
[49] E. E. Underwood and K. Banerji, Fractals in fractography, Mater Sci, Eng. 80, 1 14 (1986).
[50] I. M. Sokolov, Dimensions and other geometric critical exponents in percolation theory, Uspekhi Fi:. Nauk, 15(2),
221-248 (1986).
[51] D. Bergman and E. Duering, Universal Poisson's ratios in two-dimensional random network of rigid and nonrigid
bonds, Phys. Rev. B, 34(11), 8199-8201 (1986).
[52] D. Bergman, Elastic moduli near percolation in a two dimensional random network of rigid and nonrigid bonds, Phys.
Rev. B, 33(3)2013-2016 (1986).
[53] A. E. Morozovskii and A. A. Snarskii, Critical behavior of elasticity moduli in percolation systems, Preprint IMF
11-90, Kiev (1990).
[54] S. Feng and P. N. Sen, Phys. Rev. Lett., 52, 216-218 (1989).
[55] P. G. De Gennes, Scaling Concepts in Polymer Physics, Cornell University Press, London (1979).
[56] A. Yu. Grosberg and A. R. Khokhlov, Statistieal Physics of Macromolecules (in Russian), Nauka, Moscow (1989).
[57] G. M. Bartenev and S. Ya. Frenkel', Physics of Polymers (in Russian), Khimiya, Leningrad, (1990).
[58] V. E. Gul' and V. N. Kuleznev, Structure and Mechanical Properties of Polymers, Vysshaya Shkola, Moscow (1979).
[59] I. Webman, The elasticity of fractals structures, In Proc. 6th Trieste Int. Syrup. on Fractals in Physics, ICTP, Trieste,
Italy, 9--12 July 1985 (Edited by L. Pietronero and E. Tosatti), pp. 488-497. Elsevier (1985).
[60] M. Muthukamar, Phys. Rev. Lett 52(3), 212.-214 (1984).
[61] E. Louis, F. Guinea and F. Flores, Fractal nature of cracks, In Proc. 6th Trieste Int. Syrup. on Fractals in Physics,
1CTP, Trieste, Italy, 9 12 July 1985 (Edited by L. Pietronero and E. Tosatti), pp. 487-497. Elsevier (1985).
[62] P. Meakin, G. Li, L. M. Sander, E. Louis and F. Guinea, A simple two-dimensional model for crack propagation,
J. Phys, A, 22(9), 1393-1403 (1987).
[63] P. H. Dauskardt, F. Houbensak and R. D. Ritchie, On interpretation of the fractal character of fracture surfaces, Acta
metall. Mat. 38(2), 143 159 (1990).
[64] E. Hornbogen, Fractals in microstructure of metals, Int. Met. Rev. 34(5), 277 296 (1989).
[65] G. P. Cherepanov, A. S. Balankin and V. S. Ivanova, Fractals and fracture, In Fracture: A Topical Encyclopedia of
Current Knowledge Dedicated to Alan Arnold Griffith (Edited by G. P. Cherepanov), Krieger, Melbourne, U.S.A. (1996).

(Received 3 April 1994)

You might also like