Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Cosserat’s effect (microstruture) coupled to Biot’s poroelasticity

in stability of circular openings in rocks

D. M. S. Lopes
National Nuclear Energy Comission, Caetité, Bahia, Brazil
R. P. Figueiredo
Federal University of Ouro Preto, Ouro Preto, Minas Gerais, Brazil

ABSTRACT: In stability analysis in wells and other underground openings the most used model
in practice is consider the rock as an elastic solid, through solution proposed by Kirsch in 1898.
Usually in several engineering applications, the rock has been treated as a homogeneous materi-
al. However, rocks are generally composite materials, and hence inhomogeneous on a micro-
scopic scale. This study has proposed another solution, in which the coupling of the poroelas-
ticity and Cosserat’s effects in stability of circular openings in rocks has been investigated. An
analytical solution to the stress concentration factor on neighborhood of a circular opening in
rocks was achieved and it applied in a simple and biaxial tensile field using two different rocks:
granite and sandstone. The results demonstrate that there were reductions on stress concentra-
tion factor of up to 55% in comparison with the traditional solution of Kirsch.

1 INTRODUCTION

1.1 Cosserat’s theory


The generalized mechanics continuum proposed by Cosserat brothers in 1909 considers that the
material points have all degrees of freedom of a rigid element, in fact, those points besides
evolving translations and rotations as well. Consequently, the static fields appear in the relation-
ship between couple-stress and gradient of rotation, such action introduces implicitly the dimen-
sions and shape of the particles in constitutive relations, which allows considering effects of the
microstructure in behavior of field.
In the generalized continuum description, the material particle is rigid and it has micro rota-
tion wc3 as additional degrees of freedom to ui. This way, in such a continuum, the particle will
have the degrees of freedom of a rigid body positioned in xi Figure 1.

Figure 1. (a) Displacement vector (ui) and micro rotation (wc3); (b) Curvatures - gradients of micro-
rotation (Figueiredo, 1999).
The kinematic of the generalized continuum is defined as:
a) Macroscopic displacement: symmetric part of macro-strain (ε(ij)) and antisymmetric part of
macro-rotation (Ω(ij)).
1
 (ij )  ( j ui   iu j ) (1)
2
1
(ij )  ( j ui   i u j ) (2)
2
b) Microscopic displacement: symmetric part of micro-strain (g(ij)) and antisymmetric part of
micro-rotation (g[ij]):
1
g (ij )  ( 'j ui'   i' u 'j ) (3)
2
1
g[ij ]  ( 'jui'  i' u 'j ) (4)
2
As the particle of Cosserat is considered rigid, the micro-strain becomes null (g(ij) = 0) and the
gradient of microscopic displacement is antisymmetric, coinciding with the Cosserat’s micro-
rotation (g[ij] = Ω(ij)).
According to Figueiredo (1999), the objectivity is requirements to formulation of constitute
laws and the only the symmetric part of macro- and micro-displacement gradients are objective
magnitudes. Thus, it is convenient to define a relative gradient (γ ij) that will correspond to the
difference between the macro- and micro-displacement tensor:
 ij   j u i  g ij (5)
Some examples of the kinematics represented by the expression γij are shown in Figure 2.

Figure 2. Kinematic relationship for γij: (a) γ22 = ∂2u2 - g22; (b) γ21 = ∂1u2 - g21; (c) γ12 = ∂2u1 - g12 (Figueire-
do, 1999).

The micro-rotation gradients that represent the curvatures (κi) are defined by:
 i   j w3c (6)
Thus, the gradients (γij e κi) are defined in 2D as:
 11  1u1  12   2u1  w3c
 22   2u2  21  1u2  w3c (7)

1  1w3c  2   2 w3c
Note that γij ≠ γji.
1.1.1 Stress-strain relationships
The equations for the linear elastic isotropic behavior of a 2D Cosserat continuum are defined
as:
 ij   kk ij  (G  Gc ) ij  (G  Gc ) ji (8)

m3i  mi  Bc i (9)


where λ = Lame’s parameter; G = shear modulus; Gc = antisymmetric shear modulus; and Bc =
flexural modulus.

1.2 Biot’s theory


This theory describes the coupled behavior between stress and pore pressure. Biot (1941) intro-
duced a quantity he called the variation in water content, which he defined as “the increment of
water volume per unit volume of soil”. The increment of water content is the volume of the wa-
ter exchanged by flow into or out of the control volume. In other words, Biot’s increment of wa-
ter content is the volume of water added to storage as used in the earlier hydrogeologic work of
Theis (1938) or Jacob (1940), although Biot’s work appears to be independent of theirs. Accord-
ing to Geertsma (1966), the mathematical description of macroscopic theory of poroelasticity is
similar to that used in the theory of thermoelasticity.
The linear relations in terms of the strain and pore pressures are taken the form:
 ij   kk  ij  2G ij  p (10)

 
 ( kk )  p (11)
3K KB
where λ = Lame’s parameter; G = shear modulus; α = Biot-Willis coefficient; K = bulk modu-
lus; p = pore pressure; θ = increment of fluid; and B = Skempton’s coefficient.

The Biot’s parameters are necessary to relate the deformations and the increment of fluids to
the stress and the pore pressure. The Biot-Willis coefficient (α) is defined as:
K
  1 (12)
Ks
where K = bulk modulus; and Ks = bulk modulus of the solid phase.

The Skempton’s coefficient (B) is defined to be the ratio of the induced pore pressure to the
change in applied stress for undrained conditions - that is, no fluid is allowed to move into or
out of the control volume:
p
B (13)

where p = pore pressure; and σ = stress for undrained condition.

A negative sign is included in the definition because the sign convention for stress means that
an increase in compressive stress, which induces a pore pressure increase, is a decrease in σ.

2 FORMULATION

2.1 Cosserat’s effects coupled to Biot’s poroelasticity


Based on the knowledge that the Cosserat’s and Biot’s theories modify the way to consider the
shear and normal stresses, respectively, and that these modifications occur independently, that
is, the effect of the Cosserat’s theory modifies only the treatment of the shear stresses and the
effect of the Biot’s theory modifies only the treatment of the normal stresses. The coupling of
the effects would not violate any prescription intrinsic to the said theories. It is anticipated that
the coupling proposed here decouple readily when one of the theories is not satisfied and this
does not cause interference to other.
It is important to note that the transient dissipation process of fluid pressures occurring in se-
quence to the aperture excavation, as treated by Detournay & Cheng (1988), has not been ad-
dressed here. This is only a short-term condition.
The equations that characterize the linear elastic behavior of a poroelastic Cosserat medium
can be written follows:
 ij   kk  ij  2G kk  2Gc [ij ]  p (14)

m3i  mi  Bc i (15)


 
 ( kk )  p (16)
3K KB
Considering the rectangular components of force per unit area (σ ij) and couple per unit area
(mi) on a 2D plane, the condition of equilibrium of moments is taken by:
m1 m2
   12   21  0 (17)
x1 x2
The equilibrium of forces is:
 11  21  12  22
 0  0 (18)
x1 x2 x1 x2
For plane strain condition (plane x1x2), the third component of displacement is zero (ε33 = 0).
Then, in an isotropic and elastic medium, the normal components of strains ε11 and ε22 are relat-
ed to the normal components of stress by:
1 
11   11  ( 11   22 )  (1  2 )p (19)
E
1 
 22   22  ( 11   22 )  (1  2 )p (20)
E
The equations (19) and (20) are similar to the equations found by Wang (2000). On the other
hand, the shear strains γ12 and γ21 may be written as:
 12   21  12   21
 12   (21)
4G 4Gc
 12   21  12   21
 21   (22)
4G 4Gc
The equations (21) and (22) are similar to the equations found by Pal’mov (1964).
The formulation proposed in this paper it is considered that there is no movement of fluid in
and out of the reference volume. According to Wang (2000) for this condition we have the fol-
lowing relation:
( u  )( 11   22 )   (1  2 )p (23)
where ν = Poisson’s ratio; νu = undrained Poisson’s ratio.

2.1.1 Equations of compatibility

The equations of compatibility, for a particular case of plane strain, can be written as:
 21  11
   11  0 (24)
x1 x2
 22  12
   22  0 (25)
x1 x2
 22  11
 0 (26)
x1 x2
As shown in equation (9) the couple-stresses are related to the curvatures (κi) and flexural
modulus (Bc). Substituting expressions (9), (19), (20), (21), (22) into (24), (25) and (26), we ob-
tain:
  12   21  12   21   1  
     11  ( 11   22 )  (1  2 )p  m1  0 (27)
x1  4G 4Gc  x2  E  Bc

 
 1  
  22  ( 11   22 )  (1  2 )p    12   21   12   21   m2  0 (28)
x1  E  x2  4G 4Gc  Bc

  m2    m1 
   0 (29)
x1  Bc  x2  Bc 
The equations (18) and (29) are necessary and sufficient conditions for the existence of the
functions φ1, φ2 and ω, such that:
1  2 1  2
 11   22   21    12   (30)
x2 x1 x1 x2
 
m1  m2  (31)
x1 x2
If the expressions (30) and (31) are inserted in equation (17), the latter becomes:
       
  1      2   0 (32)
x1  x1  x2  x1 
Thus, the equation (32) is necessary and sufficient conditions for the existence of a function
φ, such that:
   
 1   2  (33)
x1 x2 x2 x1
Rewriting the equations (30) in terms of φ and ω, we have:
 2  2  2  2  2  2  2  2
 11              (34)
x12 x1x2 x12 x1x2 x1x2 x22 x1x2 x12
22 21 12

In this way, all the components of stress and couple-stress are expressed in terms of two
stress function φ and ω.
Considering the undrained condition (equation 23) and substituting equations (31) and (34)
into equations (27) and (28), we will determine the stress functions φ and ω:

x1
 
  l 22  2(1    1  2 )h2
 2
x2
 (35)


x 2
 
  l 2  2   2(1    1  2 )h 2
 2
x1
 (36)

where
Bc Bc (G  Gc ) 3  B(1  2 )    
h2  l2  u   2   2  2 
4G 4GGc 3  B(1  2 )  x1 x2 
where Bc = flexural modulus; G = shear modulus; Gc = antisymmetric shear modulus; h and l =
characteristic lengths; ν = Poisson’s ratio; νu = undrained Poisson’s ratio; B = Skempton’s coef-
ficient; and α = Biot-Willis coefficient.

From equations (35) and (36) we can easily arrive at the equations:
 4  0 (37)

 2 (  l 2 2 )  0 (38)

2.1.1.1 Stress function in polar coordinates

The stress functions in polar coordinates are taken by:


1  1  2 1  2 1   2 1  2 1 
r        
r r r 2  2 r r r 2  r 2 r 2 r r 
1  2 1  1  1  2 1  2 1   2
 r       r     (39)
r r r 2  r r r 2  2 r r r 2  r 2
 1 
mr  m  (40)
r r 
Thus, the equations (35) and (36) transformed to polar coordinates become:

r
 
  l 22  2(1    1  2 )h2
1  2
r 
 (41)

1 
r 
  
  l 22  2(1    1  2 )h2 2
r
(42)

where
 2 1  1 2 
 2   2   
 r r r r 2  2 

3 STRESS CONCENTRATION FACTOR AROUND A CIRCULAR OPENING


3.1 Circular opening in a uniaxial field of stress
Consider a cylindrical opening of radius a in a uniaxial field of stress P, at right angles to the ax-
is of the opening, as shown in Figure 3a. For the solution, we take:
 D
   A  Br 2  2  cos(2 ) (43)
 r 
E  r 
   2  H .BesselK  2,   sin(2 ) (44)
r  l 
where K = modified Bessel function of the second kind and second order.
Figure 3. Circular opening in (a) uniaxial field of stress; (b) biaxial field of stress

The stress functions given in equations (43) and (44) satisfy equations (37) and (38).
Considering stress concentration in the neighborhood of a circular opening and assume at the
periphery of the hole r = a is free of stresses and couple-stress and that at infinity we have the
state of stress:
 r   r  mr  m  0 (45)

 
P
1  cos  (46)
2
Substituting from equations (43) and (44) into equations (39) and (40), and considering the
state of stress from equations (45) and (46), we find for the stress σθ on the periphery of the
opening
 2 cos    3  F  for θ = ± π/2
   P1   or   max  P  (47)
 1 F   1 F 
where
h2
81     1  2  Bc Bc (G  Gc )
F l2 h2  l2 
  a  4G 4GGc
 a 2 2a K 0  l  
4  2  .   
 l l a
 K1   
  l 
Thus, the stress concentration factor (θ = ± π/2) for the neighborhood of the opening is
  max 3  F
  (48)
P 1 F
3.2 Circular opening in a biaxial field of stress
Considering now a biaxial field of stress as shown in the Figure 3b and also considering the
same conditions of the previous section, we take:
3P  S  P  S F
  max  (49)
1 F

3.3 Examples
From equations (47) and (49), it is seen that the stress concentration factor depends on the elas-
tic constants (G, Gc, Bc, υ, ) and the radius of the hole (a). For example, considering the circu-
lar opening in a sandstone (properties Table 1) the maximum tangential stress concentration fac-
tor (δ), inserted in a uniaxial and biaxial field of stress, is shown in Figure 4.

Table 1. Some properties of sandstone.


____________________________________________
Rock type υ B
____________________________________________
Berea Sandstone 0,20 0,79 0,62
____________________________________________
Figure 4. Stress concentration factor on circular hole: (a) uniaxial field of stress; (b) biaxial field of stress.

4 CONCLUSION

According of results obtain above it is clear that the largest the stress concentration factor is ob-
tained for the smallest F. But we can note it follows that minimum value of F is zero and that it
is attained for a/l → ∞. In this case one attains the classical value of δ = 3 (uniaxial) or δ = 4
(biaxial field of stress).
The smallest δ is attained at the largest F. But we can also note that the maximum F is real-
ized for a/l → 0 and h/l = 1. In these cases, the stress concentration factor is 1.9 inserted in uni-
axial field of stress, and 1.8 when inserted in biaxial field. That is means almost 40% and 55%
of reduction of stress concentration factors, respectively, when compared with traditional solu-
tion of Kirsch (1898).
It is worth remembering that when we disregard the effects of poroelasticity the coupling is
undone and the solution resembles the solution found by Mindlin (1963).

5 REFERENCES

Biot, M.A. (1941). General Theory of Three-Dimensional Consolidation. Journal of Applied Physics, vol.
12, pp. 155-164.
Cosserat, E.; Cosserat, F. (1909). Théorie des Corps Déformables. Paris: Hermann et Fils, pp. 226.
Detournay, E. & Cheng, A. H.D. (1988). Poroelastic response of a borehole in a non-hydrostatic stress
field. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., vol. 25, nº 3, pp. 171-182.
Figueiredo, R.P. (1999). Modelagem de Maciços Rochosos como Meios Contínuos Generalizados de
Cosserat. Tese de Doutorado, Departamento de Engenharia Civil, PUC-Rio Brasil.
Geertsma, J. (1966). Problems of rock mechanics in petroleum production engineering. In Proc. 1º Cong.
Int. Soc. Rock Mech., vol. 1, pp. 585-594, Lisboa.
Jacob, C.E. (1940). On the flow of water in an elastic artesian aquifer. Trans. Am. Geophys. Union, vol.
22, pp. 783-787.
Kirsch, E.G. (1898). Die Theorie der Elastizität und die Bedürfnisse der Festigkeitslehre. Zeitschrift des
Vereines deutscher Ingenieure, vol. 42, pp. 797–807.
Mindlin, R.D. (1963). Influence of Couple-stresses on Stress Concentrations. Experimental Mechanics,
vol. 1, nº 1, pp. 1-7.
Pal’mov, V.A. (1964). The plane problem in the theory of nonsymmetrical elasticity. PMM, vol. 28, nº 6,
pp. 1117-1120.
Theis, C.V. (1938). The significance and nature of the cone of depression in ground-water bodies. Eco-
nomic Geology, vol. 33, pp. 889-902.
Wang, H.F. (2000). Theory of Linear Poroelasticity with Applications to Geomechanics and Hydrogeolo-
gy. Princeton University Press.

You might also like