PhD Thesis FAN2014
PhD Thesis FAN2014
PhD Thesis FAN2014
formations
Shock-induced borehole waves in fractured
formations
PROEFSCHRIFT
Samenstelling promotiecommissie:
Rector Magnificus, Technische Universiteit Delft, voorzitter
Prof.dr.ir. D.M.J. Smeulders, Technische Universiteit Eindhoven, promotor
Prof.dr. N. Li, PetroChina Research Institute, Beijing
Prof.dr.-ing. H. Steeb, Ruhr-Universität Bochum
Prof.dr. G. Bertotti, Technische Universiteit Delft
Prof.dr. A.V. Metrikine, Technische Universiteit Delft
Prof.dr. R.J. Schotting, Universiteit Utrecht
Prof.dr.ir. C.P.A. Wapenaar, Technische Universiteit Delft
ISBN 978-90-8891-910-7
c 2014 by H. FAN.
All rights reserved. No part of the material protected by this copyright notice
may be reproduced or utilized in any form or by any means, electronic or
mechanical, including photocopying, recording or by any information storage
and retrieval system, without permission from the author.
Summary xi
Samenvatting xiii
1 Introduction 1
1.1 Ray tracing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Stoneley wave . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Literature survey . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.1 Bulk waves in porous media . . . . . . . . . . . . . . . 5
1.3.2 Waves along flat interfaces . . . . . . . . . . . . . . . . 5
1.3.3 Stoneley waves in a borehole . . . . . . . . . . . . . . 6
1.3.4 Stoneley waves in a borehole intersected by fractures . 7
1.4 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
vii
viii Contents
4 Experimental setup 47
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.2 Vacuum procedure . . . . . . . . . . . . . . . . . . . . . . . . 50
4.3 Sample overview and preparation . . . . . . . . . . . . . . . . 52
8 Conclusions 97
Bibliography 107
Acknowledgements 117
xi
xii Summary
profiles.
Reflection from the water-sample interface and from the free water inter-
face can be recorded by means of a fixed pressure transducer mounted in the
wall of the shock tube above the sample in the water layer. The fractures
in the formation are represented by small horizontal slits in composite cyl-
inders whose upper and lower parts are separated by small spacer poles. In
this way, a variable horizontal fracture (slit) aperture can be obtained. Obvi-
ously these fractures form an open connection between the borehole fluid and
the fluid outside the cylinder. Also mandrel samples are used for horizontal
slits that are not open to the fluid outside the cylinder, thus representing
fractures with finite radial extension. Wave experiments show that varying
fracture widths significantly alter the recorded Stoneley wave pressure signal
at fixed depth. The reflection and transmission of borehole tube waves over
1 and 5 mm fractures are correctly predicted by theory. Other wave experi-
ments show that attenuation in boreholes adjacent to porous zones without
fractures can be predicted by theory. This technique even allows a direct
measurement of the permeability, although the acoustically measured per-
meability and the permeability measured by falling-head technique still show
a significant discrepancy. This technique is directly applicable to fractured
porous reservoir core samples.
Samenvatting
xiii
xiv Samenvatting
Introduction
Formaon
Head wave wave front
Wave front
Formaon
wave front
Source Source
Figure 1.1: Sketches of a borehole and the adjacent formation. A source generates a fluid
wave in the borehole. At the interface between the formation and the borehole, the fluid
wave partially reflects back into the borehole and partially transmits into the formation
to be converted into formation body waves (left panel). A head wave in the borehole is
generated (right panel).
1
2 1. Introduction
the weight of the tool string and also transmits measurement data to the
surface. Acoustic logging is one of the borehole logging techniques. In this
technique, sound chirps in the kHz frequency range are generated that travel
through the formation and thus carry information on its consistency. The
sound waves are recorded by the tool and interpreted for reservoir properties.
Here we introduce the basic concepts of monopole acoustic logging.
In Fig. 1.1, a borehole is sketched as well as the adjacent formation. In the
left panel of Fig. 1.1, a source generates a fluid wave which propagates as a
spherical wave. When the fluid wave reaches the interface between the form-
ation and the borehole, the wave partially reflects back into the borehole and
partially transmits into the formation to be converted into formation body
waves. The formation body waves comprise compressional waves (also called
P-waves) and shear waves (also called S-waves). The propagation speeds Vp
and Vs of the compressional and shear waves depend on the formation prop-
erties. In a so-called fast formation, Vs > Vf , in a slow formation, Vs < Vf ,
where Vf is the borehole fluid wave speed.
According to Huygens’ principle, each point at the fluid-solid interface
becomes a secondary wave source. The secondary source generates two kinds
of waves: the borehole head wave (velocity Vp or Vs ) and the secondary
compressional and shear waves in the formation. In most situations, we ignore
the secondary body waves, but focus on the head waves in the borehole. By
using this method, one can measure the Vp and Vs of the formation in the
borehole. In a slow formation, there is no shear head wave present.
Strictly speaking, the ray tracing method is only valid when the fluid wavelen-
gth in the borehole is much smaller than the borehole diameter or when the
fluid wavefront can be considered plane. Although these conditions are not
always fully met, ray tracing is still a very powerful technique for problem
visualization. The ray tracing method helps to better understand the travel
paths of the waves and provides valuable information for logging tool design.
Borehole Formation 2 1 1 2
Reflected
P-wave Transmied Receiver array
P(S)-wave
2
1
Incident
P-wave
Transmitter
Source
Figure 1.2: A fluid wave in the borehole propagates to the interface between the formation
and the borehole with incident angle θ1 . This wave partially reflects back into the borehole
with the same angle θ1 , and partially transmits into the formation with angle θ2 . The
compressional body waves and shear body waves can be distinguished by using the Snell’s
law (left panel of this figure). In the right panel of this figure, different ray paths can be
identified. By using a sophisticated acoustic tool with multiple receivers (receiver array),
detailed information on formation 1 and 2 can be obtained, as well as the position of the
interface between them.
Acoustic logging can also be used for more complicated formations, where,
for example, the formation properties are altered with radial penetration
depth. In the right panel of Fig. 1.2, different ray paths can be identified. By
using a sophisticated acoustic tool with multiple receivers (receiver array),
detailed information on formation 1 and 2 can be obtained, as well as the
position of the interface between them.
4 1. Introduction
Receiver
Fracture Fracture
Transmitter
Figure 1.3: A Stoneley wave propagating along the borehole wall is affected by permeable
zones and fractures intersecting the borehole.
1.3. Literature survey 5
permeable zones, the wave partially reflects and partially transmits as in-
dicated in Fig. 1.3. The presence of a permeable zone causes the Stoneley
wave to become dispersive, so that velocity and attenuation are induced as
a function of frequency. When the Stoneley wave crosses a fracture, similar
effects occur. One can obtain information on fractures and permeable zones
by analyzing the reflected and transmitted Stoneley waves.
from the Stoneley wave, also pseudo-Rayleigh wave exists. Viktorov (1967),
Überall (1973), and Brekhovskikh (1980) investigated acoustic surface waves
in general. The first studies of surface wave propagation at the interface
between fluid and fluid-saturated porous media can be attributed to Der-
esiewicz (1960, 1961, 1962) and Deresiewicz and Skalak (1963). By invest-
igating the high-frequency range of Biot’s theory where all body waves in
the porous medium are non-attenuated, Feng and Johnson (1983a,b) found
that there are three types of surface waves existing at the interface between
fluid and fluid-saturated porous media. These are pseudo-Rayleigh waves,
pseudo-Stoneley waves and true Stoneley waves. The pseudo-Stoneley wave
is attenuated as it leaks energy into the slow compressional wave. The true
Stoneley wave is non-attenuated and it has a velocity lower than the slow
compressional wave (Chao, 2005).
Other studies were performed by Edelman and Wilmanski (2002), Albers
(2006), Markov (2009) and van Dalen et al. (2010, 2011a). For the experi-
mental investigations, we refer to the work of Mayes et al. (1986), Adler and
Nagy (1994), Allard et al. (2004) and van Dalen (2011b).
and Toksöz (1981), Hardin et al. (1987), Hornby et al. (1987), Hsu et al.
(1987), Hsu and Esmersoy (1992), and Brie et al. (2000).
Hardin et al. (1987) reported an approach for borehole Stoneley wave at-
tenuation based on the steady fluid flow into the fracture. Tang and Cheng
(1989) developed a wave propagation theory in a single fracture bounded by
two rigid half spaces based on the Navier-Stokes equations. The viscous ef-
fects of the fluid were taken into account. Ferrazzini and Aki (1987) presented
a theory of wave propagation in a fracture bounded in two elastic half spaces
where viscous effects were neglected. Kostek et al. (1998b) and Korneev
(2008, 2010) discussed wave propagation in a fracture which includes both
the fluid viscous effects and the fracture surfaces elasticity effects.
• The displacements of both the fluid and solid phases are assumed to be
sufficiently small, so that the equations can be linearized;
• The fluid does not react to a shear force in the solid. The fluid cannot
sustain shear forces;
We consider a cube of unit size of bulk material. The total normal tension
force per unit bulk area applied to the fluid faces of the unit cube τ can be
described by:
pAf
τ =− = −φp, (2.1)
Ab
where φ is porosity of the system, p is the pressure of the fluid in the pores,
Af and Ab are total surfaces of the interconnected pores and the bulk area
9
10 2. Constitutive relations and momentum equations
respectively. The total normal tension force per unit bulk area applied to
that portion of the cube faces occupied by the solid τij can be described by:
τij = −σij − (1 − φ)pδij , (2.2)
where the intergranular stress σij refers to the intergranular forces percent
bulk area. The Kronecker symbol δij is introduced because the pore fluid
cannot exert nor sustain any shear forces on the macroscopic scale.
The stress-strain relations can be written as follows (Biot, 1955):
τij = 2μeij + Aekk δij + Qkk δij , (2.3)
τ = Qekk + Rkk , (2.4)
1−φ 1−φ
σij = −2μeij − A − Q ekk δij − Q − R kk δij , (2.5)
φ φ
where eij = 1/2(∂usi /∂xj + ∂usj /∂xi ), ij = 1/2(∂uf i /∂xj + ∂uf j /∂xi ), and
summation over repeated indices is assumed. A, Q and R are generalized
elastic parameters which can be related via Gedanken experiments to poros-
ity, bulk modulus of the solid Ks , bulk modulus of the fluid Kf , bulk modulus
of the porous drained matrix Kb , and shear modulus μ of both the drained
matrix and of the composite (Biot and Willis, 1957):
A = Kb − 2μ/3 + Kf (1 − φ − Kb /Ks )2 /φeff , (2.6)
Q = φKf (1 − φ − Kb /Ks )/φeff , (2.7)
R = φ2 Kf /φeff , (2.8)
where the effective porosity φeff = φ + Kf /Ks (1 − φ − Kb /Ks ). If we assume
that the frame and pore fluid are much more compressible than the solids
themselves (Kb /Ks 1 and Kf /Ks 1), we can write
(1 − φ)2 2
A= Kf + Kb − μ, (2.9)
φ 3
Q = (1 − φ)Kf , (2.10)
R = φKf . (2.11)
For this case, the stress-strain relations are simplified. Substituting (2.9),
(2.10) and (2.11) into (2.4) and (2.5), we obtain:
2
σij = −(Kb − μ)ekk δij − 2μeij , (2.12)
3
and
1−φ
p=− Kf ekk − Kf kk . (2.13)
φ
2.2. Momentum equations 11
∂2 ∂p
φρf 2
uf i = −φ − fi . (2.15)
∂t ∂xi
The viscous damping factor b0 = ηφ2 /k0 , with η the dynamic fluid viscosity
and k0 the permeability. The tortuosity is denoted α∞ . The equations of
motion resulting from momentum conservation and the stress-strain relations
can now be written as:
∂2
μ∇2 us + (A + μ)∇∇ · us + Q∇∇ · uf = (ρ11 us + ρ12 uf )
∂t2
∂
+b0 (us − uf ), (2.17)
∂t
∂2 ∂
Q∇∇ · us + R∇∇ · uf = 2
(ρ12 us + ρ22 uf ) − b0 (us − uf ), (2.18)
∂t ∂t
where the density terms are given by
We now consider harmonic waves of the form exp i(ωt − κx), where κ is the
wavenumber. Substitution into (2.30) and introduction of the phase speed
c = ω/κ, yields that
4 iθ −1 2 2 2 iθ −1 2
c 1− − c V+ + V− − V + V+2 V−2 = 0. (2.31)
ω ω 0
From (2.31), the squared phase speed is expressed as
√
2 V+2 + V−2 − i(ωθ)−1 V02 ± D
c = , (2.32)
2[1 − i(ωθ)−1 ]
with
D = (V+2 − V−2 )2 + 2i(ωθ)−1 [2V+2 V−2 − V02 (V+2 + V−2 )] − (ωθ)−2 V04 . (2.33)
The result of the dispersion relation (2.32) is that there are two distinct
longitudinal modes which we call mode 1 (first wave) and mode 2 (second
wave). The solution (2.32) is somewhat modified by introducing a frequency-
dependent friction factor. It corrects for the fact that above the characteristic
frequency θ −1 , deviations from low-frequency Stokes’ flow become important
owing to inertia
effects. In the limit of high frequencies, the viscous skin
depth δ = 2ν/ω (ν denotes the kinematic viscosity η/ρf ) eventually be-
comes much smaller than a characteristic viscous length scale Λ (Smeulders,
1992a). The steady-state friction factor b0 in (2.29) is then replaced by a
more realistic frequency-dependent friction factor b(ω) = b0 F (ω), where the
viscous correction factor F (ω) is given by:
b(ω) 1 ω
F (ω) = = 1 + iM , (2.34)
b0 2 ωc
where
ηφ
ωc = (2.35)
α∞ k0 ρf
14 2. Constitutive relations and momentum equations
3500
3000 Fast
Phase Velocity (m/s)
2500
2000
1500 Shear
1000
500 Slow
0 −2 −1 0 1 2
10 10 10 10 10
ω/ωc
2
10
Slow
0 Shear
10
Attenuation (m 1)
−
Fast
−2
10
−4
10
−6
10 −2 −1 0 1 2
10 10 10 10 10
ω/ω
c
Figure 2.1: Phase velocity and attenuation factors of the fast, slow and shear wave for a
water-saturated porous medium. The parameters of the porous medium are listed in Table
6.1.
is the frequency for which inertia and viscous forces are of equal importance
in rigid porous solids, and M is the viscous shape factor defined as
8α∞ k0
M= . (2.36)
φΛ2
2.3. One-dimensional field equations 15
It was shown that for many porous materials the shape factor M = 1 in
good approximation, so that the characteristic viscous
length scale
is directly
related to other material properties by Λ = 8k0 α∞ /φ = 8ν/ωc . We
remark that there is direct relation between the characteristic frequencies ωc
and θ −1 :
(1 − φ)ρs + (1 − α−1 )φρf
ωc = θ −1 . (2.37)
(1 − φ)ρs + φρf
∂2 ∂
0= 2
(ρ12 usy + ρ22 uf y ) − b0 (usy − uf y ). (2.39)
∂t ∂t
The above equations can also be written as
∂2 ∂ ∂2 ∂ ∂2
ρ11 2 + b0 − μ 2 usy = b0 − ρ12 2 uf y , (2.40)
∂t ∂t ∂x ∂t ∂t
∂ ∂2 ∂2 ∂
b0 − ρ12 2 usy = ρ22 2 + b0 uf y . (2.41)
∂t ∂t ∂t ∂t
where ρ = ρ11 + ρ22 + 2ρ12 . Dividing by ρ11 ρ22 − ρ212 , we find that
4 2
∂ 2 ∂4 −1 ∂ ∂ 2 ∂
2
− Vs 2 2 + θ − V0 usy = 0, (2.44)
∂t4 ∂t ∂x ∂t ∂t2 ∂x2
16 2. Constitutive relations and momentum equations
with
μρ22
Vs2 = , (2.45)
ρ11 ρ22 − ρ212
μ
V02 = . (2.46)
ρ
3.1 Introduction
We consider a single horizontal layer, having a height h, intersecting the
borehole (see Fig. 3.1). The layer can be a fracture with aperture h, or a
permeable zone. The layer is of infinite extent in the radial (r) direction.
The borehole has radius b. The plane z = 0 is in the center of the layer.
The z-coordinate is pointing downward. We assume that the borehole fluid
pressure is uniform across the borehole. The 1-D borehole wave equation is
d2 ψ
+ κ2 ψ = 0, (3.1)
dz 2
where ψ is the displacement potential, and κ is the wavenumber. The fluid
pressure p and the axial fluid displacement U in the borehole are given by
p = ρf ω 2 ψ, (3.2)
dψ
U= , (3.3)
dz
where ρf is the fluid density, and ω is the angular frequency. Borehole wave
propagation in the z-direction is described by
17
18 3. Tube wave propagation in a fluid-filled borehole
2b
Δu z
Borehole
ub
Formation Formation
r
Δz h Viscous fluid
Formation Formation
z
Figure 3.1: Borehole intersected by a single horizontal fracture. The distortion of an
elementary volume of borehole fluid by fluid compression and wall expansion is indicated
by the dashed box. The axial compression of the volume element is denoted by Δuz . The
radial wall displacement is denoted ub .
In the region z < −h/2, A+ e−iκ1 z and A− eiκ1 z represent the incident wave
propagating in the positive z direction and the reflected wave propagating in
the negative z direction, respectively. Note that κ1 is the fluid wavenumber in
the undisturbed borehole, and κ2 is the fluid wavenumber where the borehole
is intersected by the layer. A+ is the incident amplitude, and A− is the
reflected amplitude (see Fig. 3.2). In the region −h/2 < z < h/2, B + and
B − are the amplitudes of the waves propagating in the positive and negative
z directions, respectively. C + is the amplitude of the transmitted wave in
the region z > h/2. The fluid displacement and the pressure should be
continuous at z = h/2 and z = −h/2. The coefficients A− , B + , B − and C +
can now be calculated as a function of the incident amplitude coefficient term
A+ :
2b
A+ A-
B+ B- r
h
C+
z
Figure 3.2: Stoneley wave propagation in the borehole intersected by a single horizontal
fracture. A+ and A− represent incident and reflected Stoneley wave amplitude in z <
−h/2 respectively; B + and B − represent incident and reflected Stoneley wave amplitude
in −h/2 < z < h/2 respectively; C + represents transmitted Stoneley wave amplitude in
z > h/2.
where D is given by
D = (κ1 + κ2 )2 eiκ2 h − (κ1 − κ2 )2 e−iκ2 h . (3.11)
The above equations were also found by Tang and Cheng (1993). We will
now discuss tube wave propagation over different types of layers.
where ρf is the density of the fluid inside the pore. Combining (3.22) and
(3.23), we find that
2
∂p ∂ p 1 ∂p
= Dh + , (3.24)
∂t ∂r 2 r ∂r
where Dh = k0 Kf /(ηφ) is the hydraulic diffusivity, with Kf the bulk modulus
of the pore fluid. For harmonic variations p = p̂ exp(iωt), (3.24) becomes
2
∂ 1 ∂ 2
+ − κ p̂ = 0, (3.25)
∂r 2 r ∂r
with
ω2 iω
κ2 = =− , (3.26)
c2f Dh
where cf is the speed of sound in the fluid. The solution to (3.25), which is
finite at large distance is
2
10
1
10
ηb/(k0Zb)
0
10 Re
−1 Im
10
−2
10 −2 −1 0 1 2
10 10 10 10 10
b(ω/Dh)1/2
Figure 3.3: Real and imaginary parts of the reduced wall impedance.
A plot of the reduced wall impedance is given in Fig. 3.3. For high frequencies,
√
both the real and imaginary parts tend to infinity as ω. From (3.19) and
(3.31), the tube wave number can be obtained
ω ω 2k0 Kbf κ K1 (κb) Kbf
κT = = 1+ + . (3.32)
cT cf iωηb K0 (κb) G
hole is filled with the same fluid as the reservoir and that ρf = ρbf = 1000
kgm−3 , we now compute the wave speed and attenuation of the tube wave as
a function of frequency using (3.32). For the shear modulus we take G = 3.0
24 3. Tube wave propagation in a fluid-filled borehole
1200
1000
Phase Velocity (m/s)
a
800
b
600
c
400
200
0 −2 0 2 4 6
10 10 10 10 10
f (Hz)
1
10
0
10
Attenuation (m−1)
−1
10 c
b
−2
10 a
−3
10
−4
10 −2 0 2 4 6
10 10 10 10 10
f (Hz)
Figure 3.4: Tube wave phase velocity and attenuation along a pore formation from (3.32).
a: k0 =10 mD, φ=16%, b: k0 =100 mD, φ=20%, c: k0 =1000 mD, φ=26%.
GPa. Results are plotted in Fig. 3.4. The dependence on porosity and per-
meability is clearly visible. The higher the permeability, the more attenuation
is observed due to the increased infiltration into the formation.
When the permeable layer is replaced by a horizontal fracture (see Fig. 3.1),
the wavenumber κ2 in the fracture zone can also be obtained from (3.19), but
3.2. Borehole waves in flexible and fractured formations 25
with different Zb . The difference with respect to the fluid flow in the per-
meable formation is that the relation between fracture velocity and fracture
pressure is now given by Poiseuille’s law:
h2 ∂ p̂
ŵr = − . (3.33)
12η ∂r
The wall impedance Zb can then be written as
1 h2 K1 (κb)
= κb , (3.34)
Zb 12ηb K0 (κb)
and the effective fluid bulk modulus is given by
1 1 1 h2 κ K1 (κb)
= + + . (3.35)
Keff Kbf D 6iωηb K0 (κb)
We thus easily obtain the wavenumber κ2 in the borehole adjacent to the
fracture:
ω ω h2 Kbf κ K1 (κb) Kbf
κ2 = = 1+ + . (3.36)
cT cf 6iωηb K0 (κb) D
Obviously it is assumed here that 1-D approach is valid which is realistic for
λ/2b 1, for b the borehole radius. The vertical motion of the layer’s wall
is discussed in section 3.3.2.
1200
1000
Phase Velocity (m/s)
a
800
b
600
c
400
200
0 −2 0 2 4 6
10 10 10 10 10
f (Hz)
1
10
0
10
Attenuation (m )
−1
−1
10 c
b
−2
10 a
−3
10
−4
10 −2 0 2 4 6
10 10 10 10 10
f (Hz)
Figure 3.5: Tube wave phase velocity and attenuation along a porous formation from
(3.32). The solid lines are for constant permeability and the dashed lines are for dy-
namic permeability (3.37) and wavenumber (3.40). a: k0 =10 mD, φ=16%, b: k0 =100 mD,
φ=20%, c: k0 =1000 mD, φ=26%.
1200
1000
Phase Velocity (m/s)
800
a
600
b
400
200
c
0 −2 0 2 4 6
10 10 10 10 10
f (Hz)
4
10
2
10
Attenuation (m )
−1
0 c
10
b
a
−2
10
−4
10 −2 0 2 4 6
10 10 10 10 10
f (Hz)
Figure 3.6: Tube wave phase velocity and attenuation in the borehole adjacent to the
fracture from (3.36) (solid lines) compared with where the dynamic parameters are included
(dashed lines). Fracture apertures are a: 1 μm, b: 10 μm, c: 100 μm.
The introduction of inertia effects will also affect the wave number κ,
which was so far diffusive in nature (3.26). The modified wave number in the
porous medium can be written as (Johnson et al., 1987)
ω
κ̃ = α(ω). (3.40)
cf
28 3. Tube wave propagation in a fluid-filled borehole
4
10
Phase Velocity (m/s)
2
10
0
10
−2
10 −2 0 2 4
10 10 10 10
f (Hz)
4
10
3
10
Attenuation (m−1)
2
10
1
10
0
10 −2 0 2 4
10 10 10 10
f (Hz)
Figure 3.7: Tube wave phase velocity and attenuation in the borehole adjacent to the
fracture from (3.36) (solid lines) compared with where the dynamic parameters are included
(dashed lines). Fracture aperture is 1 cm.
We noted that the dynamic interaction between solid and fluid in porous
media can be described by a dynamic permeability which extends the low-
frequency viscous effects toward high-frequency inertia effects. Likewise, all
these effects are also comprised in a dynamic tortuosity, which extends the
3.3. Exact solutions for fracture and borehole waves 29
where b0 = ηφ2 /k0 is the viscous damping factor of the porous media.
For a fracture, (3.41) becomes
12iη 12iη iωh2 ρf
α(ω) = 1 − 2 F (ω) = 1 − 2 1+ . (3.42)
h ωρf h ωρf 36η
This means that eqs. (3.31), (3.32), (3.34) and (3.36) still hold, but we simply
replace k0 by k(ω) and κ by κ̃.
The phase velocity and attenuation of the tube wave in porous forma-
tions from (3.32) are compared with and without dynamic parameters (see
Fig. 3.5). For the phase velocity the two situations agree very well in both
low- and high- frequency ranges. For the attenuation, they agree very well
in low frequencies but have some discrepancy in the high-frequency range.
Above the critical frequencies ωc (see Table 3.1), inertia effects become dom-
inant and the attenuation departures from the viscosity-dominated region
(see Fig. 3.5).
Tube wave phase velocity and attenuation in the borehole adjacent to the
fracture from (3.36) are compared with where the dynamic parameters are
included (see Fig. 3.6). For 1 μm and 10 μm fractures, the two situations
agree well in all frequencies with regard to both the phase velocity and atten-
uation. When the fracture aperture increases to 100 μm, they begin to show
different behavior around 100 Hz. When the fracture aperture increases to 1
cm (see Fig. 3.7), discrepancies become very large as inertia effects become
dominant.
∂ 2 Φ̂ 1 ∂ Φ̂ ∂ 2 Φ̂ ω2
+ + + Φ̂ = 0, (3.51)
∂r 2 r ∂r ∂z 2 c2f + 43 iων
∂ 2 ψ̂ 1 ∂ ψ̂ ∂ 2 ψ̂ ψ̂ ω
2
+ + 2
− 2 + ψ̂ = 0. (3.52)
∂r r ∂r ∂z r iν
The solution to (3.51) and (3.52) is given by
Φ̂ = H0 (κr)[A cos(f z) + B sin(f z)], (3.53)
3.3. Exact solutions for fracture and borehole waves 31
1500
c
Phase Velocity (m/s)
1000
b
500
0 1 2 3 4 5
10 10 10 10 10
f (Hz)
4
10
3
10
Attenuation (m 1)
−
2
10 a
1 b
10
0
10 c
−1
10 1 2 3 4 5
10 10 10 10 10
f (Hz)
Figure 3.8: Phase velocity and attenuation of the wave in the fracture from the exact
solution (3.61) (solid lines) compared with the generalized solution (3.40) (dashed lines).
Fracture apertures are a: 1 μm, b: 10 μm, c: 100 μm.
1500
1 mm
Phase Velocity (m/s)
1000 100 μm
10 μm
1 mm
500
100 μm
10 μm
0 1 2 3 4 5
10 10 10 10 10
f (Hz)
Figure 3.9: Phase velocity of the wave in the fracture bounded by two elastic half spaces
from (3.65) (dashed lines) compared with two rigid half spaces from (3.61) (solid lines).
Three fracture apertures 10 μm, 100 μm and 1 mm are considered.
Other theories for the fracture waves have been given by Korneev (2008,
2010), who takes both the viscosity effect of the fracture fluid, and the form-
ation elasticity into account.
1200
a
1000
Phase Velocity (m/s)
800
b
600
400
c
200
0 1 2 3 4 5
10 10 10 10 10
f (Hz)
1
10
c
Attenuation (m 1)
0
−
10
b
−1
10 a
−2
10 1 2 3 4 5
10 10 10 10 10
f (Hz)
Figure 3.10: Tube wave phase velocity and attenuation in the borehole adjacent to the
fracture from (3.75) (solid lines) compared with (3.36) in which the dynamic parameters
are included (dashed lines). Fracture apertures are a: 1 μm, b: 10 μm, c: 100 μm.
Now we can obtain the average fluid velocity at the fracture opening using
(3.69) and (3.71):
iω H1 (κb)
< ω̂b >= 2 2 − < p̂b > κ , (3.72)
κ cf ρf H0 (κb)
36 3. Tube wave propagation in a fluid-filled borehole
1200
1000
Phase Velocity (m/s)
800
600
400
c b a
200
0 1 2 3 4
10 10 10 10
f (Hz)
0.9
10
c
0.7 b
Attenuation (m−1)
10
0.5
10
a
0.3
10
0.1
10
1 2 3 4
10 10 10 10
f (Hz)
Figure 3.11: Tube wave phase velocity and attenuation in the borehole adjacent to the
fracture from (3.75) (solid lines) compared with (3.36) in which the dynamic parameters
are included (dashed lines). Fracture apertures are a: 0.1 mm, b: 1 mm, c: 1 cm.
p+ = ρf ω 2 (C + e−iκz ). (3.77)
2b
A+ A-
r
h
C+
z
Figure 3.12: Stoneley wave propagation in the borehole intersected by a single horizontal
fracture of Hornby’s approach. A+ and A− represent incident and reflected Stoneley wave
amplitude in z < 0 respectively; C + represents transmitted Stoneley wave amplitude in
z > 0.
T = 1 + R. (3.79)
1 C+
[C + − (A+ − A− )]πb2 + 2πbh = 0, (3.81)
Zf Zb
which can be rewritten as
2h Zf
1−R=T 1+ . (3.82)
b Zb
3.4. Reflection and transmission 39
0.8
0.6
|R|
5 cm
0.4
3 cm
0.2
1 cm
0
0 500 1000 1500 2000
f (Hz)
1
1 cm
0.8
3 cm
0.6
5 cm
|T|
0.4
0.2
0
0 500 1000 1500 2000
f (Hz)
Figure 3.13: Amplitude of reflection and transmission coefficients over a horizontal frac-
ture bounded by rigid solid formations. The fracture apertures are 1 cm, 3 cm and 5 cm
respectively. Solid lines are from (3.7) and (3.10) where borehole flexibility is taken into
account, dashed lines are from (3.83) where borehole flexibility is neglected.
Combination of (3.79) and (3.82) allows us to find expressions for the reflec-
tion and transmission coefficients:
h Zf h Zf h Zf
R=− / 1+ ; T = 1/ 1 + . (3.83)
b Zb b Zb b Zb
40 3. Tube wave propagation in a fluid-filled borehole
180
170
Phase of R (degree)
160
150 1 cm
3 cm
140
5 cm
130
0 500 1000 1500 2000
f (Hz)
80
70
Phase of T (degree)
60
50
40
30
20 5 cm
10 3 cm
1 cm
0
0 500 1000 1500 2000
f (Hz)
Figure 3.14: Phase of reflection and transmission coefficients over a horizontal fracture
bounded by rigid solid formations. The fracture apertures are 1 cm, 3 cm and 5 cm
respectively. Solid lines are from (3.7) and (3.10) where borehole flexibility is taken into
account, dashed lines are from (3.83) where borehole flexibility is neglected.
For the wall impedance, we adopt here the expression given by Hornby:
(1)
Zf iH1 (κ̃b)
= (1)
. (3.84)
Zb α(ω)H0 (κ̃b)
3.4. Reflection and transmission 41
0.8
0.6
|R|
5 cm
0.4 3 cm
0.2 1 cm
0
0 500 1000 1500 2000
f (Hz)
1
1 cm
0.8
3 cm
0.6
|T|
5 cm
0.4
0.2
0
0 500 1000 1500 2000
f (Hz)
Figure 3.15: Amplitude of reflection and transmission coefficients over a horizontal frac-
ture bounded by rigid solid formations. The fracture apertures are 1 cm, 3 cm and 5 cm
respectively. Solid lines are from (3.7) and (3.10), dashed lines are from (3.83). Borehole
flexibility is neglected in both cases.
We now compare expressions (3.83) with (3.7) and (3.10) for b = 10 cm. In
(3.7) and (3.10) we have used that
ω Kbf
κ1 = 1+ , (3.85)
cf G
42 3. Tube wave propagation in a fluid-filled borehole
0.8
0.6
|R|
5 cm
0.4
0.2 3 cm
1 cm
0
0 500 1000 1500 2000
f (Hz)
1
1 cm
0.8 3 cm
0.6 5 cm
|T|
0.4
0.2
0
0 500 1000 1500 2000
f (Hz)
Figure 3.16: Amplitude of reflection and transmission coefficients over a horizontal frac-
ture bounded by rigid permeable formations. Solid lines: permeability k0 =0.01 D, porosity
φ=16%. Dashed lines: permeability k0 =10 D, porosity φ=26%. The fracture apertures are
1 cm, 3 cm and 5 cm respectively.
and κ2 is according to (3.75). Kbf = 2.0 GPa and G = 3.0 GPa, so that
cf = 1414 m/s. Results are given in Figs. 3.13 and 3.14. We note that
over the frequency domain up to 2000 Hz the absolute values are reasonably
in agreement although the borehole flexibility gives rise to lower reflectivity
3.5. Fast Fourier transform 43
ρbf η Kf G φ k0 Dh
(kg · m−3 ) (mPa · s) (GPa) (GPa) (%) (D) (m2 · s−1 )
1000 1.0 2.0 3.0 16 0.01 0.125
1000 1.0 2.0 3.0 26 10 76.92
and higher transmittivity. For the sake of completeness, in Fig. 3.14 also the
phase values of the reflection and transmission coefficients are given. Note
that the reflected waves are largely out-of-phase, whereas the transmitted
waves tend to be fully in-phase, for higher frequencies.
The same procedure is now repeated for 1/G = 0. From the results in
Fig. 3.15, it can be seen that the Hornby approach is in agreement with our
layered approach (3.7) and (3.10), for all practical purposes.
where the ratio A− /A+ is defined by (3.7). Using the boundary condition
that p̂ is p̂0 at z = −d, where d is the distance between the borehole top and
the fracture center, yields that
−iκ z
e 1 + (A− /A+ )eiκ1 z
p̂ = pˆ0 iκ1 d . (3.87)
[e + (A− /A+ )e−iκ1 d ]
In the zone below the fracture, we have that
C+ e−iκ1 z
p̂ = pˆ0 , (3.88)
A+ [eiκ1 d + (A− /A+ )e−iκ1 d ]
where the ratio C + /A+ is defined by (3.10). The wavenumber κ1 is obtained
from (3.85) and κ2 is obtained from (3.75).
Fig. 3.17 shows 3 snapshots of the borehole pressure profile at location
above and below the fracture. For the signal at z = −52.5 mm, there is
no attenuation of the step wave because the formation is impermeable. The
reflected wave from the fracture arrives around t = 0.26 ms causing a pressure
decrease in time. For the signal at z = 37.5 mm, the wave is dispersed because
it has passed the fracture.
The modeling of formation 2 can be considered as a special case of (3.88),
where C + /A+ = 1, A− /A+ = 0 and d = 0:
where κ1 is obtained from (3.32). Note that the dynamic parameters are
included. Fig. 3.18 shows 3 snapshots of the borehole pressure profile at
location z = 0, 30, and 50 mm from the borehole top surface. It can clearly
3.5. Fast Fourier transform 45
0.6
0.4
0.2 z=−200 mm
z=−52.5 mm
0 z=37.5 mm
Figure 3.17: Step wave propagation in a borehole intersected by a fracture. The fracture
aperture h = 1 mm, the center of the fracture is 200 mm below the borehole top. The
formation is defined in Table 3.3 (Formation 1).
0.8
Pressure (bar)
0.6
0.4
0.2 z=0 mm
z=30 mm
0 z=50 mm
Figure 3.18: Step wave propagation in a borehole adjacent to formation 2 (see Table 3.3).
Dispersion of the incident step wave is clearly visible.
46 3. Tube wave propagation in a fluid-filled borehole
be seen that the attenuation causes the amplitude of the borehole wave to
decrease, but that the original pressure profile is largely maintained. The
attenuation of the tube wave is due to the radial flow from the borehole into
the formation where wave energy is dissipated by friction.
Chapter 4
Experimental setup
4.1 Introduction
Wave experiments are carried out in a vertical shock-tube (see Figs. 4.1 and
4.2). This setup was first used for porous medium research by Van der
Grinten et al. (1985, 1987). Next, research on wave propagation in partially
saturated porous media was carried out in this setup (Smeulders and van
Dongen, 1997). Also, Wisse (1999) and Chao (2005) performed shock-tube
Figure 4.1: Picture of the shock tube Figure 4.2: Shock tube within support
support structure. structure.
47
48 4. Experimental setup
2.02 m
High pressure
section
77 mm diaphragm
4.32 m
test section
probe
P1
formation formation
1.1 m
fracture
0.4 m
Sample
P2
Figure 4.3: Schematic of the shock tube set up. The fractured sample is enlarged in the
inset.
positions inside the borehole (see Fig. 4.5). The Kistler transducer P1 (603B,
see Fig. 4.3) is mounted in the shock tube wall, which is used for recording
the pressure in the shock tube. We use a four-channel LeCroy 6810 data
acquisition system. Kistler 5001 amplifier is used to increase signal voltage.
The inner diameter of the tube is 77 mm. The high-pressure section is about
2 meters long. The low-pressure section is about 5.4 meters long. The wall
thickness of the tube section is 2.5 cm. The test section and the high-pressure
section are separated by a diaphragm. The membrane (diaphragm) section
is shown in Fig. 4.6. The membrane is made of polyethylene terephthalate
(PET), with thickness 0.07 mm. The diaphragm has a red tape stuck to it to
hold the debris after rupture. The black arrow is drawn on the membrane to
provide reproducible cutting from the PET sheet. Below, in the low-pressure
section, the test sample is installed (see Fig. 4.4). All elements specified
above can be found back in this picture.
A wave experiment proceeds as follows: The pressure in the high pressure
50 4. Experimental setup
Figure 4.5: The end plate of the shock Figure 4.6: Top view of the membrane
tube, PDCR 81 transducer is installed in section. To allow unobstructed view, the
the center of the end plate, and can be high-pressure section is removed. The
moved up and down. diaphragm has a red tape stuck to it to
hold the debris after rupture. The arrow
is drawn on the membrane to provide re-
producible cutting from the PET sheet.
Figure 4.9: Schematic of the mandrel sample (left) and picture of the mandrel sample
(right).
Figure 4.10: Schematic of the composed sample (left) and picture of the composed sample
(right).
120o
360o
240o
Figure 4.11: Schematic of sample #5 (left) and picture of sample #5 (right), composed
of two parts separated by a small horizontal slit.
which are used for the experiments are given in Table 4.1. A schematic and
photograph of sample #1 is given in Fig. 4.9. Sample #1 consists of a man-
drel and a host cylinder (parts A and B), in which a 10 mm diameter borehole
is drilled. The fracture aperture is determined by shim stock holding parts
A and B distance h apart. The fracture extension length L is also given
in the Table. Sample #2 is exactly the same as sample #1 except for the
borehole diameter which is 12.5 mm here. We also used PVC sample #3,
whose schematic and photograph are given in Fig. 4.10. This sample consists
of two cylindrical parts. The lower part is 195 mm and the upper part is
396 mm. The two parts are connected by three small steel poles holding
the parts distance h apart and thus creating a horizontal slit over the entire
cylinder cross-section. The three steel poles are arranged at 120o azimuthal
separation. The slit aperture can be changed by using different lengths of the
steel poles. Sample #4 is a sandstone sample and is also specified in Table
4.1. The length of the sample is 348 mm, the borehole diameter is 12.5 mm
Table 4.1: Samples that are used for the shock tube experiments.
Sample #1 #2 #3 #4 #5
type mandrel mandrel composed unfractured composed
clean clean
material PVC PVC PVC
sandstone sandstone
k0 (D) − − − 15±2 15±2
φ (%) − − − 47.6 47.6
r (mm) 5 6.25 6.1 6.25 6.25
h (mm) 0.1−5 0.1−5 0.1−5 − 0.1−5
L (mm) 25 23.75 32.15 − 32.15
54 4. Experimental setup
and the outside sample diameter is 76.8 mm. The porosity and permeability
determination of this sample is discussed in detail in the Appendix C. Sample
#5 was made from sample #4 by cutting it into 2 pieces. One piece is 30 mm
and the other is 302 mm (see Fig. 4.11). A similar procedure for attaching
both cylinder parts to each other is used as for the non-porous sample #3.
Here as well, the slit aperture is depending on the length of the steel poles.
Chapter 5
We perform wave experiments using a vertical shock tube setup. Shock waves
are generated by the rupture of a thin membrane. In the test section, the in-
cident pressure waves generate borehole-guided waves along water-saturated
samples. The tube is equipped with side wall gages and a mobile pressure
probe, so that the attenuation and reflection of the wave can be measured.
The computation for a single horizontal fracture intersecting a vertical bore-
hole gives a quantitative prediction of reflection and transmission of borehole-
guided waves. Three different fracture apertures are used for the calculation.
Fracture aperture significantly affects both reflection and transmission coef-
ficients. Large fractures increase reflectivity and decrease transmissivity. In
the experiment, we found that both pressures above and below the fracture
are influenced by the fracture aperture indeed, thus indicating the potential
for fracture detection by borehole waves.
5.1 Introduction
Fractured reservoir is ubiquitous in hydrocarbon reservoir engineering (Chilin-
garian et al., 1992; Reza Naimi-Tajdar et al., 2007; Ramirez et al., 2009). The
Stoneley wave (St) has been used for the detection and characterization of
fracture zones (Hornby et al., 1989; Kostek et al., 1998a; Qobi et al., 2001;
Saito et al., 2004). The borehole fluid flow into the fracture and scattering at
the fracture can lead to attenuation of the St (Hornby et al., 1989; Tang and
Cheng, 1989). The relationship between St propagation and permeability was
i
This chapter has been published as a journal paper in Transport in Porous Media. 93,
263-270 (Fan and Smeulders, 2012) and is slightly modified here for consistency in notation.
55
56 5. Shock-induced borehole waves and fracture effects
2.02 m
High pressure
section
77 mm diaphragm
4.32 m
test section
probe
P1
formation formation
1.1 m
fracture
0.4 m
Sample
P2
Figure 5.1: Schematic of the shock tube set up. The fractured sample is enlarged in the
inset.
L r
fracture O
borehole
frequency attained in our experiments is some 80 kHz, which means that the
wavelength λ is always larger than 2.5 cm, so that for all practical purposes,
λ/d >> 1. We thus assume that the borehole fluid pressure is uniform across
the borehole (Tang and Cheng, 1993). The borehole wave equation is
d2 ψ
+ κ2 ψ = 0, (5.1)
dz 2
58 5. Shock-induced borehole waves and fracture effects
1
Reflection coefficient amplitude
0.8
5 mm
0.6
0.4
1 mm
0.2
0.5 mm
0
0 2000 4000 6000 8000 10000
f (Hz)
Figure 5.3: Amplitude of reflection coefficient above the fracture at three different fracture
apertures.
1
0.5 mm
Transmission coefficient
0.8
1 mm
0.6
0.4 5 mm
0.2
0
0 2000 4000 6000 8000 10000
f (Hz)
Figure 5.4: Transmission coefficient over the fracture at three different fracture apertures.
where ρf is the fluid density, and ω is the angular frequency. Borehole wave
propagation is described by
ψ = A+ eiκ1 z + A− e−iκ1 z for z < −L/2, (5.3)
+ iκ2 z − −iκ2 z
ψ=B e +B e for − L/2 < z < L/2, (5.4)
+ iκ1 z
ψ=C e for z > L/2. (5.5)
In the region z < −L/2, A+ eiκ1 z and A− e−iκ1 z represent the incident wave
propagating in the positive z direction and the reflected wave propagating in
the negative z direction, respectively. Note that κ1 is the fluid wavenumber in
the undisturbed borehole, and κ2 is the fluid wavenumber where the borehole
is intersected by the fracture. A+ is the incident amplitude, and A− is the
reflected amplitude. In the region −L/2 < z < L/2, B + and B − are the
amplitude of the waves propagating in the positive and negative z directions,
respectively. C + is the amplitude of the transmitted wave in the region
z > L/2. The fluid displacement and the pressure should be continuous at
z = L/2 and z = −L/2. The coefficients A− , B + , B − and C + can now be
calculated as a function of the incident amplitude coefficient term A+ :
A− /A+ = 2i(κ22 − κ21 )sin(κ2 L)/D, (5.6)
+ + −iκ2 L
B /A = 2κ1 (κ1 + κ2 )e /D, (5.7)
− +
B /A = 2κ1 (κ2 − κ1 )e iκ2 L
/D, (5.8)
+ + −iκ1 L
C /A = 4κ1 κ2 e /D, (5.9)
where D is given by
D = (κ1 + κ2 )2 e−iκ2 L − (κ1 − κ2 )2 eiκ2 L . (5.10)
The above equations were also found by Tang and Cheng (1993).
Assuming a rigid formation, the wavenumber in the region z < −L/2
and z > L/2 is simply κ1 = ω/cf = ω ρf /Kf , where cf is acoustic velocity
in the fluid, and Kf is the fluid bulk modulus. The wavenumber κ in the
fracture is given by the fracture dispersion equation (Tang and Cheng, 1989)
L L
κ2 tan( f¯) + f ftan(
¯ f ) = 0. (5.11)
2 2
The parameters f and f¯ can be expressed as follows:
ω2
f2 = 2 4 − κ2 , (5.12)
cf − 3 iων
ω
f¯2 = − κ2 . (5.13)
−iν
60 5. Shock-induced borehole waves and fracture effects
The kinematic viscosity ν = μ/ρf , with μ the dynamic viscosity. The effective
wavenumber κ2 at region −L/2 < z < L/2 in the borehole can now be
expressed as follows:
ω 2 H1 (κR)
κ2 = 1− (5.14)
cf κR H0 (κR)
where H0 and H1 are Hankel functions of zeroth and first orders, respectively.
This relation is derived from the assumption that an oscillatory Poiseuille flow
exists in the fracture satisfying the no-slip boundary conditions at z = ∓L/2.
We assume the borehole radius R to be 10 mm and the fluid velocity cf to be
1.5 km/s. Using equations (5.6) and (5.9), we plot the reflection coefficient
A− /A+ and the transmission coefficient C + /A+ as a function of frequency
for different fracture apertures (see Figs. 5.3, 5.4). The amplitude of the
reflection coefficient increases with fracture aperture, but decreases with fre-
quency. The trends of the transmission coefficient are obviously opposite.
1.5
1
a.u.
0.5
−0.5
−0.2 0 0.2 0.4 0.6 0.8 1 1.2
time (ms)
Figure 5.5: Recorded pressure signal P1, which is used to trigger the data acquisition
system.
2.5
2
Pressure (bar)
1.5
0.5
St
Figure 5.6: Recorded pressure signals from the mobile probe P2 at 225- and 245-mm
TVD.
the individual wave modes. Here, the focus is on the St mode only.
0.6
0.4
Pressure (bar)
0.3
0.2
0.1
−0.1
P St
−0.2
0.3 0.4 0.5 0.6 0.7
time (ms)
Figure 5.7: Recorded pressure signal from the mobile probe P2 at 230-mm TVD. The
fracture aperture is 5 mm. The fracture is at 200-mm TVD.
In that respect, our 5-mm fracture is much larger (taking borehole scaling
into consideration). For that reason also a much smaller fracture aperture
(0.1 mm) is studied in our setup (see Fig. 5.8). We also note that Tang and
Cheng (1993) modeled centimeter scale fractures that we want to compare
against our experimental results. The recorded pressure profile with a 5-mm
fracture is given in Fig. 5.7. In this figure, the P wave arrival is at 0.369 ms
and the St arrival is at 0.526 ms. The pressure between 0.423 and 0.526 ms
is dramatically decreasing because of the presence of the fracture. We also
note that in the previous Fig. (5.6) the influence of the (closed) fracture was
already visible as a small pressure rise at t ≈ 0.43 ms, indicating that a frac-
ture once created cannot easily be undone. The theoretical predictions have
transmission coefficients always below 1, which indicates that there should be
a pressure decrease over the fracture. In the experiment, the pressure drop
is almost complete, which is much larger than predicted by theory.
5.4. Conclusion 63
0.5
0 0.1mm fracture
St 5mm fracture
−0.5
0.3 0.4 0.5 0.6 0.7
P time (ms)
Figure 5.8: Recorded pressure signals from the mobile probe P2 at 90-mm TVD. The
fracture apertures are 0.1 and 5 mm. The fracture is at 200-mm TVD.
5.4 Conclusion
The computation for a single horizontal fracture intersecting a vertical bore-
hole gives a quantitative prediction of reflection and transmission of borehole-
guided waves. Three different fracture apertures are used for the calculation.
Fracture aperture significantly affects both reflection and transmission coef-
ficients, as was also predicted in previous literature.
The shock tube can generate borehole-guided waves in a broad frequency
band. Reflection from the water-sample interface and from the free water
interface can be recorded by means of a fixed pressure transducer mounted
in the wall of the shock tube above the sample in the water layer. We also
64 5. Shock-induced borehole waves and fracture effects
6.1 Introduction
The Stoneley wave has been used to detect and characterize fracture zones
(Hornby et al., 1989; Kostek et al., 1998a,b; Qobi et al., 2001; Saito et al.,
2004). In the horizontal fracture case, Stoneley waves are attenuated because
i
This chapter (except Section 6.7) has been published as a journal paper in Journal of
the Acoustical Society of America. 134 (6), 4792-4800 (Fan and Smeulders, 2013). With
respect to the publication, additional figures were added here.
65
66 6. Wave propagation over porous and fractured borehole zones
of flow into the fracture and scattering by the fracture (Tang and Cheng,
1989). The fracture zone is traditionally modeled as a fluid-filled narrow
parallel-plate channel in which fracture waves propagate. These fracture
waves carry part of the energy of the borehole wave radially outward away
from the borehole and thus attenuate the borehole wave itself. If the viscous
skin depth of the fracture wave is on the order of the fracture aperture,
these fracture waves are also attenuated by viscous effects, and thus some
of the energy of the borehole wave is dissipated. Moreover, because the
fracture zone is experienced by the borehole Stoneley wave as a zone with
a contrast in borehole impedance, Stoneley wave reflection and transmission
will be induced. The reflected Stoneley waves carry information about the
size, shape, and orientation of the fracture and are thus of potential interest
for fractured reservoir characterization.
In this paper, first theory and application of Stoneley wave propagation in
a single horizontal fracture plane are discussed. Next we describe the shock
tube facility. This facility was used previously to detect borehole surface wave
modes (Chao et al., 2004, 2006; Smeulders and van Dongen, 1997; Sniekers
et al., 1989), and is comparable with the setup used by Winkler et al. (1989).
Measurements of the wave experiments in the shock tube are presented and
discussed. Finally, conclusions are drawn.
p = ρf ω 2 ψ, (6.2)
6.2. Theoretical formulation 67
2R
Borehole
Formation Formation
r
h Viscous fluid Viscous fluid
Formation Formation
z
Figure 6.1: Borehole intersected by a single horizontal gap representing the fracture.
dψ
U= , (6.3)
dz
where ρf is the fluid density and ω is the angular frequency. We now con-
sider wave propagation and reflection in the borehole. Wave propagation is
described by
1
ψ = A+ e−iκ1 z + A− eiκ1 z for z < − h, (6.4)
2
1 1
ψ = B + e−iκ2 z + B − eiκ2 z for − h < z < h, (6.5)
2 2
1
ψ = C + e−iκ1 z for z > h. (6.6)
2
In the region z < −h/2, where A+ e−iκ1 z represents the incident wave propagat-
ing in the positive z direction and A+ is the incident amplitude coefficient,
A− eiκ1 z is the reflected wave propagating in the negative z direction and A−
is the reflected amplitude coefficient. In the region −h/2 < z < h/2, B + and
B − are the amplitude coefficients for waves propagating in the positive z and
negative z directions, respectively. In the region z > h/2, C + is the amp-
litude coefficient of the transmitted waves. As for the boundary conditions,
68 6. Wave propagation over porous and fractured borehole zones
at z = h/2 and z = −h/2, the fluid displacement and the pressure should
be continuous. We then obtain the coefficients A− , B + , B − , and C + as a
function of the incident amplitude coefficient term A+ :
where D is given by
The above equations were also found by Tang and Cheng (1993). In the
upper zone, i.e., the zone above the fracture, the pressure in the borehole is
given by
p̂ = ρf ω 2 A+ e−iκ1 z + (A− /A+ )eiκ1 z , (6.12)
where the ratio A− /A+ is defined by (6.7). Using the boundary condition
that p̂ is p̂0 at z = −d, where d is the distance between the sample top and
the fracture center, yields
−iκ z
e 1 + (A− /A+ )eiκ1 z
p̂ = pˆ0 iκ1 d , (6.13)
[e + (A− /A+ )e−iκ1 d ]
in the upper zone. In the lower zone (below the fracture), we have that
C+ e−iκ1 z
p̂ = pˆ0 , (6.14)
A+ [eiκ1 d + (A− /A+ )e−iκ1 d ]
where the ratio C + /A+ is defined by (6.10).
where Kf is the fluid bulk modulus, G is the shear modulus of the formation,
and ZR is the wall impedance, describing the pressure-velocity ratio at the
borehole wall. It can be expressed as (Chang et al., 1988)
1 k0 K1 (κr R)
= κr R , (6.16)
ZR ηR K0 (κr R)
where k0 is the permeability of the formation, η is the viscosity of the fluid,
and K1 and K0 are modified Bessel functions of first and zeroth order, re-
spectively. In the above equation, the radial wavenumber κr is given by
κ2r = −iω/Dh , (6.17)
for incompressible dynamic (Darcy) fluid motion in a rigid formation. Dh =
k0 Kf /(ηφ) is the hydraulic diffusivity, with φ the porosity of the formation.
As indicated by Kostek et al. (1998b),
for wave propagation in the fracture,
κr can simply be computed from α(ω)ω/cf , where cf is the fluid wave
speed and α(ω) is the dynamic tortuosity as defined by Johnson et al. (1987).
Therefore, in the limiting case for low frequencies, we have that
α(ω)ω iω
lim =− . (6.18)
ω→0 cf Dh
Here we have assumed that the attenuation of the borehole wave is governed
by viscous effects due to the oscillatory radial “breathing” fluid motion, where
the elasticity of the formation is of minor importance. The elasticity, how-
ever, cannot be ignored in the expression for the effective borehole fluid bulk
modulus (6.15). We have to note that this approach [Eqs. (6.15), (6.16), and
(6.17)] is only valid for low frequencies, where the Stoneley wave becomes a
tube wave. The wavenumber and velocity of the tube wave are now easily
given by
ω ρf
κ= =ω . (6.19)
cT Keff
For a borehole radius of 10 cm, a fluid bulk modulus Kf = 2.0 GPa, and
a fluid density is ρf = 1000 kgm−3 , we compute the phase speed and the
attenuation of the tube wave as a function of frequency, for three different
porosity-permeability combinations. For the shear modulus we use G = 3.0
GPa. Results are plotted in Fig. 6.2. Curves “a” refer to a porosity of
16% and a permeability of 10 mD, curves “b” to a porosity of 20% and a
permeability of 100 mD, and curves “c” to the highest values for porosity and
permeability of 26% and 1000 mD, respectively. The dependence on porosity
and permeability is clearly visible. The higher the permeability, the more
70 6. Wave propagation over porous and fractured borehole zones
1200
1000
Phase Velocity (m/s)
a
800
b
600
c
400
200
0 −2 0 2 4 6
10 10 10 10 10
f (Hz)
1
10
0
10
Attenuation (m−1)
−1
10 c
b
−2
10 a
−3
10
−4
10 −2 0 2 4 6
10 10 10 10 10
f (Hz)
Figure 6.2: Phase velocity and attenuation of the tube wave along a porous formation.
a: k0 =10 mD, φ=16%, b: k0 =100 mD, φ=20%, c: k0 =1000 mD, φ=26%.
−iωρf A
p(r, z) = H0 (κr r) cos(f z), (6.23)
1 + 43 iων/c2f
where H0 and H1 are Hankel functions of zeroth order and first order, respect-
ively, and ν = η/ρf is the kinematic viscosity. A and D are dimension-full
arbitrary constants. The modified wavenumbers f and f¯ are expressed as
follows:
ω2
f2 = − κ2r , (6.24)
c2f + 43 iων
ω
f¯2 = − κ2r . (6.25)
iν
The wavenumber κr in the fracture is given by the dispersion equation (Tang
and Cheng, 1989)
h h
κ2r tan( f¯) + f f¯ tan( f ) = 0. (6.26)
2 2
This equation is solved numerically, and the resulting pressure and velocity
field in the fracture is obtained from (6.21) and (6.23). Next, averaging over
the fracture opening is performed and the fracture impedance is calculated
from (6.20):
iκr c2f ρf H0 (κr R)
ZR = . (6.27)
ωH1 (κr R)
72 6. Wave propagation over porous and fractured borehole zones
1200
a
1000
Phase Velocity (m/s)
800
b
600
400
c
200
0 1 2 3 4 5
10 10 10 10 10
f (Hz)
1
10
c
Attenuation (m 1)
0
−
10
b
−1 a
10
−2
10 1 2 3 4 5
10 10 10 10 10
f (Hz)
Figure 6.3: Phase velocity and attenuation of the tube wave along a fracture. Fracture
apertures are a: 1 μm, b: 10 μm, c: 100 μm.
The effective bulk modulus and hence the borehole wavenumber are now
known. For the same borehole configuration as before, the results for the
phase velocity and attenuation are plotted in Fig. 6.3, for identical borehole
parameters as before, and fracture apertures 1 μm, 10 μm and 100 μm, re-
spectively. From the figures, it becomes clear that similar trends are observed
6.4. Experimental setup 73
as for the porous reservoir. The larger the aperture, the larger the attenu-
ation becomes, which corresponds with larger permeability effects. However,
the absolute attenuation values are significantly larger for the fracture case,
at corresponding frequencies. This isbecause the fracture apertures are on
the order of the viscous skin depth 2η/(ρf ω) in these examples. Other,
perhaps more complete theories have been derived by Korneev (2008, 2010),
who relaxes the assumption of wall rigidity in the derivation of the dispersion
equation for fracture waves.
61 mm
2.02 m
High pressure
77 mm
section 12.6 mm
77 mm diaphragm
4.32 m
mm
water
tube
P1
1.1 m
water fracture
funnel
P2
L
Shock
fracture
probe
L
formation formation
Sample
P2
Figure 6.4: Schematic of the shock tube setup. An acoustic funnel is installed on top of
the sample for borehole wave enhancement (see inset for details).
74 6. Wave propagation over porous and fractured borehole zones
the borehole. In order to enhance the excitation of borehole waves and sup-
press body wave generation in the sample itself, an acoustic funnel (see inset
of Fig. 6.4) is installed approximately 1 mm above the top of the sample,
consisting of a thick-walled cylinder with decreasing internal cross-sectional
open area in the downward direction.
A wave experiment proceeds as follows: The pressure in the high pressure
section is increased to 1 to 5 bars. Rupture of the diaphragm is caused by
means of an electric current pulse. A shock wave in air is generated which
travels downward and is transmitted into the water layer (see Fig. 6.4). The
wave is partially reflected and partially transmitted into the funnel and the
borehole. The pressure at different positions inside the borehole is meas-
ured by P2. The shock tube wall is equipped with pressure transducer P1,
which is used to trigger the data acquisition system. By repeating the wave
experiments, we can measure the full pressure profiles in the borehole.
2.5
1.5
Pressure (bar)
0.5
−0.5
−1
−0.2 0 0.2 0.4 0.6 0.8
time (ms)
Figure 6.5: Recorded pressure signal P1, which is used to trigger the data acquisition
system. At t = 0, the incident step wave arrives. Between t = 0.37 and 0.54 ms, the
combined reflections from within the funnel cause a gradual pressure increase.
2.5
1.5 a
Pressure (bar)
1 b
0.5
c
0
−0.5
−1
0.2 0.3 0.4 0.5 0.6 0.7 0.8
time (ms)
Figure 6.6: Development of the measured pressure profile in the funnel: 5 mm above the
funnel (a), in the middle of the funnel (b), and 1 mm below the funnel (c).
76.5
12.2
120o
A 396
Separator
pole
360o
h
B 195
240o
Figure 6.7: Schematic of sample #3. The three separator poles are arranged at 120o
azimuthal separation. The fracture aperture is h. Dimensions are in mm.
microseismogram is shown in Fig. 6.8. The tube wave is clearly visible (St).
The slope of the line St connecting all first arrivals of the tube waves in
Fig. 6.8 corresponds with a tube wave speed of 960±40 m/s. In Fig. 6.8, also
fluid wave mode E1 is indicated which propagates with a speed of 1500 m/s.
The identification of the different wave modes was performed by using the
so-called semblance cross correlation method (Kimball and Marzetta, 1984).
This method picks wave arrivals by computing the scalar semblance in a
time window for a large number of possible arrival times and slownesses. The
maximum values of semblance are interpreted as arrivals and their associated
slownesses are plotted in a slowness-time coherence graph in Fig. 6.9. In
Fig. 6.9, the different colors stand for the different values of the coherence.
The value is from 0 (black) to 100 (white), which means that in the black
area there is no coherence and in the white area the coherence is maximal.
L φ k0 Kb G ρb ρs
Sample
(mm) (%) (D) (GPa) (GPa) (kg/m3 ) (kg/m3 )
#3 591+h 0 0 7.8 1.7 1427 1427
#4 348 47.6 15 ± 2 7.1 3.0 1310 2495
6.5. Experiment results 77
Figure 6.8: Microseismogram comprising 23 shock wave experiments in sample #3. The
fracture center is at 396.5 mm from the sample top. The fracture aperture is 1 mm. The
line St connects first arrivals of the Stoneley wave. E1 is the fluid wave arriving earlier
than the Stoneley wave and having a velocity of 1500 m/s.
In Fig. 6.9, there is also a coherence plot maximum E2 that cannot clearly
be distinguished in the microseismogram (Fig. 6.8).
Two selected pressure recordings are compared with theory in Figs. 6.10.
In Fig. 6.10(a), the position of the transducer is 67.5 mm above the fracture
center. In Fig. 6.10(b), the position of the transducer is 27.5 mm below the
fracture center. In Fig. 6.10(a), the agreement between experiment and the-
ory is very good. The amplitude of the first peak perfectly matches theory.
Note that the theoretical result stems from (6.13) and (6.14), followed by a
standard inverse FFT routine to convert the signals back to the temporal
domain. The input parameters for the theory were obtained from independ-
ent laboratory experiments and no data fitting procedure was applied. The
pressure trough A is not measured. It is associated with precursor tube mode
(Van der Grinten et al., 1987) in the input signal p0 (Fig. 6.6). In Fig. 6.10(b),
where we compare theory and experiment at some distance below the frac-
ture, the agreement between experiment and theory is even better. We note
that in Figs. 6.10(a) and 6.10(b) event E1 does not appear in the theoretical
prediction. This is because E1 is associated with a fluid bulk mode that is
not part of the theoretical description given in Section 6.2.
Next, all amplitudes of the tube waves are compared with theory in
Figs. 6.11. These amplitudes are the maximum pressure values in all 23
78 6. Wave propagation over porous and fractured borehole zones
5 90
4.5 80
4 70
Slowness (s/km)
3.5
60
3
50
2.5
40
2
St 30
1.5 E1 E2 20
1
0.5 10
Figure 6.9: Coherence plot for wave identification within sample #3. Event E1 is the
fluid wave arriving earlier than the Stoneley wave (St) and having a velocity of 1500 m/s.
E2 is a second fluid wave event that cannot clearly be identified in the microseismogram.
snapshots. In both Figs. 6.11(a) and 6.11(b), the vertical lines represent
the fracture position. We note that there is a strong decrease in amplitude
caused by the presence of the fracture. Surprisingly, the decrease in amplitude
starts somewhat earlier than where the fracture is located. In Fig. 6.11(b),
the modeled amplitude also predicts this sharp decrease over the fracture
very well. The pressure level before pressure decay (1.8 bar) is in agreement
with theory. The pressure level after pressure decay is around 1.2 bar for the
experiment and around 1.4 bar for the modeling. This means that the tube
wave loses somewhat more energy over the fracture in the experiment than
predicted by theory.
Next, new separator poles were used to obtain a fracture aperture of 5
mm. The center of the fracture is now at 398.5 mm from the sample top.
Again, 23 shock wave experiments were carried out. The resulting micro-
seismogram is shown in Fig. 6.12. Apart from the arriving tube wave, also
a reflection from the fracture can now be distinguished. The line St again
connects all first arrivals of the tube wave. The velocity is determined to
be 960±40 m/s, which is in agreement with the previous velocity measure-
ments. The dotted horizontal line indicates the position of the fracture. The
slowness-time coherence is plotted in Figs. 6.13. In Fig. 6.13(a), besides the
arrival of the Stoneley wave, again event E1 can be identified. The reflected
6.5. Experiment results 79
2
Experiment
1.5 Modeling
Pressure (bar)
1
0.5 E1
−0.5 A
−1
0 0.2 0.4 0.6 0.8
time (ms)
(a) z=-67.5 mm
2
Experiment
Modeling
1.5
Pressure (bar)
E1
0.5
−0.5 A
−1
0 0.2 0.4 0.6 0.8
time (ms)
(b) z=27.5 mm
Figure 6.10: Experimental and modeled pressure signals in sample #3 with 1 mm fracture
aperture at 67.5 mm above the fracture center (a), and 27.5 mm below the fracture center
(b). E1 is fluid wave arriving earlier than the Stoneley wave (St) and having a velocity of
1500 m/s. The precursor mode A is also visible.
1.8
Pressure (bar)
1.6
1.4
1.2
1
320 340 360 380 400 420 440
Depth from the sample top (mm)
(a) Experiment
1.9
1.8
Pressure (bar)
1.7
1.6
1.5
1.4
1.3
320 340 360 380 400 420 440
Depth from the sample top (mm)
(b) Theory
Figure 6.11: Amplitude of the Stoneley wave at different positions in the borehole of
sample #3 with 1 mm fracture. The dashed vertical line is the center of the fracture and
the two solid lines are the borders of the fracture.
6.14(a), the position of the transducer is 72.5 mm above the fracture center.
In Fig. 6.14(b), the position of the transducer is 27.5 mm below the fracture
6.5. Experiment results 81
Figure 6.12: Microseismogram comprising 23 shock wave experiments in sample #3. The
fracture center is at 398.5 mm from the sample top. The fracture aperture is 5 mm. The
line St connects first arrivals of the Stoneley wave; the line RSt connects the inflection
points representing the reflected Stoneley wave. E1 is the fluid wave arriving earlier than
the Stoneley wave and having a velocity of 1500 m/s.
5 90 5 90
4.5 80 4.5 80
4 70 4 70
Slowness (s/km)
Slowness (s/km)
3.5 3.5
60 60
3 3
50 50
2.5 2.5
2
40
2
RSt 40
E1 St 30 30
1.5 1.5
1 20 20
1
0.5 10 0.5 10
0.35 0.4 0.45 0.5 0.55 0.6 0.4 0.45 0.5 0.55 0.6 0.65
time (ms) time (ms)
(a) (b)
Figure 6.13: Coherence plots for wave transmission (a) and wave reflection (b) within
sample #3. The fracture aperture is 5 mm. Events E1, St, and Rst can clearly be identified.
center. In Fig. 6.14(a), the amplitude of the first peak is perfectly predicted
by theory, and also the gradual oscillatory pressure decrease, albeit some
time lag between predicted and recorded peaks and troughs. Note that also
reflectivity from the fracture is included in the theory (see Section 6.2). In
Fig. 6.14(b), where we compare theory and experiment at some distance
82 6. Wave propagation over porous and fractured borehole zones
2
Experiment
Modeling
1.5
Pressure (bar)
0.5 E1
−0.5 A
−1
0 0.2 0.4 0.6 0.8
time (ms)
(a) z=-72.5 mm
2
Experiment
1.5 Modeling
Pressure (bar)
1
E1
0.5
−0.5
A
−1
0 0.2 0.4 0.6 0.8
time (ms)
(b) z=27.5 mm
Figure 6.14: Experimental and modeled pressure signals in sample #3 with 5 mm fracture
aperture at 72.5 mm above the fracture center (a), and 27.5 mm below the fracture center
(b). E1 is the fluid wave, and A is the precursor pressure trough.
below the fracture, the agreement between experiment and theory is also
good. Again, event E1 does not appear in the theoretical prediction because
6.5. Experiment results 83
1.8
1.4
1.2
0.8
0.6
0.4
320 340 360 380 400 420 440
Depth from the sample top (mm)
(a) Experiment
2.2
1.8
Pressure (bar)
1.6
1.4
1.2
0.8
Figure 6.15: Amplitude of the Stoneley wave at different positions in the borehole of
sample #3 with 5 mm fracture. The dashed vertical line is the center of the fracture and
the two solid lines are the borders of the fracture.
E1 is associated with a bulk water mode that is not part of the theoretical
description given in Section 6.2. It can be seen in Fig. 6.14(b) that the first
84 6. Wave propagation over porous and fractured borehole zones
Figure 6.16: Microseismogram comprising 22 shock wave experiments in sample #4. The
line St connects the first arrivals of the Stoneley wave. E1 is the fluid wave traveling with
a speed of 1500 m/s.
4 80
3.5 70
Slowness (s/km) 3 60
2.5 50
2 40
E1 St
1.5 30
1 20
0.5 10
Figure 6.17: Coherence plot for wave identification in sample #4. Both Events E1 and
St are clearly identified.
0.8
0.6
a
Pressure (bar)
0.4
b
0.2
c
−0.2
croseismogram is shown in Fig. 6.16. The slope of the line St connecting all
first arrivals of the tube waves in Fig. 6.16 corresponds with a wave speed of
86 6. Wave propagation over porous and fractured borehole zones
905±40 m/s. The slowness-time coherence is given in Fig. 6.17. In Fig. 6.17,
besides the arrival of the Stoneley wave, again fluid wave event E1 can be
distinguished.
In Fig. 6.18, one pressure snapshot at 48 mm from the sample top is
plotted. In the plot, the results are compared with theory using three different
permeabilities: 0.5 Darcy (D), 5 D, and 50 D. It can be seen from Fig. 6.18
that permeability around 5 D would accurately predict the measured pressure
curve. The permeability was also determined in an independent falling head
laboratory experiment, from which the permeability was actually found to
be 15 ± 2 D. The discrepancy between the effective permeability of 5 D and
the actual permeability of 15 D can be attributed to the fact that in the
theory so far only low-frequency viscous effects are incorporated, whereas
also high-frequency tortuosity effects need be taken into account. Moreover,
due to long residence times of the sample in the water-filled shock tube, we
measured that fouling of the borehole wall decreased permeability over time.
We thus argue that the permeability from the separate falling head test was
probably too high.
6.6 Conclusions
Tube waves are strongly affected by fractures intersecting the borehole. A
theoretical description for both porous samples and fracture zones is given
based on the introduction of an effective borehole fluid bulk modulus. This
effective fluid bulk modulus contributes to the wave attenuation through the
borehole wall impedance. This impedance can be calculated for both porous
and fracture zones adjacent to the borehole, thus predicting borehole wave
attenuation, transmission and reflection over such zones. Our shock tube
setup generates borehole tube waves that are used for porous and fracture
zone characterization. We use sample #3 to introduce and vary fractures
in a cylindrical sample. Shock wave experiments show that attenuation in
boreholes adjacent to porous zones can be predicted by theory, although
the permeability fit still has a significant discrepancy. The reflection and
transmission of borehole tube wave over 1 and 5 mm fractures are correctly
predicted by theory, thus showing the potential of borehole wave experiments
for fracture detection and characterization.
6.7. Fractured porous sample 87
0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
time(ms)
Figure 6.19: Microseismogram comprising 23 shock wave experiments in sample #5. The
fracture center is at 32.8 mm from the sample top. The fracture aperture is 5.3 mm. For
comparison see Fig. 6.16.
5 90 5
4.5 80 4.5 80
4 70 4 70
Slowness (s/km)
Slowness (s/km)
3.5 3.5
60 60
3 3
50 50
2.5 2.5
40 40
2 2
30 30
1.5 1.5
20 20
1 1
0.5 10 0.5 10
(a) (b)
Figure 6.20: Coherence plots for wave transmission (a) and wave reflection (b) within
sample #5. The fracture aperture is 5.3 mm. Events E1, St, and Rst can not be identified
(for comparison see Fig. 6.17).
plots are given in Figs. 6.20 for the incident and reflected waves, respectively.
Note that the original coherence method can only identify downgoing waves.
For reflected upgoing waves, the interpretation method was modified so that
also these wave types could be captured. In the coherence plots, no clear
identification of the Stoneley wave could be made, although in Fig. 6.19
wave-type behavior is clearly visible, albeit only in the downgoing direction.
We also notice that the downgoing disturbance is strongly attenuated, and
that there is a strong agreement with the recordings in Fig. 6.16. The fact
that the Stoneley waves have disappeared from the coherence plots is prob-
ably caused by the relatively long residence time in the shock tube which
has caused sample degradation and fouling of the pores. Unfortunately, also
in Fig. 6.19, no upgoing reflected waves can be traced. The presence of the
fracture obviously causes attenuation of the transmitted borehole waves, as
can be seen from the fact that not even the water wave is recognized any
more in the coherence plots.
Chapter 7
76.5
60
12.5
Fracture
width
400
200
Shim
stock
h
10
89
90 7. Fracture effects in mandrel sample
Fig. 6.4. The physical properties of sample #2 are shown in Table 7.1.
Probe P2 is used to measure the borehole tube waves. It was translated
axially from 160 to 270 mm from the sample top. In this interval 23 shock
wave experiments were carried out. We started with experiments where all
shim stock was removed so that the fracture was fully closed. The resulting
microseismogram is shown in Fig. 7.2. The slope of the line St connecting
all first arrivals of the tube waves in Fig. 7.2 corresponds with a tube wave
speed of 900±40 m/s. The identification of the different wave modes was
performed by using the semblance cross correlation method (Kimball and
Marzetta, 1984). This method picks wave arrivals by computing the scalar
semblance in a time window for a large number of possible arrival times and
slownesses. The maximum values of semblance are interpreted as arrivals and
their associated slownesses are plotted in a slowness-time coherence graph in
Fig. 7.3. In Fig. 7.3, the different colors stand for the different values of the
coherence. The value is from 0 (black) to 100 (white), which means that in
the black area there is no coherence and in the white area the coherence is
Probe position from sample top (mm)
St
160
170
180
190
200
210
220
230
240
250
260
270
0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
time (ms)
Figure 7.2: Microseismogram comprising 23 shock wave experiments in sample #2. The
fracture is closed. The line St connects first arrivals of the Stoneley wave.
91
5 90
4.5 80
4
Slowness (s/km) 70
3.5
3 60
2.5 50
2
St 40
1.5
30
1
20
0.5
0.4 0.45 0.5 0.55 0.6
time (ms)
St RSt
135 −67.5
145 −57.5
155 −47.5
165 −37.5
175 −27.5
185 −17.5
195 −7.5
202.5 0
210 7.5
220 17.5
230 27.5
E1
Figure 7.4: Microseismogram comprising 23 shock wave experiments in sample #2. The
fracture center is at 202.5 mm from the sample top. The fracture aperture is 5 mm. The
line St connects first arrivals of the Stoneley wave; the line RSt represents the reflected
Stoneley wave. E1 is the fluid wave arriving earlier than the Stoneley wave and having a
velocity of 1500 m/s.
maximum.
Next, shim stock was added to obtain a fracture aperture of 5 mm. The
92 7. Fracture effects in mandrel sample
5 90 5 90
4.5 80 4.5 80
4 70 4 70
Slowness (s/km)
Slowness (s/km)
3.5 3.5
60 60
3 3
50 50
2.5 2.5
40 40
2 2 RSt
30 30
1.5 St 1.5
E1 20 20
1 1
10 0.5 10
0.5
0.35 0.4 0.45 0.5 0.55 0.45 0.5 0.55 0.6 0.65
time (ms) time (ms)
(a) (b)
Figure 7.5: Coherence plots for wave transmission (a) and wave reflection (b) within
sample #2. The fracture aperture is 5 mm. Events E1, St, and RSt can clearly be identified.
center of the fracture is now at 202.5 mm from the sample top (see Fig. 7.1).
Again, 23 shock wave experiments were carried out with probe P2 each time
at a different axial position along the borehole. The resulting microseismo-
gram is shown in Fig. 7.4. Apart from the arriving tube wave, now also
a reflection from the fracture can be distinguished. The velocity is now
determined to be 921±40 m/s, which is in agreement with the previous ve-
locity measurements, within experimental accuracy. The dotted horizontal
line indicates the position of the fracture. The maximum values of semblance
are interpreted as arrivals and their associated slownesses are plotted in the
slowness-time coherence graph in Fig. 7.5. The incident Stoneley wave in
Fig. 7.5(a) and the reflected Stoneley wave from the fracture in Fig. 7.5(b)
have the same speed which is identical to the Stoneley wave speed in Fig. 7.4,
within experimental accuracy.
Two selected pressure recordings are compared with theory in Figs. 7.6.
In Fig. 7.6(a), the position of the transducer is 62.5 mm above the fracture
center. In Fig. 7.6(b), the position of the transducer is 27.5 mm below the
fracture center. In Fig. 7.6(a), the agreement between experiment and the-
ory is very good. The amplitude of the first peak perfectly matches theory.
Note that the theoretical result stems from (6.13) and (6.14), followed by a
standard inverse FFT routine to convert the signals back to the temporal
domain. Also reflectivity from the fracture is included in the theory (see
Section 6.2). The pressure trough A is not measured. It is associated with
precursor tube mode (Van der Grinten et al., 1987) in the input signal p0
(Fig 6.6). In Fig. 7.6(b), where we compare theory and experiment at some
distance below the fracture, the agreement between experiment and theory
93
2
Experiment
Modeling
1.5
Pressure (bar)
1
0.5
−0.5 A
−1
0 0.2 0.4 0.6 0.8
time (ms)
(a) z=-62.5 mm
2
Experiment
Modeling
1.5
Pressure (bar)
1
E1
0.5
−0.5
A
−1
0 0.2 0.4 0.6 0.8
time (ms)
(b) z=27.5 mm
Figure 7.6: Experimental and modeled pressure signals in sample #2 with 5 mm fracture
aperture at 62.5 mm above the fracture center (a), and 27.5 mm below the fracture center
(b). E1 is the fluid wave, and A is the precursor pressure trough.
is less good. Event E1 does not appear in the theoretical prediction because
E1 is associated with a bulk water mode that is not part of the theoretical
94 7. Fracture effects in mandrel sample
1.8
1.6
Pressure (bar)
1.4
1.2
0.8
0.6
0.4
120 140 160 180 200 220 240
Depth from the sample top (mm)
(a) Experiment
2.2
1.8
Pressure (bar)
1.6
1.4
1.2
0.8
Figure 7.7: Amplitude of the Stoneley wave at different positions in the borehole of
sample #2 with 5 mm fracture. The dashed vertical line is the center of the fracture and
the two solid lines are the borders of the fracture.
description given in Section 6.2. It can be seen in Fig. 7.6(b) that the first
peak of the Stoneley wave is slightly overpredicted by theory, and the exper-
95
Conclusions
97
98 8. Conclusions
the dynamic wall impedance over the full frequency domain. Exact solutions
for slit flow confirm that fractures have a frequency-dependent permeability
indeed.
Straightforward computation on a single horizontal fracture gives a quant-
itative prediction of velocity, attenuation and reflection of borehole Stoneley
waves. Fracture aperture significantly affects both reflection and transmission
coefficients. Stoneley wave propagation in porous and fractured formations
is studied experimentally by means of a vertical shock tube facility. In this
set-up, shock wave in air are generated that travel downwards into a water-
saturated cylindrical rock sample that has a borehole drilled along the center
axis. In this way, high-energy borehole waves can be generated with excel-
lent repeatability. A logging probe is installed in the borehole to measure the
pressure profiles. We show that our shock tube facility can generate borehole
Stoneley waves in a broad frequency band under conditioned circumstances.
The wave can be analyzed for velocity and attenuation determination. Re-
flection from the water-sample interface and from the free water interface
can be recorded by means of a fixed pressure transducer mounted in the wall
of the shock tube above the sample in the water layer. The fractures in the
formation are manufactured by means of composite cylinders whose upper
and lower parts are separated by small spacer poles so that a variable ho-
rizontal fracture (slit) aperture can be obtained. Obviously these fractures
form an open connection between the borehole fluid and the fluid outside the
cylinder. Also mandrel samples are used for horizontal slits that are not open
to the fluid outside the cylinder, thus representing fractures with finite radial
extension. Wave experiments on PVC samples show that varying fracture
widths significantly alter the recorded Stoneley wave pressure signal at fixed
depth. The reflection and transmission of borehole tube waves over 1 and
5 mm fractures are correctly predicted by theory, thus showing the poten-
tial of borehole wave experiments for fracture detection and characterization.
Other wave experiments show that attenuation in boreholes adjacent to por-
ous zones can be predicted by theory, although the permeability fit still has
a significant discrepancy. The technique is easily extensible for detecting the
presence of multiple fractures in rock samples, and has a potential for per-
meability prediction of fractured porous samples. This research has already
successfully been applied for fracture evaluation of carbonate samples.
Appendix A
99
100 A. Derivation of the plane fracture wave equation
! "
¯ h¯ h kcos( h2 f ) f¯cos( h2 f¯)
+2f sin( f ) −2f cos( f )
2 2 −f sin( h2 f ) ksin( h2 f¯)
h h
4f f¯sin( f¯)cos( f )] = 0.
2 2
From the above equations, we find that
! 2
k sin( h2 f )cos( h2 f¯) + f fsin(
¯ h ¯ h
2 f )cos( 2 f ) = 0
k 2 cos( h2 f )sin( h2 f¯) + f f¯sin( h2 f )cos( h2 f¯) = 0,
so that
!
k 2 tan( h2 f ) + f f¯tan( h2 f¯) = 0
k 2 tan( h2 f¯) + f f¯tan( h2 f ) = 0.
As the first equation does not have zeros in the complex plane, the second
characteristic relation remains to be solved.
Appendix B
Coherence method
An artificial signal consisting of two wave types was generated to test the
coherence method. The frequency of the first wave is 50 kHz, that of the
second wave is 20 kHz. We loaded 10 traces into the program. The distance
between two traces is 5 cm (see Fig. B.1). The velocity of the first wave is
2500 m/s, that of the second one is 500 m/s. The identification of the different
wave modes was performed by using the so-called semblance cross correlation
method (Kimball and Marzetta, 1984). This method picks wave arrivals by
computing the scalar semblance in a time window for a large number of
Probe position from sample top (cm)
0
5
10
15
20
25
30
35
40
45 First Wave Second Wave
0 0.5 1 1.5 2
time (ms)
101
102 B. Coherence method
90
80
2
70
Second Wave
Slowness (s/km)
60
1.5
50
40
1 30
20
First Wave
0.5 10
Figure B.2: Coherence plot of the artificial signal. The two waves are clearly identified
by the program.
L At Asample η ρ g
(mm) (m2 ) (m2 ) (Pa · s) (kg/m3 ) (m/s2 )
302 3.22 · 10−3 4.51 · 10−3 10−3 998 9.8
mm. The length of the sample is 302 mm. We first dry the sample in an
oven. The dry weight of the sample was found to be 1779 g. Next, the sample
was carefully saturated with water and the buoyancy weight was measured
to be 1067.5 g. The porosity of the sample can now be obtained from
Mdry − Mbuoy (1779 − 1067.5) · 10−3
φ = 1− = 1− = 0.47±0.01 (C.1)
V · ρwater 1361.3 · 10−6 · 998
103
104 C. Porosity and permeability determination
Air tube
Water tube
h0 h
Pressure gauge
L
Porous sample
Water tank
The water head is measured as a function of time as the water flows through
the sample by means of gravity forces.
Obviously water must not escape from the side walls of the sample. There-
fore, the inner and outer walls are sealed with rubber hoses. The inner hose
(in the borehole) is pushed against the side wall by means of a small over-
pressure through an air tube (see Fig. C.1). The original water height in the
water tube h0 = 140 cm. After time t the water height reaches h. A pressure
gauge (Endress+Hauser Deltabar) is connected to the top of the sample to
determine the water height as a function of time. As the cross-sectional area
of the air tube is very small compared with that of the water tube, the flow
rate can be obtained from
At · dh
Q= , (C.2)
dt
where At is the cross-sectional area of the water tube. We have that
k0 (h + L)ρg
Q=− · · Asample , (C.3)
η L
where Asample is the cross-sectional area of the sample, k0 is the permeability
of the sample, ρ and η are the density and viscosity of the water, respectively.
Their values can be found in Table C.1. Combination of (C.2) and (C.3) gives:
1 −k0 ρgAsample
· dh = · dt. (C.4)
h+L LηAt
C.2. Permeability determination 105
−8
x 10
3.5
2.5
H (m2⋅ s)
1.5
0.5
0
0 500 1000 1500 2000 2500
time (s)
Figure C.2: Reduced water height recorded as a function of time. The slope is a measure
for permeability.
Adler, L., and P.B. Nagy (1994). Measurements of acoustic surface waves on
fluid–filled porous rocks. J. Geophys. Res., 99, 17863–17869.
Berryman, J.G. (1980). Confirmation of Biots theory. Appl. Phys. Lett., 37,
382-384.
Berryman, J.G., and H.F. Wang (1995). The elastic coefficients of double-
porosity models for fluid transport in jointed rock. J. Geophys. Res., 100,
24611-24627.
Berryman, J.G., and H.F.Wang (2000). Elastic wave propagation and atten-
uation in a double-porosity dual-permeability medium. Int. J. Rock Mech.
Min. Sci., 37, 63-78.
Beydoun, W.B., C.H. Cheng, and M.N. Toksöz (1985). Detection of open
fractures with vertical seismic profiling. J. Geophys. Res., 90, 4557–4566.
Biot, M.A. (1955). Theory of elasticity and consolidation for a porous aniso-
tropic solid. J. Appl. Phys., 26, 182–185.
107
108 BIBLIOGRAPHY
Biot, M.A., and D.G. Willis (1957). The elastic coefficients of the theory of
consolidation. J. Appl. Mech., 24, 594–601.
Brie, A., T. Endo, D.L. Johnson, and F. Pampuri (2000). Quantitative Form-
ation Permeability Evaluation from Stoneley Waves. SPE Reservoir Eval.
& Eng., 3, 109–117.
Chang, S.K., H.L. Liu, and D.L. Johnson (1988). Low-frequency tube waves
in permeable rocks. Geophysics, 53, 519–527.
Chao, G., D.M.J. Smeulders, and M.E.H. van Dongen (2004). Shock-induced
borehole waves in porous formations: Theory and experiments. J. Acoust.
Soc. Am., 116, 693-702.
Chao, G., D.M.J. Smeulders, and M.E.H. van Dongen (2006). Measurements
of shock-induced guided and surface acoustic waves along boreholes in
poroelastic materials. J. Appl. Phys., 99, 094904.
Cheng, C.H., and M.N. Toksöz (1981). Elastic wave propagation in a fluid-
filled borehole and synthetic acoustic logs. Geophysics, 46, 1042–1053.
Cheng, C.H., J. Zhang, and D.R. Burns (1987). Effects of in-situ permeability
on the propagation of Stoneley (tube) waves in a borehole. Geophysics, 52,
1279–1289.
Chilingarian, G.V., S.J. Mazzullo, and H.H. Rieke (1992). Carbonate Reser-
voir Characterization: A Geologic-Engineering Analysis, Part I. Elsevier,
Amsterdam, 1–35.
Cicerone, R.D., and M.N. Toksöz (1990). A dynamic model of tube wave
generation at fracture in hydrophone VSP data. Reservoir delineation –
Vertical Seismic Profiling Consortium, Annual Report.
Fan, H., and D.M.J. Smeulders (2012). Shock-induced borehole waves and
fracture effects. Transp Porous Med, 93, 263–270.
Fan, H., and D.M.J. Smeulders (2013). Shock-induced wave propagation over
porous and fractured borehole zones: Theory and experiments. J. Acoust.
Soc. Am., 134, 4792–4800.
Ferrazzini, V., and K. Aki (1987). Slow waves trapped in a fluid-filled infinite
crack: implication for volcanic tremor. J. Geophys. Res., 92, 9215–9223.
Gelinski, S., and C.H. Cheng (1998). Integrated borehole acoustic fracture
characterization. Proceedings, 4th SEGJ Tokyo, 111–116.
110 BIBLIOGRAPHY
Hardin, E.L., C.H. Cheng, F.L. Paillet, and J.D. Mendelson (1987). Fracture
characterization by means of attenuation and generation of tube waves in
fractured crystalline rock at Mirror Lake, New Hampshire. J. Geophys.
Res., 92, 7989–8006.
Hornby, B.E., D.L. Johnson, K.W. Winkler, and R.A. Plumb (1987). Fracture
evaluation from the borehole Stoneley wave. SEG Expanded Abstracts, 688–
691.
Hornby, B.E., D.L. Johnson, K.W. Winkler, and R.A. Plumb (1989). Fracture
evaluation using reflected Stoneley–wave arrivals. Geophysics, 54, 1274–
1288.
Hsu, K., A. Brie, and R.A. Plumb (1987). A new method for fracture iden-
tification using array sonic tools. J. Petrol. Tech., 39, 677–683.
Hsu, K., and C. Esmersoy (1992). Parametric estimation of phase and group
slownesses from sonic logging waveforms. Geophysics, 57, 978–985.
Johnson, D.L., D.L. Hemmick, and H. Kojima (1994). Probing porous media
with first and second sound. I. Dynamic permeability. J. Appl. Phys., 76,
104–114.
Korneev, V. (2008). Slow waves in fractures filled with viscous fluid. Geo-
physics, 73, N1–N7.
Kostek, S., D.L. Johnson, and C.J. Randall (1998a). The interaction of tube
waves with borehole fractures, Part I: Numerical models. Geophysics, 63,
800-808.
Kostek, S., D.L. Johnson, K.W. Winkler, and B.E. Hornby (1998b). The
interaction of the tube waves with borehole fractures, Part II: Analytical
models. Geophysics, 63, 809–815.
Lang, S.W., A.L. Kurkjian, J.H. McClellan, C.F. Morris, and T.W. Parks
(1987). Estimating slowness dispersion from arrays of sonic logging wave-
forms. Geophysics, 52, 530–544.
Liu, E., S. Crampin, and J.A. Hudson (1997). Diffraction of seismic waves by
cracks with application to hydraulic fracturing. Geophysics, 62, 253–265.
Mayes, M.J., P.B. Nagy, L. Adler, B.P. Bonner, and R. Streit (1986). Excit-
ation of surface waves of different modes at fluid-porous solid interface. J.
Acoust. Soc. Am., 79, 249–252.
Paige, R.W., J.D.M. Roberts, L.R. Murray, and D.W. Mellor (1992). Fracture
measurement using hydraulic impedance testing. SPE 24824, 605–614.
112 BIBLIOGRAPHY
Paige, R.W., L.R. Murray, and J.D.M. Roberts (1995). Field application of
hydraulic impedance testing for fracture measurement. SPE Production &
Facilities, 7–12.
Paillet, F.L., and J.E. White (1982). Acoustic modes of propagation in the
borehole and their relationship to rock properties. Geophysics, 47, 1215–
1228.
Pride, S.R., and J.G. Berryman (2003a). Linear dynamics of double-porosity
dual-permeability materials. I. Governing equations and acoustic attenu-
ation. Physical Review E, 68, 036603.01-036603.10.
Pride, S.R., and J.G. Berryman (2003b). Linear dynamics of double-porosity
dual-permeability materials. II. Fluid transport equations. Physical Review
E, 68, 036604.01-036604.10.
Qobi, L., A. Kuijper, X.M. Tang, and J. Strauss (2001). Permeability determ-
ination from Stoneley waves in the Ara group carbonates, Oman. GeoAr-
abia, 6, 649-666.
Ramirez, B., H. Kazemi, M. Al-Kobaisi, E. Ozkan, and S. Atan (2009). A
critical review for proper use of water/oil/gas transfer functions in dual-
porosity naturally fractured reservoirs: Part I. SPE Reservoir Evaluation
& Engineering, 12, 200–210.
Rayleigh, L. (1885). On waves propagated along the plane surface of an elastic
solid. Proc. London Math. Soc., s1–17, 4–11.
Rosenbaum, J.H. (1974). Synthetic microseismograms: logging in porous
formation. Geophysics, 39, 14–32.
Saito, H., K. Hayashi, and Y. Iikura (2004). Detection of formation bound-
aries and permeable fractures based on frequency-domain Stoneley wave
logs. Exploration Geophysics, 35, 45-50.
Schmitt, D.P., M. Bouchon, and G. Bonnet (1988). Full-wave synthetic acous-
tic logs in radially semiinfinite saturated porous media. Geophysics, 53,
807–823.
Schoenberg, M. (1984). Wave propagation in alternating solid and fluid layers.
Wave Motion, 6, 303–320.
Schoenberg, M., and P.N. Sen (1986). Wave propagation in alternating solid
and viscous fluid layers: Size effects in attenuation and dispersion of fast
and slow waves. Appl. Phys. Lett., 48, 1249–1251.
BIBLIOGRAPHY 113
Smeulders, D.M.J., R.L.G.M. Eggels, and M.E.H. van Dongen. (1992c). Dy-
namic permeability: reformulation of theory and new experimental and
numerical data. J. Fluid Mech., 245, 211–227.
Sniekers, R.W.J.M., D.M.J. Smeulders, M.E.H. van Dongen, and H. van der
Kogel (1989). Pressure wave propagation in a partially water–saturated
porous medium. J. Appl. Phys., 66, 4522-4524.
Tang, X.M., and C.H. Cheng (1989). A dynamic model for fluid flow in open
borehole fractures. J. Geophys. Res., 94, 7567–7576.
Tang, X.M., C.H. Cheng, and M.N. Toksöz (1991a). Dynamic permeabil-
ity and borehole Stoneley waves: A simplified Biot–Rosenbaum model. J.
Acoust. Soc. Am., 90, 1632–1646.
Tang, X.M., C.H. Cheng, and M.N. Toksöz (1991b). Stoneley–wave propaga-
tion in a fluid–filled borehole with a vertical fracture. Geophysics, 56, 447–
460.
Tang, X.M., and C.H. Cheng (1993). Borehole Stoneley wave propagation
across permeable structures. Geophysical Prospecting, 41, 165–187.
Tang, X.M., and C.H. Cheng (1996). Fast inversion of formation permeab-
ility from Stoneley wave logs using a simplified Biot–Rosenbaum model.
Geophysics, 61, 639–645.
114 BIBLIOGRAPHY
Toksöz, M.N., C.H. Cheng, and R.D. Cicerone (1992). Fracture detection and
characterization from hydrophone vertical seismic profiling data. Interna-
tional Geophysics, 51, 389–414.
Van Dalen, K.N., G.G. Drijkoningen, and D.M.J. Smeulders (2010). On wave-
modes at the interface of a fluid and a fluid-saturated poroelastic solid. J.
Acoust. Soc. Am., 127, 2240-2251.
Van Dalen, K.N., G.G. Drijkoningen, and D.M.J. Smeulders (2011a). Pseudo
interface waves observed at the fluid/porous–medium interface. A compar-
ison of two methods. J. Acoust. Soc. Am., 129, 2912–2922.
Van der Grinten, J.G., M.E.H. van Dongen, and H. van der Kogel (1985). A
shock–tube technique for studying pore–pressure propagation in a dry and
water–saturated porous medium. J. Appl. Phys., 58, 2937–2942.
Van der Grinten, J.G., M.E.H. van Dongen, and H. van der Kogel (1987).
Strain and pore pressure propagation in a water–saturated porous medium.
J. Appl. Phys., 62, 4682–4687.
Van der Hijden, J.H.M.T., and F.L. Neerhoff (1984). Scattering of elastic
waves by a plane crack of finite width. J. Appl. Mech., 51, 646–651.
Viktorov, I.A. (1967). Rayleigh and Lamb waves: physical theory and applic-
ations. Plenum Press, New York.
Winkler, K.W., H.L. Liu, and D.L. Johnson (1989). Permeability and bore-
hole Stoneley waves: Comparison between experiment and theory. Geo-
physics, 54, 66–75.
Zwikker, C., and C.W. Kosten (1941). Extended theory of the absorption of
sound by compressible wall–coverings. Physica VIII, 9, 968–978.
Zwikker, C., and C.W. Kosten (1949). Sound Absorbing Materials. Elsevier,
New York.
116 BIBLIOGRAPHY
Acknowledgements
117
118 Acknowledgements
ziker and Joost van der Neut. My thanks also go to my Chinese colleagues
Dong Yufei, He Yuanyuan and Guo Hua. I appreciate your help during my
PhD study and my stay in Delft. You helped me quickly getting used to
working in the laboratory and living in Delft. Alimzhan Zhubayev and Alex
Kirichek, the interesting discussions with you makes working in the laborat-
ory very lively. Ralph Feld, it is a pleasant to work in the same office with
you. I appreciate your help to translate Dutch letters for me. Especially, I
would thank you for translating the Summary part of my thesis into Dutch.
I also find it is interesting to teach you Chinese language.
I enjoyed the activities organized by DOGS (Delft Organization of Geo-
physics Students), e.g. the Thursday talks, the DOGS drinks, and the Com-
pany visits. I would also like to thank my alumni from China University of
Geosciences (Beijing) in the Netherlands: Yu Yanqing, Zhang Mengmeng,
Tan Shuhong, Li Tianqi, Li Jiaguang. I enjoy all the gatherings with you.
I would like to give my thanks to Wang Meng, Cui Haiyang, Xu Min, Yang
Xiaogang and other friends in Delft. It is nice to meet you in Delft and thank
you for your help to me and my family.
Finally, I address my appreciation to my families. I appreciate the great
help and encouragement from my parents and parents-in-law during my PhD
study. Thank you for coming to the Netherlands when my son was born. I
would like to thank my wife, Liao Fang. Dear, thank you for your continuous
support and encouragement since we met in the first year of our university
life in Beijing. I also appreciate it so much that you gave up the job oppor-
tunity in the USA after you obtained a MSc degree there and joined me in
the Netherlands. I am grateful for your company during the good time and
also the tough time in the Netherlands. My little son Fan Yuxiu, you bring
me so much joy. Your crying, laughing and your struggling to talk are a gift
for me. I enjoy every moment with you. I love you so much!
Huajun Fan
Delft, May, 2014
Curriculum Vitae
Personal information
Huajun Fan was born on 7 September 1983 in Suichang, Zhejiang Province
of China. He went to China University of Geosciences (Beijing) in 2001 to
study as a Bachelor student and obtain his BSc in Exploration and Engin-
eering in 2005. In the same year, he started as a Master student at the same
university. From March 2006 to October 2006, he interned as a geophysicist
in the Research Institute of Petroleum Exploration & Development (RIPED)
of PetroChina in Beijing. He obtained his MSc in Geological Engineering in
2008, and afterward became a PhD student at the Faculty of Civil Engin-
eering and Geosciences of Delft University of Technology. Since 2013, he has
been working in a project cooperating with RIPED as a research scientist at
Delft University of Technology.
Journal publications
• Fan, H., and D.M.J. Smeulders (2013). Shock–induced wave propaga-
tion over porous and fractured borehole zones: Theory and experi-
ments. Journal of the Acoustical Society of America, 134, 4792–4800.
• Fan, H., and D.M.J. Smeulders (2012). Shock–induced borehole waves
and fracture effects. Transport in Porous Media, 93, 263–270.
Conference publications
• Fan, H., and D.M.J. Smeulders (2013). Interaction between fracture
zones and shock–induced borehole waves. Poromechanics V, 267–275.
• Fan, H., and D.M.J. Smeulders (2012). Fracture evaluation using
shock–induced borehole waves. 28th International Symposium on Shock
Waves, 805–810.
119
120 Curriculum Vitae