You are on page 1of 8

Powder Technology 192 (2009) 195202

Contents lists available at ScienceDirect



Powder Technology

journal homepage: www. el sevier. com/l oca te/ powtec




Experimental study and mathematical model of nanoparticle transport in porous
media

Binshan Ju , Tailiang Fan

School of Energy Resources, Key Laboratory of Marine Reservoir Evolution and Hydrocarbon Accumulation Mechanism, Ministry of Education, China University of Geosciences (Beijing), Beijing
100083, China


a r t i c l e i n f o

Article history:
Received 22 August 2008
Received in revised form 7 November 2008
Accepted 22 December 2008
Available online 7 January 2009

Keywords:
Nanoparticle
Oil recovery
Water injection
Porous media
Mathematical model


a b s t r a c t

Two types of polysilicon nanoparticles (PN) were used in oil fields to improve oil recovery and enhance water injection
respectively in this work. The physical properties of the nanoparticles were studied experimentally, and pore characteristics of
sandstone were investigated by mercury injection experiments. The adsorption experiments of lipophobic and hydrophilic
polysilicon nanoparticles (LHPN) were conducted to testify wettability change (from oil wetting to water wetting) of
sandstone surface, and the nanoparticles attached to pore walls were observed bya transmission electron microscope (TEM). A
mathematical model to describe the nanoparticles transport carried by two-phase flow in random porous media was presented
and a numerical simulator was developed to simulate two application examples of the nanoparticles in oilfields. An important
discovery is that water-phase permeabilities of these sandstones increase from 1.6 to 2.1 times of their original values.
However, there are decreases in their absolute permeabilities because of nanoparticle adsorption on pore surfaces and
nanoparticle capture at pore throats. The important parameters such as the distributions of porosities and permeabilities, the
changes in water injection capability and oil recovery are obtained successfully by numerical simulation approach.
Furthermore, the permeabilities obtained from numerical simulation have a good match with experimental data. The
conclusion that polysilicon nanoparticles are effective agents for enhancing water injection capability or improving oil
recovery can be safely drawn.

2009 Elsevier B.V. All rights reserved.



1. Introduction

Nanometer particles have many special physical effects [1] and they can
be made in different ways [25]. Their applications in chemicals, metallurgy,
ceramics, medicines and other fields have been reported frequently in recent
years. Ding and Wen [6] presented a theoretical model for predicting particle
concentration and velocity fields of nanofluids flowing through a pipe. By
contrast, the flow paths in random porous media look like network
interconnected by pore throats and pore bodies. The sandstone in oil
formation can be regarded as a kind of complicated random porous media and
the transport process of nanoparticles carried by fluids in sandstone belongs to
multiphase flow.
As far as it goes, only few papers address the issues of the application of
nanopowders in oilfields to enhance water injection by virtue of changing the
wettability of reservoir rock through their adsorption on porous walls of
sandstones. Ju and Dai [7] reported that one nanometer-scale polysilicon
material could change the wettability of porous surfaces of sandstone and
consequently have effects on the flows of water and oil in oil formation when
the suspension of the nanoparticles is injected into an oil reservoir. There are
only few papers [811] dealing with mathematical modeling of fine particles
migration in formation;

Corresponding author.
E-mail address: jubs2936@163.com (B. Ju).

0032-5910/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2008.12.017



however, none of which deals with the migration and adsorption of
nanometer-scale materials in porous media.
During PN transport in porous media, there is a mass exchange between
the PN on pore framework and the PN in fluids by adhering to pore walls,
detaching from pore walls and blocking at pore throats. In addition, the
covering of PN on pore walls will change the wettability of pore surfaces. To
understand the PN transport behaviors in random porous media, theoretical
and experimental approaches were used in the present investigation and a
mathematical model for predicting PN transport performances and its effects
on two-phase flow behaviors was developed. Two examples, enhancing water
injection capacity of low permeability reservoirs and improving oil recovery
of high permeability reservoirs, were also studied by numerical simulation
approach on the simulator developed in current work.


2. Experimental studies on physical properties of the nanoparticles, random
porous media and flowability with HLPN treatment

2.1. Physical properties of the polysilicon nanoparticles

The PN in this study is a kind of modified ultra-fine powder (see Fig. 1),
which is made from SiO
2
and an additive. The shape of a nanoparticle looks
like an approximate sphere when observed under a TEM and the particle
diameters are from 10 to 500 nm (see Fig. 2).
196 B. Ju, T. Fan / Powder Technology 192 (2009) 195202















Fig. 3. Particle size distribution curve.
Fig. 1. The nanopower in a beaker.

Fig. 3 indicates that the sizes of the particles have a quasi-Gaussian
distribution. The bulk density of the nanopowder is 0.056 g/cm
3
. According
to wettability of the surface of the polysilicon nanoparticles, they can be
classified into two types: lipophobic and hydrophilic polysilicon nanoparticle
(LHPN) and hydrophobic and lipophilic polysilicon nanoparticle (HLPN).


2.2. Physical properties of random porous media (sandstone)

Sandstone oil and gas reservoir is one major type of reservoirs discovered
by the petroleum companies in the world. Therefore sandstone was selected to
be as an example of random porous media. As we know, sandstone,
composed of grains of different sizes, is porous media deposited under the
combination of consolidation and compaction through a long geological
period. Sandstone contains voids dispersed in a solid matrix and it can be
considered equivalent to a system in which the solid particles and void phases
are randomly dispersed in such a way that both phases form continuous
conducting paths through the medium. Void space is generally known as the
pore space as it is known to consist of randomly distribution pores of various
shapes and sizes. The void spaces in sandstone can be divided into pore
throats (the narrowest segments of pores) and pore chambers (the widest
segments of pores). The sizes of pore throats in sandstone are from 0.5 to 5.0
m, and the size of pore chamber is from 5.0 to 50.0 m [12,13].

In the laboratory, the physical properties of sandstone cores, obtained
from the drilling wells in H.Z.J oil field, China, are studied by mercury
injection experiments. The main data of the sandstone cores used in these
experiments are shown in Table 1. The data obtained by mercury injection
experiments are shown in Fig. 4. It indicates that the radii of most pores fall in
the range of 0.4 to 10 m. The pores of radii less than 0.4 m have little
contribution for mercury permeation.

















Fig. 2. The images of polysilicon nanoparticles observed under TEM.

2.3. Experimental studies on the effects of nanoparticles on the physical
properties of the porous media

The experimental data from the above section show that the sizes of
polysilicon particles are from 1010
0
to 5.010
2
nm, and pore radii of
sandstone are in the range of 6.010
0
6.310
4
nm. In the process of the
nanoparticle transport with flow in porous media, a particle larger than a pore
throat may block at the pore throat. Two or more than two particles that sizes
are slightly less than a pore throat may bridge at the pore throat. If the sizes of
particles are far less than pore sizes, the nanoparticles can be adhered to the
pore walls. The theoretical analysis of adsorption is given in the Ref. [7].
Generally, the wettability of polysilicon particles and the wettability of
sandstone pore walls are different. Therefore, the adsorption of polysilicon on
sandstone pore walls leads to wettability change of pore walls. The following
experiments were conducted to study the effects of polysilicon nanoparticles
on the physical properties of the porous media.


2.4. Macroscopic experiment on wettability change

In order to study the wettability change of sandstone surface, the
measurement of wetting angles was conducted. First, a rock slice sawed from
a block of sandstone was furbished on ultra-fine sand paper; then, a drop of
water was placed on the rock slice surface and the wetting angle (
1
) was
measured; finally, the rock slice was immersed in an aqueous LHPN solution
for 2 h and the wetting angle (
2
) was measured. Fig. 5 (a) indicates that the
wetting angle (
1
) is much larger than /2, while (b) shows that the wetting
angle (
2
) after LHPN treatment becomes much smaller than /2. The change
in wetting angles indicates the wettability of surface of reservoir rock can be
changed from oil-wet to water-wet by adsorbing LHPN. The image (c) of
wetting angle (
3
) of the rock immerged in pure water for 2 h shows that
3
is
slightly higher than /2. It indicates that the wettability in this case is weak
oil-wetting.

2.5. Microscopic adsorption observation under a TEM

Two furbished slices from an oil-wet sand core were first extracted with
benzene, then, one was directly observed under a TEM, the other was
observed under a TEM after it was dipped in an aqueous LHPN solution for 2
h. Fig. 6 (a) is an image of sandstone surface without adsorbing LHPN. Fig. 6
(b) shows that the adsorbed LHPN looks like a layer of white frost.


2.6. Experimental study on flowability with HLPN treatment

Dynamic core displacements were conducted to study the fluid flowability
in porous media after treatment with HLPN under the reservoir temperature
and pressure.
B. Ju, T. Fan / Powder Technology 192 (2009) 195202 197
Table 1

The parameters of sandstone cores.

Sandstone core name Depth
a
of core, m Length of Rock density, Diameter of Porosity,% Permeability, Sedimentary
sandstone core, cm g/cm
3
sandstone core, cm 10


3
m
2
microfacies
RC3-1 2271.9 2.20 2.15 2.496 21.40 352.1 River channel face
SS6-1 2365.1 2.40 2.04 2.495 25.12 150.5 Sand sheet face
DB4-1 2374.4 2.42 2.00 2.495 22.25 79.3 Distal Bar face


a
The depth of core is the location of the core before it was drilled. For instance, RC3-1 core was buried underground 2271.9 m from well head before it was drilled.


The displacement routine for each core is listed in the following steps:

(1) Heat the core holder with core sample up to a constant temperature of
80 C.
(2) Flooded by diesel oil until the flow rate reaches at a constant and the
outflow shows no water.
(3) Flooded by brine until the flow rate reaches at a constant and
the outflow shows no hydrocarbon, then record overall differential
pressure, P
i
(=P
in
P
out
) and flow rate q
iw
(ml/s).
(4) Flooded by the suspension of HLPN in diesel oil until 10 pore volumes
(PV) of the suspension pass the core.
(5) Flooded by brine again until the flow rate reaches at a constant
and the outflow shows no diesel oil, then record overall differential
pressure, P (=P
in
P
out
) and flow rate q
w
(ml/s).

As we know, the formation near the wellbore is very important where the
pressure drop of the oil-field mainly depletes [6]. The mobile oil in the
vicinity of injection well has been displaced by water, so the flow in the pores
of the reservoir rock around wellbore can be regarded as single-phase flow.
Therefore, the water injectability can be evaluated by comparing the effective
perme-ability of water, K
wb
, (= KK
rwb
), before HLPN treatment, with the
effective permeability of water, K
wa
(= KK
rwa
) after the treatment as long as
the maximum water saturation is reached (only water phase flow). The
effective permeability of water can be calculated by Darcy's law on condition
that the length (L), cross-sectional area (A) of a core, viscosity of water (
w
),
overall differential pressure ( P) and injection rate (q
w
) are measured.


K
w
=
q
A
w

w
P
L
:

The parameters of rock cores and experimental results are shown in Table
2. The data in Table 2 show that the effective permeabilities of water after
HLPN flooding are improved 1.6272.136 times for the four cores.


3. Mathematical model to describe the nanoparticle transport process in
random porous media

3.1. Assumptions

The model simulating two-phase displacement is based on the following
assumptions:

(1) The flow is one-dimensional under isothermal condition. The rock and
fluids are supposed to be incompressible.
(2) The porous media is heterogeneous.
(3) The oil and water flows in porous media follow Darcy's law and the
gravity force is neglected.
(4) The nanoparticles are discretized into n size intervals.
(5) The viscosity and density of the fluids are constant and oil and water
are Newtonian fluids.

3.2. Transport of fluids in porous media

Since the flows of oil and water in porous media follow Darcy's law, the
continuity equations of oil (o) and water (w) phases for incompressible
Newtonian flow are given by the following equation:

A

K P


_Sl _
l
x
l
_ = 0 ;l = o; w; 1 At x
l







where x is the distance from the inlet of the sand core, t is time, is the
porosity of the porous media, S
l
,
l
and P
l
are saturation, viscosity, and
pressure of phase l, respectively, and K
l
(=K
rl
) is the effective permeability of
phase l. The expression [14] for capillary force is

P
c
= P
o
P
w
= a + bs
w
=1 + cs
w
; 2

where a, b and c are empirical parameters. s
w
is water saturation.





















Fig. 4. Mercury saturation histogram and cumulative permeability contribution curve.
198 B. Ju, T. Fan / Powder Technology 192 (2009) 195202









Fig. 5. Wettability change of sandstone after absorbing LHPN.


3.3. Transport of PN in porous media

Since PN have wettabilities, LHPN exist in the water, and HLPN exist in
the oil phase. Inasmuch as the sizes of PN are in the range of 10 to 500 nm,
Brownian diffusion should be considered. Thus, the continuity equation for
size interval i of PN can be expressed as
AC AC

_

C
_


u
l
i;l
+ _S
l

i;l

_S
l
D
i;l
i;l
+ R
i;l
= 0; 3 Ax At x

x



where i =1,2n.
The initial and boundary conditions, respectively, for Eq. (3) are given by

C
i;l
= 0; t = 0; 4
C
i;l
=

C
i;l;in
;

x

= 0;

5
where C
i,l
is the volume concentration of PN in interval i in phase l, D
i,l
is the
dispersion coefficient of PN in size interval i in phase l, R
i,l
is the net rate of
loss of PN in interval i in phase l, and C
i,in
is the con-centration of the
interval i of PN in the injected fluids.

3.4. Net loss rate of PN in transport process

The pore spaces in sandstone mainly consist of interconnected pore bodies
and pore throats. For the PN transport carried by fluid stream in the porous
media, two types of particle retention in the pores may occur: deposition on
pore surfaces and blockage in pore throats. For the retained particles on pore
surfaces, they may desorb for hydrodynamic forces, and then possibly adsorb
on other sites of the pore bodies or get entrapped at other pore throats. By
modifying the Ju and Dai's model [7], R
i,l
in Eq. (3) is given by
R
i;l
=

A
i;l
+
A
i;l

; 6 At At

where
i,l
is the volume of PN i in contact with phase l available on the pore
surfaces per unit bulk volume of sandstone,
i,l
is the volume of PN i
entrapped in pore throats from phase l per unit bulk volume of sandstone due
to plugging and bridging.
According to Gruesbeck and Collins [15], there exists a critical velocity
for surface deposition, below which only particle retention occurs and above
which retention and entrainment of PN particles take place simultaneously. A
modified Gruesbeck and Collins's model for the surface deposition is
expressed by
A
i;l

d i l
u
l
C
i l
; when u
l
bu
lc


=

_
d;i;
;
lu
;
lCi;l
;

e;i;l

i;l
u
l
u
lc

; when ulNulc : 7 At




The initial condition for Eq. (7) is

i;l
= 0; t = 0: 8

In Eq. (7),
d,i,l
and
e,i,l
are rate coefficients for surface retention and
entrainment respectively of PN in interval i in the phase l, and u
lc
is the
critical velocity for the phase l to entrain particles.


The rate equation for the entrapment of the particles in interval i in pore
throats in phase l can be written as
A
i;l

=

pt;i;l
u
l
C
i;l
;

9 At

where
p,i,l
is a constant for pore throat blocking. The initial condition for Eq.
(9) is

i;l
= 0; t = 0:
10

3.5. Change of porosity and absolute permeability

Both PN deposition on the pore surfaces and blocking in pore throats may
lead to the reduction in porosity and permeability. The instantaneous porosity
is expressed by

_ = _
0
_; 11
where denotes the variation of porosity by release and retention of PN in
the porous media, and it is expressed by


_

=

i;l
+

i;l
:

12
According to Ju and Dai's model [7], the expression for the instantaneous
permeability due to the deposition and blocking of particles can thus be
written as

K = K
0
1f k
f
+ f _=_
0
&
n
; 13
where K
0
and
0
are initial permeability and porosity, K and are
instantaneous local permeability and porosity of the porous media, k
f
is a
constant for fluid seepage allowed by the plugged pores, and f is the fraction
of the original cross-sectional area open to flow.

3.6. Evaluation of relative permeability

As we know, the wettability of pore surfaces is the most important factor
to determine the relative permeability of a porous media. The PN retention in
porous media may induce wettability changes and the shape of relative
permeability curve can also be changed. If V
i,l
is the














Fig. 6. TEM images of the sandstone surfaces.
B. Ju, T. Fan / Powder Technology 192 (2009) 195202 199

Table 2
Parameters of cores and experimental results.

The core Cross-sectional Length of The effective The effective
K
wa
/K
wb
name area, A, cm
2
core, L, cm permeability permeability
measured before measured after
HLPN flooding, HLPN flooding,
Kwb, 10
3
m
2
Kwa 10
3
m
2

N1 4.91 7.91 4.434 9.471 2.136
M1 4.91 7.23 2.133 4.212 1.975
M2 4.91 7.62 4.345 8.783 2.021
M3 4.91 7.14 1.162 1.891 1.627




size interval i entrapped in pores, the total PN retention volume satisfies the
following equation:

n
V = V
i;l
: 14
i = 1;l = w;o

Supposing the spherical particles in size interval i touching each other in
the form of point contact and using the real volume of particles as the
denominator, the specific area of the particles in size interval i can be defined
as
A
t
n
3
d
2

6

s
bi
=

=
i i
= : 15 V 6
1
n
i
3
d
i
3
d
i



Supposing
i,l
is the PN volume in interval i adsorbed on the pore
surfaces and is the volume of PN in interval i entrapped in pore
i,l
throats per unit bulk volume of the porous media. PN adhered to the pore
walls first spread as a single layer, the surface area for particles in interval i is
given by

s
i
=
_

i;l
+
i;l
_
s
bi
:
16
The total surface area in contact with fluids for all the size intervals of PN
per unit bulk volume of the porous media is calculated by

n n
6


s = s
i
=

i;l
+

i;l
_
;

17


i = 1
i = 1;l = w;o
_
d
i


where is the surface area coefficient. The specific area of a sand core can be
calculated by the following empirical equation [16],

_

s = 7000_
r
:

18
K


We suppose that the relative permeabilities of water and oil phases are,
respectively, K
rwj
and K
roj
at a water saturation, S
wj
. When s s

, the total
surfaces per unit bulk volume of the porous media are



Table 3
Parameters of HLPN used for numerical simulation.
HLPN Diameter HLPN d,i,l
,
e,i,l
,
p,i,l
,

D
i,l
,

u
lc
,

composition of HLPN, concentration, cm
1
cm
1
cm
1
cm
2
s
1
cm s
1

number nm cm
3
/cm
3

C1 40 0.004 0.16 0.3 0.0128 0.00056 0.00046
C2 50 0.0065 0.2 0.24 0.02 0.00036 0.00058
C3 60 0.009 0.24 0.2 0.0288 0.00025 0.00069
C4 70 0.01045 0.28 0.17143 0.0392 0.00018 0.00081
C5 80 0.0075 0.32 0.15 0.0512 0.00014 0.00092
C6 90 0.005 0.36 0.13333 0.0648 0.00011 0.00103
C7 100 0.0035 0.4 0.12 0.08 0.00009 0.00115
C8 150 0.002 0.6 0.08 0.18 0.00004 0.00172
C9 200 0.00115 0.8 0.06 0.32 0.00002 0.0023
C10 300 0.0009 1.2 0.04 0.72 0.00001 0.00345


Fig. 7. Concentration distribution of HLPN particles along the dimensionless distance at
different injection PV.


completely covered by PN adsorbed on pore body surfaces or entrapped in
pore throats, and wettability is determined by PN. However, when s b s

, only
part of the surfaces per unit bulk volume of the porous media is occupied by
PN.
When the surfaces per unit bulk volume of the porous media are
completely occupied by PN, the relative permeabilities of water and oil
phases are taken as K
rwj
and K
roj
respectively; otherwise, the relative
permeabilities of water and oil phases are taken as a linear function of the
surfaces covered by PN, that is, when 0 b s b s

, the relative permeabilities of


water and oil are given by
K
rwjp
V
= K
rwj
+
K
rwj
V

K
rjw
s 19 s



and


KrojV Krow


K
rojp
V
= K
roj
+

s: 20 s




4. The solution method to mathematical model

The overall mathematical model is a nonlinear system that includes the
continuity equations of oil (o) and water (w) phases (Eq. (1)), the convection
diffusionadsorption equation (Eq. (3)), and a series of auxiliary equations.
The finite-difference method is used for solving the nonlinear equation
system. In this work, the Implicit-Pressure/Explicit-Saturation (IMPES)
technique was used to solve the pressuresaturation equation (Eq. (1)) and an
explicit method was employed to solve the convection-diffusion-adsorption
equation (Eq. (3)). The solving procedures are: first, the pressure
distribution is obtained by solving the mass balance equation; then, the
velocity is calculated by Darcy's law; the PN concentration distribution in
interval i is obtained by solving the convection diffusionadsorption Eq. (3);
the new porosity , absolute permeability K, relative permeabilities of oil
and water phases for
















Fig. 8. Porosity distribution along the dimensionless distance at different injection PV.
200 B. Ju, T. Fan / Powder Technology 192 (2009) 195202

Table 4
Comparisons of permeabilities between experimental and numerical results.
Core name

N101A N2B RZ1

Before treatment, Kwb, 10
3
m
2
Experiment 1.400 4.330 1.130
After treatment, Kwa, 10
3
m
2
Kwa/Kwb
Numericals 1.378 4.336 1.127
Experiment 0.570 9.470 4.100
Numericals 0.572 9.480 4.107
Experiment 0.407 2.187 3.628
Numericals 0.415 2.186 3.644




Fig. 9. Permeability distribution along the dimensionless distance at different injection PV.

each grid are calculated, and then return to the first step if maximum
simulated time is not reached.

5. Application examples and discussion

This section gives two examples concerning oil field development. Since
PN can adhere to sandstone pore walls and change the wettabilities of pore
walls, it can be used in oil field-development to enhance oil recovery. The first
example is that HLPN is used in a low permeability oil reservoir to enhance
water injection capacity. The second one is that LHPN is used in a high
permeability oil reservoir to improve oil recovery.


5.1. Examples 1: Enhancing water injection capacity of a well in a low
permeability reservoir

Water injection into a low permeability reservoir either for pressure
maintenance or for secondary oil recovery is very difficult for the following
conflicts. On one hand, injection rates must be low enough to prevent
formation damages from over pressuring and inducing unwanted fractures.
On the other hand, these rates must be high enough to make the costly fluid
injection process economic. Formation damages caused by clay minerals
(Illite, Kaolinite, Calcium montmorillonite and Sodium montmorillonite)
easily occur in low permeability reservoirs. Although some conventional
stimulations, such as hydraulic fracturing [17,18] and acidizing [19,20], are
used to improve the flow conductivity of low permeability reservoirs, the
stimulations may fail to acquire expected designation for geological
complexity. Wettability of reservoir rock pore walls can be changed into
hydrophobic by HLPN adsorption, which supplies a new approach to enhance
water injection capacity of wells in low permeability reservoirs.


Numerical simulation approach is also used in the present work to study
the transport performances of HLPN in porous media and its effects on
physical characteristics of sandstone. The parameters of

each composition of HLPN used for numerical simulation are shown in Table
3.
Fig. 7 gives the distribution of dimensionless concentration of C
1
of
HLPN from inlet (dimensionless distance is equal to 0.0) to outlet
(dimensionless distance is equal to 1.0) of the sand core at different injection
PV (1 PV is defined as the total porous volume of the simulation model at
initial time). It indicates that the wave of HLPN concentration travels toward
the outlet and the concentration curves become flatter and flatter with the
increasing injection PV.
Figs. 8 and 9 show that both the ratios of porosity (/
0
) and the ratios of
permeability (K/K
0
) decline with the increasing injection PV. For an injection
volume, the ratios of porosity and permeability are smaller at the vicinity of
inlet than those at the vicinity of outlet. The decrease in porosity and
permeability results from the HLPN adsorption on pore walls and capture in
pore throats.
Fig. 10 indicates that water injection capability (I
w
/I
w0
= K
w
K
rw
/
(K
wo
K
rw0
)) doesn't increases linearly with HLPN injection volume. The
water injection capability reaches maximum at injection volume of 1.8 PV, so
the 1.52.0 PV of injection with total concentration of 5 vol.% of HLPN is
recommended to enhance water injection capability for low permeability
oilfields. It is very difficult to measure porosity and permeability of each point
along a sand core by experimental approach; however, the average porosity of
a sand core can be obtained by experiments and the average permeability can
be calculated by Darcy's law when having experimental data. Numerical
results and experimental data are shown in Table 4.


5.2. Examples 2: Improving oil recovery of high permeability reservoirs

It is well known that the water-flood sweep efficiency in a slight water-
wetting reservoir is lower than that in a strong water-wetting one. Since
LHPN has an ability to increase the tendency of strong water-wetting by
adsorption of LHPN on porous surfaces, it can be used to improve oil
recovery in the oil fields flooded by water injection. The following simulation
example is conducted to predict production performance with injection
LHPN. The main parameters used the numerical simulation runs are shown in
Table 5.
This example clearly shows how we can use LHPN to enhance oil
recovery. The one-dimensional numerical simulator developed in this work
was used to study flooding performances displaced by LHPN solution of 5.0
vol.%.
Fig. 11 shows that the numerical results have good matches with
experimental data. Fig. 12 indicates that permeability declines











Fig. 10. The relations between water injection capability (Iw/Iw0) and HLPN injection
volume Iw/Iw0 = KwKrw/(KwoKrw0).


Table 5
Parameters for simulation.
Number of grid 40
Grid size, dx/m 0.025
Cross-sectional area/10
4
m
2
5.100
Original porosity, 0.254
Original permeability/m
2
0.300
Original saturation of oil 0.730
Viscosity of reservoir oil/mPa s 5.100
Viscosity of injection water/mPa s 0.515
Injection rate of water/10
8
m
3
s
1
1.500
Production rate of fluid / /10
8
m
3
s
1
1.500
B. Ju, T. Fan / Powder Technology 192 (2009) 195202 201















Fig. 11. The comparison of permeability ratios between experimental and numerical results.

(7) HLPN is suitable for enhancing water injection capacity for low

permeability reservoirs, and LHPN can be used to improve oil

considerably in the vicinity of injection inlet. The mechanism of this
recovery.


kind of formation damage caused by LHPN injection is same as that
Nomenclature

cased by HLPN. Fig. 13 shows that oil recovery has been improved to a

A cross sectional area, m
2


considerable extent from 52.2% to 69.8% after injection of 2.0 PV LHPN.
a,b constants for capillary pressure correlations, Pa

The recovery is improved 17.6%.



c constants for capillary pressure correlations

The simulation model is one-dimension, so the sweep efficiency is

C
i
volume concentration of interval i of PN particles, m
3
m
3


almost up to 100%. In an oil reservoir, the sweep efficiency can be only

up to 4060% due to heterogeneity and the existence of well pattern D
i
dispersion coefficient of interval i in oil phase, m
2
s
1


dead area. Therefore, the recovery is approximately improved 7.04 to d
i
diameter of interval i, m

10.56% in an oil reservoir.
f flow efficiency factor


K absolute permeability of porous media, m
2




6. Conclusion K
r
relative permeability of porous media


k
f constant for fluid seepage allowed by the plugged pores

(1) The experimental data show that the sizes of the nanoparticles in
P pressure, Pa

q injection rate or production rate, m
3
s
1


this study are in the range of 10 to 500 nm, and the diameters of
R
i
volume changing rate of PN particles in interval i in the

the nanoparticles approximately follow a normal distribution.

(2) The mercury injection tests show that the pore radii of the
water phase per unit bulk volume of the porous media


m
3
m
3
s
1


sandstone fall in the range of 6.06.310
4
nm.



S saturation

(3) The change of wetting angles indicates that the wettability of
s total surface area in contact with fluids for all particles of PN

surface sandstone can be changed from oil-wet to water-wet by


per unit bulk volume of the porous media, m
2
m
3


adsorbing LHPN.



s
v
specific area of sand core, m
2
m
3


(4) Microscopic adsorption tests imply that these nanoparticles can

be adsorbed on pore surfaces of sandstones and in turn reduce
t time, s

u flow velocity, m s
1


the pore radii.


u
wc critical velocity for water phase, m s
1


(5) An important discovery is the sandstones' effective permeabil-

ities of water increase from 1.6 to 2.1 times of their original
V total volume of retention of PN per unit bulk volume of the


porous media, m
3
m
3


values in spite of the decease in their absolute permeabilities.

i
volume of particles i of PN available on pore surfaces per

(6) The mathematical model presented in this paper is able to

simulate successfully the transport process of nanoparticles in
unit bulk volume of the porous media, m
3
m
3

i
volume of particles i of PN entrapped in pore throats per

random porous media, and numerical results have good match

with experimental data.
unit bulk volume of the porous media, m
3
m
3



x distance, m



rate constant, m
1


surface area coefficient

porosity of porous media

viscosity of fluid, Pa s

wetting angle


Subscripts

0 initial value

c critical value or capillary pressure

d deposition

e entrainment

fe flow efficiency

i one composition

pt pore throat

Fig. 12. The distribution of permeability ratio by numerical simulating approach along o oil

distance at different injection volume. w water



Fig. 13. The relations between oil recovery and injecting volume of LHPN.
202 B. Ju, T. Fan / Powder Technology 192 (2009) 195202

SI metric conversion factors
1 cm 110
2
m
1 m 110
6
m
1 nm 110
9
m
1 MPa 110
6
Pa
1 mPa 110
3
Pa

Acknowledgments

The authors would like to thank Prof. Zhian Luan, the teachers and
graduates, laboratory of displacement mechanism, U.P.C, East China, for
their partial experimental work. The authors would also like to thank Dr.
Guodong Jin, Institute of Mechanics, Chinese Academy of Sciences, and Dr.
Faisal Qureshi, Department of Physics, to thank Faisal Qureshi, Department
of Physics, Institute of Heavy Ion Physics Peking University, China, for their
proof reading and improving the English expression.


References

[1] L.D. Zhang, Preparation and Application Technology for Ultrafine Powder, Sinopec Press,
2001, pp. 2330, ISBN:780043970.
[2] G.G. Chen, G.S. Luo, J.H. Xu, J.D. Wang, Membrane dispersion precipitation method to
prepare nanoparticles, Powder Technol. 139 (2004) 180185.
[3] J.F. Chen, Y.H. Guo, F. Wang, X.M. Wang, C. Zheng, Synthesis of nanoparticles with
novel technology: high-gravity reactive precipitation, Ind. Eng. Chem. Res. 39 (2000)
948954.
[4] R.Y. Hong, J.M. Ding, G.L. Zheng, Thermodynamic and particle-dynamic studies on
synthesis of silica nanoparticles using microwave-induced plasma CVD, China
Particuology 2 (2004) 207214.
[5] J.F. Chen, L. Shao, Mass production of nanoparticles by high gravity reactive precipitation
technology with low cost, China Particuology 1 (2003) 6469.
[6] Y.L. Ding, D.S. Wen, Particle migration in a flow of nanoparticle suspensions, Powder
Technol. 149 (2005) 8492.
[7] B.S. Ju, S.G. Dai, A study of wettability and permeability change caused by adsorption of
nanometer structured polysilicon on the surface of porous media[C],

SPE77938, SPE Asia Pacific Oil and Gas Conference and Exhibition held in Melbourne,
Australia, 810 October, 2002, pp. 915926.
[8] X.H. Liu, Civian Faruk, Characterization and prediction of formation damage in two-phase
flow systems, Paper SPE25429 Presented at the Production Operations Symposium Held
in Oklahoma City, OK, U.S.A., March 2123, 1993, pp. 231238.
[9] X.H. Liu, Civan Faruk, A multiphase mud fluid infiltration and filter cake formation model,
paper SPE25215, SPE International Symposium on Oilfield Chemistry Held in New
Orleans, LA, U.S.A., March 25, 1993, pp. 607614.
[10] X.H. Liu, Civan Faruk, Formation damage and skin factor due to filter cake formation and
fines migration in the Near-Wellbore Region, paper SPE 27364, SPE Symposium on
Formation Damage Control Held in Lafayette, Louisiana, 710 February, 1994, pp. 259
265.
[11] S. Vitthal, M.M. Sharma, K. Sepehrnoori, A one-dimensional formation damage simulator
for damage due to fines migration, paper SPE 17146, SPE Formation Damage Control
Symposium held in Bakersfield, California, February 89, 1988, pp. 2936.
[12] F.A.L. Dullien, G.K. Dhawan, Characterization of pore structure by a combination of
quantitative photomicrography and mercury porosimetry, J. Colloid Interface Sci. 47
(1974) 337349.
[13] Kartic C. Khilar, Fogler Hscott, Migration of Fines in Porous Media, 1 edition, Springer,
1999, pp. 3052, SBN-13: 978-0792352846.
[14] E.C. Donaldson, N. Ewall, B. Singh, Characteristics of capillary pressure curves, J. Pet.
Sci. Eng. 6 (1991) 249258.
[15] C. Gruesbeck, R.E. Collins, Entrainment and deposition of fines particles in porous media,
Soc. Pet. Eng. J. 24 (1982) 847855.
[16] J.S. Qin, A.F. Li, Physics of Oil Reservoir, Publishing Company, U. P. C, 2001, pp.151152.
[17] M. James, B. Amar McGowen, Ziada Abdelhak, Increasing oil production by hydraulic
fracturing in the Hassi Messaoud Cambrian Formation, Algeria, SPE paper 36904, SPE
European Petroleum Conference held in Milan, Italy, 2224 October, 1996, pp. 303309.

[18] M.R. Jackson, M. Rylance, L.G. Acosta, Hydraulic fracturing of high productivity wells in
a tectonically active area, SPE paper 38608, SPE Annual Technical Conference and
Exhibition Held in San Antonio, Texas, U.S.A., 58 October, 1997, pp. 425429.

[19] K.M. Bartko, A.M. Acock, J.A. Robert, R.L. Thomas, A field validated matrix acidizing
simulator for production enhancement in sandstone and carbonates, SPE paper 38170, SPE
European Formation Damage Conference Held in The Hague, The Netherlands, 23 June,
1997, pp. 283288.
[20] C.N. Fredd, SPE, H.S. Fogler, Alternative stimulation fluids and their impact on carbonate
acidizing, SPE paper 31074, the SPE International Symposium on Formation Damage
Control Held in Lafayette, Louisiana, U.S.A., 1415 February, 1996, pp. 2122.

You might also like