You are on page 1of 16

SPE-174637-MS

Numerical Modelling of Microbial Enhanced Oil Recovery Process Under


the Effect of Reservoir Temperature, pH and Microbial Sorption Kinetics
P. Sivasankar, and Suresh Kumar Govindarajan, Petroleum Engineering Programme, Department of Ocean
Engineering, Indian Institute of Technology-Madras, Chennai, India.

Copyright 2015, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Enhanced Oil Recovery Conference held in Kuala Lumpur, Malaysia, 11–13 August 2015.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
The present work numerically investigates the effect of reservoir temperature and pH on microbial growth
and its transport within the reservoir which undergoes the reversible sorption kinetics. Further, the present
work also studies the influence of reservoir temperature and pH on changes in interfacial tension between
oil and water, capillary pressure and its impact on microscopic oil displacement efficiency. The microbe
used is strain of Bacillus sp and the nutrient supplied to microbe is molasses. For this purpose, a novel
mathematical model is developed which describes the coupled multiphase fluid flow and multispecies
reactive transport in porous media which occurs during the MEOR process. Moreover, in the present
work, the first order Monod kinetics equation is expressed as a function of temperature and pH which
dictates the microbial growth rate. The developed mathematical model is sloved numerically by finite
volume discretization technique and the results are found to be numerically stable and validated with the
experimental results. The numerical data used for validation and for numerical simulation studies are
presented. The results suggest that the oil displacement efficiency increases as the reservoir temperature
and pH approaches the optimum temperature and pH required for microbes to reach its maximum growth.
The present numerical model may be applied as an effective screening tool before the application of
MEOR process and also serves as a reservoir simulator tool to predict the performance of MEOR process.

Introduction
Microbial enhanced oil recovery (MEOR) technique has several advantages over other enhanced oil
recovery (EOR) techniques by being cost effective, easy to implement in the field and also being
environment friendly (Sen 2008; Nielsen et al. 2014). Globally, MEOR field trials have been conducted
in 17 countries and 80% of the MEOR fiels trials in USA have reported an increament in oil recovery
(Lazar et al. 2007). Experimental studies (Sun et al. 2011; Xiao et al. 2013; Arora et al. 2014) and
modelling studies (Li et al. 2011; Nielsen 2014; Desouky et al. 1996) have been widely carried out at
constant temperature and pH conditions to understand the effect of various microbial parameters on
MEOR and all studies reported an enhancement in oil recovery. It has been reported (Gudina et al. 2012;
Vaz et al. 2012) that the biosurfactants produced by Bacillus sp can be used as an alternative to the
commercial chemical surfactants. It is also been reported (Al-Sulaimani et al. 2012) that the performance
2 SPE-174637-MS

of oil recovery is increased by mixing of biosurfactants with the chemical surfactants. Despite possessing
several merits, MEOR technique is scarcely implemented at the full reservoir scale owing to lack in
understanding of interaction between the reservoir conditions and the microbial kinetic processes (Bryant
and Lockhart 2002). The reservoir conditions that affect the growth of microbes are temperature, pH,
salinity and pressure. This sensitiveness of microbe to temperature and pH conditions affects the microbial
growth rate which in turn affects the microbial concentration and its associated oil recovery during
microbial flooding process. Further, the lack in quantifying the effect of reservoir conditions on MEOR
process (Bryant and Lockhart 2002) creates uncertainty in predicting the performance of MEOR process
(Brown 2010) which causes further retardation in the implementation of MEOR technique at reservoir
scale. Hence, it is necessary to better asses the performance of MEOR process to improve its recovery
efficiency during field implementation (Sen 2008). Experimental studies (Xia et al. 2011; Gudina et al.
2012; Vaz et al. 2012, Pereira et al. 2013) have been carried out to investigate the effect of temperature
and pH on microbial growth kinetics and biosurfactant stability but the studies lacked to quantify the
coupled effect of temperature and pH on oil recovery. Recently, to quantify the effect of reservoir
temperature on the oil recovery processes and on performance of MEOR, modelling study (Sivasankar and
Suresh Kumar 2014) has been carried out at nonisothermal conditions and computed the displacement
efficiency. However, to quantify the coupled effect of temperature and pH on oil recovery processes and
on performance of the MEOR, no study has been carried out yet at least to author’s knowledge. Hence,
still there is a lack in underatnding of MEOR process under coupled effects of temperature and pH.
Therefore, in present study, the main objective is to evaluate the coupled effect of temperature and pH
on MEOR mechanisms and on its displacement efficiency. For this purpose, a one dimensonal coupled
multiphase flow, multispecies reactive transport model is developed that mathematically describes the
MEOR process under the coupled effects of temperature and pH. The developed mathematical model is
solved numerically by using finite volume discretization technique. The developed one dimensional model
in the present work is simpler to adopt and provides faster results as compared to complex multi-
dimensional simulation studies. Hence, the present one dimensional model results helps to evaluate the
performance of MEOR and analyze its feasibility before implementing MEOR technique in the field at
relatively lesser computational cost. The present MEOR model simulates the flow of water and oil along
with the transport of nutrients and microbes for different temperature and pH conditions. In the present
model, the maximum growth rate of microbe is determined from the coupled temperature and pH values
and this helps to evaluate the temperature and pH dependent microbial concentration and its related oil
recovery performance. The robustness of the developed MEOR model is checked by verifying and
validating the present model results with the experimental and analytical results.
Physical System and Governing Equations
In present study, the MEOR simulation is carried out in rock core sample having an inlet and outlet at its
left and boundary of the core. Facultative anaerobic Bacillus licheniformis microbe is used in the present
study because the Bacillus species microbes can grow at reservoir conditions (Nielsen 2010) and also their
produced biosurfactants are highly efficient in recovery of residual oil as compared to other commercially
used chemical surfactants (Pererira et al 2013). Bacillus licheniformis grows between the temperature
range of 20 °C and 60 °C with the maximum growth rate is observed at 50 °C (Rodionova et al. 2003).
The pH range at which the Bacillus licheniformis grows is between pH4 and pH9.5 with the maximum
growth rate is observed at pH5. Molasses is the electron donar and it is used as the nutrient to the microbe
in the present study because molasses favours more growth of Bacillus species microbes (Desouky et al.
1996). Before the start of the microbial flooding process, the core is water flooded which leaves the core
with the residual oil saturation of 40%. At the start of MEOR process, the rock is water wet and the
residual oil (nonwetting phase) is held within the pore due to higher capillary pressure. In the present
model, following assumptions have been incorporated:
SPE-174637-MS 3

(1) The flow of fluid is laminar and incompressible across a uniform cross section homogeneous core
sample (Nielsen 2014); (2) The core contains only the injected microbes and nutrients and it is devoid of
any other microbes or nutients; (3) Since, the temperature difference between the injected fluid and the
core is insignificant, isothermal condition prevails within the core; (3) The growth of microbe is
insensitive to salinity pressure and variations as the salinity of the injected water is mainted constant and
the pressure variation causes insinficant variation in microbial growth rate; (4) The bisurfacants produced
during the MEOR process acts at the interface between oil and water and reduces the IFT between them
(Nielsen 2014); (5) The main mechanism helps in the recovery of residual oil from the reservoir is the IFT
reduction between oil and water by the biosurfactants, while, other recovery mechanism such as fluid
diversion due to microbial plugging is considered insignificant because the plugging of microbes within
the core is neglible; (6) Since, biosurfactant is the predominant metabolite produced by the Bacillus
microbe (Bryant and Bucherfield 1989), the oil recovery is modelled due to reduction in Interfacial tension
(IFT) by the biosurfactants, while, the production of other metabolites and its influence in oil recovery is
considered negligible.

Mathematical Formulation

Multiphase Fluid Flow Model


The injection of water along with the microbes and nutrients through the inlet of the core at flow rate, Qw
[m3 s⫺1] causes the injected water to flow from inle towards outlet with the pressure as given by Eq. (1).
The initial and boundary condition that governs the temporal and spatial variation of water pressure in Eq.
(1) is represented by Eq. (2).
(1)

(2)

In Eq. (1) and Eq. (2), PW [N m–2] is the pressure of water; Pc [N m–2] is the capillary pressure; ⌽ and
k [m2] are the updated rock porosity and rock absolute permeability; ct [m2N–1] is the total compress-
ibility; krw and kro are the relative permeability of rock to water and oil; ␮W [N s m–2] and ␮o [N s m–2]
are the viscosity of water and oil; Pini [N m–2] is the initial water pressure within the core; PR [N m–2]
is the water pressure at the right boundary of the core; A [m] is the cross sectional area of the core; x [m]
is the spatial coordinate in the core along flow direction; and t [s] is the time variable. The value of Pini
and PR are taken as 1 ⫻ 105 N m2 The value of injection flow rate (QW/A) is taken as 1 ⫻ 10– 6 N m–1.
In RHS of Eq. (1), the first and second represents the water flow and oil flow, while, the third term
represents the capillary effect during the flow. In Eq. (1), PW and Pc are dependent variables and the
capillary pressure (Pc) is represented as the pressure difference between the nonwetting (oil) phase and
wetting (water) phase at their interface and it is given by Eq. (3). Further in Eq. (3), the capillary pressure
is expresed as the function of water saturarion (Sw) which in turn defines the relative permeability of oil
(Kro) and relative permeability opf water (KrW). This interrelation between Pc, Sw, Kro and KrW is
expressed using Van Genuchten model (Van Genuchten 1980) and it is represented through Eqs (3) – (5)
respectively.
(3)

(4)

(5)
4 SPE-174637-MS

In Eqs. (3) – (6), Po [N m–2] and PW [N m–2] are the pressure of oil (nonwetting) phase and water
(wetting) phase; ␳W [kg m–3] is the density of water; g [m s–2] is the gravitational acceleration; S*w is the
effective saturation; Swir is the irreducible water saturation and its value is 0.2; Sor is the residual oil
saturation; and the constants ␤, m, n are Van Genuchten parameters and their respective values are 0.5,
0.5, 2. The pressure distribution within the core due to fluid injection causes the aqueous (water) phase
to flow with the velocity as given by Darcy’s law. Equation (6) represents the aqueous (water) phase
velocity within the core due to pressure gradient developed as aresult of fluid injection.
(6)

In Eq. (6), uw [m s⫺1] is the water velocity within the core. This water velocity (uw) plays a key role
in the transport of microbes and nutrients within the core. With the increase in uw, more of the injected
microbes and nutrients is get advected over the increased length in the domain and acts as an hyperbolic
dominant transport process, while, at lesser uw, the injected microbes and nutrients undergo more
dispersion which result in lesser concentration of microbes and nutrients transported through reduced
distance in the domain and behaves as parabolic dominant transport process (Sivasankar and Suresh
Kumar, 2014). Hence, uw is an important parameter that governs the transport of nutrient and microbe.

Multispecies Transport Model


The nutrient and microbe which are injected along with the water is get transported within the core from
inlet towards the outlet and their transport processes are represented mathematically through Eqs. (7) –
(17). The transport of nutrients and microbes are coupled processes which are interlinked by the microbial
growth rate term. Equation (7) represents the transport of nutrients in the water phase within the core,
whike Eq. (8) represents the initial and boundary condition for Eq. (7). The nutreints undergo linear
sorption sorption kinetics during their transport and it is represented by the retardation factor (Rf).
(7)

(8)

(9)

In Eq. (7) and Eq. (8), N [g l–1] is the concentration of the nutrient; Ybs is the yield coefficient for
microbe; D [m2 s⫺1] is the dispersion coefficient which is the product of dispersivity, ␣ [m] and uw as
given by Eq. (9); and the expression for dsipersivity (Schulze-Makuch 2005) is given by Eq. (9). The first
and second terms in the RHS of Eq. (7) represents the dispersion and advection of nutrients, while, the
last term represents the decline in nutrient concentration due to consumption of it by the microbes. In Eq.
(7), the term (bg - bd) [s⫺1] represents the net microbial growth rate. Here, bd [s⫺1] is the microbial decay
rate and bg [s⫺1] is the microbial growth rate. The growth rate of microbe is determined by implementing
Monod kinetic’s model which is given by Eq. (10).
(10)

In Eq. (10), Kbs [g l–1] is the half saturation constant value of nutrient; bg, max [s⫺1] is the maximum
microbial growth rate of microbe. The value of Ybs and Kbs is taken as 1 and g l– 4 respectively. It is to
be noted that the bg,max is the strong function of several factors such as temperature, pH, salinity and
pressure. In present study, it is considered that the temperature and pH are the two main factors that
dominantly governs the bg,max. The relation to determine the maximum microbial growth rate as a function
of temperature and pH (Rosso et al. 1995) is expressed through Eqs. (11) – (13).
SPE-174637-MS 5

(11)

In Eq. (11), bg, opt [s⫺1] is the optimal microbial microbial growth rate attained at optimal temperature
and at optimal ph condition; ␥(T) is the growth factor depending on the temperature value; and ␥(pH) is
the growth factor depending on the pH value. The value of ␥(T) and ␥(pH) varies between 0 and 1. At
optimal temperature and at optimal pH value, the value of ␥(T) and ␥(pH) would be 1, hence, the value
of bg, max would be equivalent to bg, opt. At temperature and pH values other than thier optimal value, the
value of ␥ (T) and ␥ (pH) is varied between 0 and less than 1 which ultimately affects the value of bg, max.
This variation in value of ␥ (T) and ␥ (pH) depending on the given temperature and pH value is expressed
by Eq. (12) and Eq. (13).
(12)

(13)

In Eq. (12) and Eq. (13), T [°C] and pH are the existing temperature and pH value within the system;
Topt [°C] and pHopt are the optimal temperature and pH value at which the growth rate of microbe is
maximum; Tmin [°C] and pHmin are the minimum temperature and pH value at which the microbe can
grow; Tmax [°C] and pHmax are the maximum temperature value at which the microbe can grow. Thus,
from Eq. (7) to Eq. (13), it is modelled in such a way that the temperature and pH condition affects the
maximum microbial growth rate (bg,max) [Eqs. (11) – (13)] which in turn affects the growth rate (bg) [Eq.
(10)] and the nutrient concentration (N) in the core [Eq. (7)]. The determined microbial growth rate also
influences the microbial concentration during its transport within the core as represented through Eqs. (14)
– (17). Equation (14) models the transport of injected microbe in the water (aqueous) phase and Eq. (15)
models the volume fraction of microbes in the sorbed (sessile) phase, while, Eqs (16) and (17) represents
the initial and boundary condition for Eq. (14) and Eq. (15).
(14)

In Eq. (14), M [g l–1] is the concentration of mmicrobe in water (aqueous) phase; ␳b [kg m–3] is the
bulk density; ␭ is the volume fraction of microbes sorbed on the solid walls of grains. The first, second
and third term in the RHS of Eq. (14) represents the dispersion, advection and net growth of microbe,
while, the last term represents the decrease in microbial concentration due to time dependent sorption of
microbes on the solid grain walls that encountered during the microbial transport. This sorption kinetics
of microbe is mathematically represented by Eq. (15). It is to be noted that the Eq. (14) and Eq. (15) are
coupled equations with the unknown parameter of M and ␭, and their corresponding values are computed
by solving Eq. (14) and (15) iteratively.
(15)

(16)

(17)

From Eqs. (15) to (17), ba [s⫺1] and b’a [s⫺1] are the reversible and irreversible attachement rate of
microbes in the solid grain wall; and bdt [s⫺1] is the detachment rate of microbes from the sorbed phase.
In RHS of Eq. (15), the first term represents the attachment of microbes from the water (aqueous) phase;
the second term represents the detachment of microbes from the attached (sorbed) phase; and the last term
represents the net microbial growth in the sorbed phase. This time dependent sorption of microbes causes
6 SPE-174637-MS

plugging or clogging within the pore which leads to the reduction in porosity and permeability of the
formation and it is represented by the relations in Eq. (18) (Clement et al. 1996)
(18)

In Eq. (18), ⌽0 and k0 [m2] are the porosity and absolute permeability of the rock at the start of the
MEOR process; ⌽ and k [m2] are the updated porosity and updated absolute permeability of the rock
during the MEOR process. Hence, from Eq. (14) to Eq. (18), the concentration of microbe in water phase
[Eq. (14)] and in sorbed phase [Eq. (15)] is computed and its consequent effect on the rock property [Eq.
(18)] is also been evaluated. The concentration of microbe in the water phase undergoes metabolic activity
to produce the biosurfactants which is dependent on the yield coefficient for biosurfactants (Ybsp) and the
relation between them is represented by Eq. (19).
(19)

In Eq. (19), Bsp [g l–1] is the concentration of biosurfactant. Equation (19) expresses that depending on
the temporal change in microbial concentration in the water phase, the biosurfactant concentration in the
water phase also changes temporally. This production of biosurfactants in the water phase acts at the
interface between oil and water and reduces the interfacial tension (IFT) between them. This reduction of
IFT between oil and water depends on the maximum and minimum biosurfactant concentration. Equation
(20) provides the relation for reduction in IFT as a function of biosurfactant concentration (Bang and
Caudle 1984).
(20)

Here, Eq. (20), ␴* [mN m–1] is the updated IFT during the MEOR process; ␴max [mN m–1] and ␴min
[mN m–1] are the maximum IFT and minimum IFT observed; Bsp, max [g l–1] and Bsp, min [g l–1] are the
maximum amd minimum biosurfactant concentration; and a is the constant. This reduction in IFT during
MEOR process in turn reduces the capillary pressure acting within the core and the relation between them
is given by Eq. (21) (Bang and Caudle 1984).
(21)

In Eq. (21), Pc,max [N m–2] is the maximum capillary pressure (or) the capillary pressure at the start of
MEOR process; p1 is the propotionality constant and its value varies from 0 to 1. The reduction in
capillary pressure because of interfacial activity of biosurfactants during MEOR process causes displace-
ment of residual oil from the pore due to weaking of capillary force, thus, leaving the core with lesser
residual oil saturation. Equation (22) relates the residual oil saturation with the capillary pressure using
Van genuchten model.
(22)

Equation (22) expresses the residual oil saturation left within the core during the MEOR process. As
MEOR process proceeds, the residual oil saturation within the core decreases which successively
increases the water saturation which in turn increases the relative permeability of oil and decreases the
relative permeability water. Hence, the fractional flow of water decreases, while the fractional flow of oil
increases during the MEOR process because of enhanced displacement of residual oil from the pores. The
input data for the present numerical model is given in Table 1.
SPE-174637-MS 7

Table 1—Input data used in present model for MEOR simulation


Parameter Value Source

Tmin 20 °C Rodionova et al. (2003)


Tmax 60 20 °C Rodionova et al. (2003)
Topt 50 °C Rodionova et al. (2003)
pHmin 4 Rodionova et al. (2003)
pHmax 9.5 Rodionova et al. (2003)
pHopt 6 Rodionova et al. (2003)
bg,max 2.78 ⫻ 10–4s⫺1 Rodionova et al. (2003)
bd 1 ⫻ 10–7s⫺1 Li et al. (2011)
Ybsp 0.4 Amani et al. (2010)
⌽o 0.4 Li et al. (2011)
ko 0.94 ⫻ 10–12m2 Li et al. (2011)
ba 2.78 ⫻ 10–5s⫺1 Li et al. (2011)
ba’ 1.72 ⫻ 10–6s⫺1 Li et al. (2011)
bdt 3.56 ⫻ 10–7s⫺1 Li et al. (2011)
␳b 1.085 ⫻ 103gl–1 Li et al. (2011)
␳w 1000 Li et al. (2011)
␮o 3.92 ⫻ 10–3 Kgm–1s–1 Li et al. (2011)
␮w 1.10 ⫻ 10–3 Kgm–1s–1 Li et al. (2011)
Binj 30 gl–1 Sivasankar and Kumar (2014)
Ninj 40 gl–1 Sivasankar and Kumar (2014)
Rf 1.11 Sivasankar and Kumar (2014)
␴max 40 mNm–1 Amani et al. (2010)
␴min 2.5 mNm–1 Amani et al. (2010)

Numerical Solution
The governing pressure equation [Eq. (1)], nutrient transport transport equation [Eq.(7)] and microbial
transport equation [Eq. (14)] are solved numerically by finite volume discretization technique. The
pressure diffusion terms in [Eq. (1)], microbial and nutrient dispersion terms in [Eq.(7)] and [Eq. (14)] are
discretized by Crank-Nicholson central differencing scheme which is unconditionally stable. The first
order PDE’s representing the advective term in nureint and microbial transport equation is discretized by
fully implicit upwind discretization scheme. Numerical stability of the present model is met by adopting
suitable cell size width and time step.

Results and Discussion


Since, there is a lack of experimental and numerical results on MEOR flooding (Li et al 2011), the validity
of the present model is checked with the results of microbial core flooding experiement which conducted
in a sand core. Figure 1 provides the verification of present microbial transport model with the
experimental result of Hendry et al. (1997) and also with the numerical result of Li et al. (2011). In this
flooding experiment, microbes are continuously injected into the saturated core for 38. 4h and its relative
concentration at the outlet of the core is measured for 500 h. The result obtained from the present
developed model fairly matches with the experimental data of Hendry et al (1999) as shown in Fig. 1. Li
et al. 2011 developed a numerical model for microbial transport in porous by considering the sorption
kinetic effects such as reversible attachment, irreversible attachment and detachment of microbes during
its transport and its results are shown in Fig. 1. It is observed from Fig. 1 that the result of the present
model is found to be in good agreement with the numerical result of Li et al. (2011).
8 SPE-174637-MS

Figure 1—Verification of present model with the experimental data (Hendry et al. 1997) and numerical result (Li et al. 2011)

To understand the coupled effect of temperature and pH on microbial growth rate, in Fig. 2, the growth
factor curves for Bacillus licheniformis is plotted for range of temperature and pH values based on the
microbe Tmin, Tmax, Topt, pHmin, pHmax, and pHopt values.

Figure 2—Effect of temperature and pH on the growth rate factor of Bacillus licheniformis

Figure 2 provides the variation in microbial growth factor for different temperature and pH values. The
growth factor in shown in Fig. 2 is computed by multiplying the ␥(T) and ␥(pH) as given by Eq. (11). The
SPE-174637-MS 9

mathematical expression for ␥(T) and ␥(pH) are given in Eq. (12) and Eq. (13) respectively. It is observed
from Fig. 2 that at pH5 and at temperature value of 50 °C, the growth factor is unity, hence at the pH and
temperature condition the microbial growth (bg) will be equal to maximum microbial growth rate (bg,max).
For Bacillus licheniformis, the bg,max value is 2.78 ⫻ 10– 4 s⫺1 (Rodionova et al. 2003). At other
temperature and pH conditions, the microbial growth factor varies between 0 and 1. It is also observed
from Fig. 2 that at Topt of 50 °C, the growth factor at pH5, pH6, pH7, pH8 and pH9 are 0.81, 1, 0.88, 0.6
and 0.21 respectively. It can be inferred that at optimal condition of pH6 and at temperature of 50 °C, the
growth factor is approximately five times as growth factor at pH9 and temperature of 50 °C. Hence, the
microbial growth rate for a given temperature can be qualitatively represented as pH6 ⬎ pH7 ⬎ pH8 ⬎
pH9 ⬎. Finally, it can be concluded from Fig. 2 that the microbial growth is sensitive to coupled
temperature and pH effects, where, the growth factor varies from 0.2 to 1 even at optimal temperature
condition for microbial growth. Thus, Fig. 2 helps to estimate the expected microbial growth rate of
Bacillus licheniformis based on the reservoir temperature and pH condition during its field implementa-
tion for EOR purpose. This temperature and pH dependent microbial growth rate in turn affects the
nutrient concentration [Eq. (7)] and microbial concentration [Eq. (14)] that exists within the system and
this in turn influences the final oil recovery. Hence, it is necessary to understand the coupled effect of
temperature and pH on microbial and nutrient concentration. For this purpose, three case studies are made
based on the pH and temperature values. In Case-I, the study is conducted at pH6 and at temperature of
50 °C which is the optimal pH and temperature condition for microbial growth. In Case-II, the study is
made at pH and temperature condition which is lesser than the optimal pH and temperature conditions;
hence, in Case II, pH5 and temperature of 40 °C is considered. In Case-III, the pH and temperature are
selected higher than the optimal pH and temperature conditions; hence, the study for Case-III is conducted
at pH9 and temperature of 60 °C. The pH and temperature condition for Case-I, II, and III are selected
in such a way so that the present study covers the wide range of pH and temperature values and also to
evaluate the MEOR process at optimal and at other than optimal conditions.
Figures (3) and (4) provides the variation in relative concentration of microbe and nutrient along the
length of the core. The results shown in Fig. (3) and Fig. (4) are obtained at the end of 3h of continuous
injection of microbes and nutrients into the core. It is observed from Fg. (3) that the microbial
concentration in Case-I is higher compared to other cases because the pH and temperature condition
favours the microbial growth in Case-I than Case-II and Case-III as reported in Fig. 2. It is inferred from
Fig. (3) that the relative concentration of microbes at relative distance of 0.2, for Case-I is 52% and 68%
greater than Cae-II and Case-III condition. It is also observed that at acidic medium of pH5 and at
relatively lower temperature to Topt, the microbial concentarion is higher than the alkaline medium of pH9
and at relatively higher temperature to Topt. Hence, it is concluded from Fig. 3 that the microbial
concentration during its transport within the system is significantly affected by the coupled pH and
temperature effects. From Fig. (4), it is observed that the pH and temperature of the system marginally
affects the nutrient concentration during its transport. Hence, it can be concluded from Fig. (4) that the
variation in nutrient concentration due to the effect of temperature and pH is insignificant.
10 SPE-174637-MS

Figure 3—Effect of pH and temperature on microbial concentration during its transport

Figure 4 —Effect of pH and temperature on nutrient concentration during its transport

The variation in microbial concentration due to different temperature and pH conditions affets the
production of biosurfactant concentration within the system [Eq.(19)]. This varied biosurfactant concen-
tration at the oil-water interface in turn affects the IFT between oil and water [Eq.(20)] and consecutively
influences the capillary pressure. Figure (5) provides the variation in IFT and relative along the core for
Case-I, Case-II and Case-III respectively.
SPE-174637-MS 11

Figure 5—Effect of pH and temperature on interfacial tension during the MEOR process

The results shown in Fig. (5) are obtained at the end of 3 h of MEOR process. The IFT at the start of
MEOR process was 40 mN m–1. From Fig. 5, it is observed that at the end of MEOR procees, the IFT
near the inlet in Case-I is lesser than the Case-II and Case-III condition because of the presence of more
biosurfactant concentration at the oil-water interface in Case-I as compared to other two cases. Further,
it is also observed from Fig. 5 that at relative distance of 0.2, the IFT is reduced by 77.5%, 65% and 54%
for Case-I, II, and III condition. This illustrates that there is a maximum possibility of 20% in the variation
of IFT depending on the pH and temperature condition of the reservoir in which the MEOR is
implemented. Thus, it is concluded from Fig. 5 that the reduction of IFT between oil and water during the
MEOR process is sensitive to the pH and temperature conditions of the system and the IFT is lesser in
the case of pH and temperature which is favourable for microbial growth. This reduction in IFT
consecutively reduces the capillary pressure within the core [Eq. (21)] which helps in the displacement of
residual oil from pores.
Figure 6 provides the variation of relative capillary pressure along the relative length of the core for
different pH and temperature conditions that exists within the core. Relative capillary pressure in Fig. 6
is defined as the ratio of capillary pressure at the start of MEOR process and the capillary pressure at the
end of MEOR process. The initial capillary pressure athe start of MEOR process was 33982. 84 N m–2
and it was computed based on the Van Genuchten model given in Eq. (3). It is observed from Fig. (6) that
the relative capillary pressure in Case-I is lesser than the Case-II and Case-III conditions because of the
reduced IFT in Case-I than Case-II, III condition as explained in Fig. 5. It is also observed from Fig. 6
that the capillary pressure near the inlet is lower for any Caseof I, II, and III, while, towards the outlet the
capillary pressure remains more unaltered from the initial capillary pressure. This is because of the
presence of microbial concentration near the inlet (Fig. 3) which produes more biosurfactants and in turn
highly reduces the IFT near the inlet (Fig. 5). Thus, it is concluded from Fig. 6 that the capillary pressure
within the system is highly influenced by the existing pH and temperature conditions during the MEOR
process which ultimately helps in the displacement of trapped or capillary held residual oil from the
system thereby enhances the displacement of residual oil.
12 SPE-174637-MS

Figure 6 —Effect of pH and temperature on relative capillary pressure during MEOR process

Figure 7 provides the spatial distribution of residual oil saturation within the core for different pH and
temperature conditions. It is observed from Fig. 7 that the residual oil saturation in Case-I is lesser than
Case-II and Case-III because in Case-I, the reduction in capillary pressure causes the displacement of
capillary held residual from the pore thereby leaving the proe with the lesser residual oil saturarion. It is
also observed that for all cases near the inlet, more residual oil is displaced compared to the outlet because
of lesser capillary pressure near the inlet as explained in Fig. 6.
SPE-174637-MS 13

Figure 7—Effect of pH and temperature on spatial distribution of residual oil saturation during MEOR process

Based on the average residual oil saturation left within the core at the end of MEOR process, the
displacement efficiency of Case-I, II and III are computed as 47.64%, 44.8%, and 41.8% respectively.
Hence, a maximum of 6% difference in displacement efficiency is observed depending on pH and
temperature conditions in which the MEOR technique is implemented. It is inferred that there is decrease
in oil displacement efficiency in the reservoir when there is an increase in the difference between the
reservoir temperature, pH and optimum temperature, pH required for maximum microbial growth. Thus,
it is concluded from Fig. 7 that the residual oil saturation and the displacement efficiency of the MEOR
process is significantly influenced by the pH and temperature conditions. Further, the minimum residual
oil and the maximum displacement efficiency is obtained at pHopt and at Topt condition. The volume of
oil displaced from the pores is successively filled by the injection water; hence the water saturation within
the core increases as the MEOR process progress as shown in Fig. 8.
14 SPE-174637-MS

Figure 8 —Effect of pH and temperature on spatial distribution of water saturation during MEOR process

Hence, it is understood from the present study that the maximum oil displacement efficiency for a
reservoir at a given temperature and pH can be achieved by selection of suitable microbe which
experiences an optimum growth rate nearer to that of reservoir temperature and pH. Hence, the
characterization of microbes based on temperature and pH in addition to other existing parameters plays
a key role in the recovery of maximum residual oil from the reservoir.

Conclusions
In present study, one dimensional multiphase multispecies reactive transport model is developed that
simulates the MEOR process under different pH and temperature conditions. In the present model, the
growth rate of microbe is expressed as the function of pH and temperature which is then coupled with the
nutrient and microbial transport equation. From the present simulation study, following conclusions are
arrived:
● The growth of microbes is highly sensitive to coupled temperature and pH effects which in turn
affect the MEOR process. The sensitivity of pH on microbial growth rate at a constant temperature
is qualitatively represented as.
● The microbial concentration during its transport is significantly affected by the coupled pH and
temperature effects, while, the effect of pH and temperature on nutrient concentration is insignif-
icant.
● The reduction of IFT and capillary pressure during the MEOR process is sensitive to the pH and
temperature conditions of the system. The IFT and consecutively the capillary pressure is lesser in
the case of pH and temperature which is favourable for microbial growth.
● The residual oil saturation and the displacement efficiency of the MEOR process is significantly
influenced by the pH and temperature conditions. The minimum residual oil and the maximum
displacement efficiency is obtained at pHopt and at Topt condition.
SPE-174637-MS 15

References
Amani, H., Sarrafzdeh, M.H., Haghighi, M. et al, 2010. Comparitive study of biosurfactant producing
bacteriain MEOR applications. .J. Pet. Sci. Eng. 75: 209 –214. http://dx.doi.org/10.1016/j.pet-
rol.2010.11.008
Al-Sulaimani, H., Al-Wahaibi, Y., Al-Bahry, S. et al, 2012. Residual-oil recovery through injection of
biosurfactant, chemical surfactant, and mixtures of both under reservoir temperatures: induced-
wettability and interfacial-tension effects. SPE Res Eval & Eng. 15 (15): 15–2. SPE-158022PA.
http://dx.doi.org/10.2118/158022-PA
Arora, P., Ranade, D.R., Dhakephalkar, P.K., 2014. Development of a microbial process for the
recovery of petroleum oil from depleted reservoirs at 91–96. Bioresour. Technol. 165: 274 –278.
http://dx.doi.org/10.1016/j.biortech.2014.03.109
Bang, H.W., Caudle, B.H., 1984. Modeling of a micellar/polymer process. SPE J. 24 (6): 617–627.
SPE-9009-PA. http://dx.doi.org/10.2118/9009-PA
Bryant, R.S. and Burchfield, T. 1989. Review of microbial technology for improving oil recovery.
SPE Reserv. Eng. 4(2): 151–154. SPE-16646-PA. http://dx.doi.org/10.2118/16646-PA
Bryant, S.L., Lockhart, T.P., 2002. Reservoir engineering analysis of microbial enhanced oil recovery.
SPE Reserv Eng & Eval. 5(5): 365–374.SPE-79719-PA. http://dx.doi.org/10.2118/79719-PA
Brown, L.R., 2010. Microbial enhanced oil recovery (MEOR). Curr. Opin. Microbiol. 13 (3):
316 –320. http://dx.doi.org/10.1016/j.mib.2010.01.011
Clement, T.P., Hooker, B.S., Skeen, R.S., 1996. Macroscopic models for predicting changes in
saturated porous media properties caused by microbial growth. Ground Water 34 (5): 934 –942.
http://dx.doi.org/10.1111/j.1745-6584.1996.tb02088.x
Desouky, S.M., Abdel-Daim, M.M., Sayouh, M.H. et al, 1996. Modelling and laboratory investigation
of microbial enhanced oil recovery. J. Pet. Sci. Eng. 15(2-4): 309 –312. http://dx.doi.org/10.1016/
0920- 4105(95)00044-5
Gudiña, E.J., Pereira, J.F.B., Rodrigues, L.R. et al, 2012. Isolation and study of microorganisms from
oil samples for application in Microbial Enhanced Oil Recovery. Int Biodeterior & Biodegrad 68:
56 –64. http://dx.doi.org/10.1016/j.ibiod.2012.01.001
Hendry, M.J., Lawrence, J.R., Maloszewski, P., 1997. The role of sorption in the transport of
Klebsiella oxytoca through saturated silica sand. Ground Water 35: 574 –584. http://dx.doi.org/
10.1111/j.1745-6584.1997.tb00122.x
Lazar, I., Petrisor, I.G., Yen, T.F. 2007. Microbial enhanced oil recovery. Pet. Sci. Technol. 25(11):
1353–1366. http://dx.doi.org/10.1080/10916460701287714
Li, J., Liu, J., Trefry, M.G. et al, 2011. Interactions of microbial enhanced oil recovery processes.
Trans. Porous Med.87: 77–104. http://dx.doi.org/10.1007/s11242-010-9669-6
Nielsen, S.M. 2010. Advanced reservoir simulation of microbial enhanced oil recovery. PhD disser-
tation, Technical University of Denmark, Denmark (July 2010).
Nielsen, S.M., Nesterov, I., Shapiro, A.A., 2014. Simulations of microbial-enhanced Oil recovery:
adsorption and filtration. Transp. Porous. Med. 102: 227–259. http://dx.doi.org/10.1007/s11242-
014-0273-z
Pereira, J.F.B., Gudiña, E.J., Costa, R. et al, 2013. Optimization and characterization of biosurfactant
production by Bacillus subtilis isolates towards microbial enhanced oil recovery applications.Fuel.
111: 259 –268. http://dx.doi.org/10.1016/j.fuel.2013.04.040
Rosso, L., Lobry, J.R., Bajard, S. et al, 1995. Convenient model to describe the combined effects of
temperature and pH on microbial growth. App. Environ. Microbiol. 61 (2): 610 –616.
16 SPE-174637-MS

Rodionova, T.A., Shekhovtsova, N.V., Panikov, N.S. et al, 2003. Effect of cultivation conditions on
growth and adhesion of Bacillus licheniformis. Microbiol. 72 (4): 521–527. http://dx.doi.org/
10.1023/A:1025052908781
Schulze-Makuch, D., 2005. Longitudinal dispersivity data and implications for scaling behaviour.
Ground water. 43(3): 443–456. http://dx.doi.org/10.1111/j.1745-6584.2005.0051.x
Sen, R., 2008. Biotechnology in petroleum recovery: The microbial EOR. Prog. Energy Combust. Sci.
34(6): 714 –726. http://dx.doi.org/10.1016/j.pecs.2008.05.001
Sun, S., Zhang, Z., Luo, Y., Zhong, W. et al, 2011. Exopolysaccharide production by a genetically
engineered Enterobacter cloacae strain for microbial enhanced oil recovery. Bioresour. Technol.
102 (102): 102–110. http://dx.doi.org/10.1016/j.biortech.2011.03.005
Sivasankar, P. Suresh, Kumar, G., 2014. Numerical modelling of enhanced oil recovery by microbial
flooding under non-isothermal conditions. J. Pet. Sci. Eng. 124: 161–172. http://dx.doi.org/
10.1016/j.petrol.2014.10.008
Van Genuchten, M.T., 1980. A closed-form equation for predicting the hydraulic conductivity of
unsaturated soils. Soil Sci. Soc. Am. J. 44 (5): 892–898. http://dx.doi.org/10.2136/
sssaj1980.03615995004400050002x
Vaz, D.A., Gudiña, E.J., Alameda, E.J. et al, 2012. Performance of a biosurfactant produced by a
Bacillus subtilis strain isolated from crude oil samples as compared to commercial chemical
surfactants. Colloids Surf., B 89: 167–174. http://dx.doi.org/10.1016/j.colsurfb.2011.09.009
Xia, W., Dong, H., Yu, L. et al, 2011. Comparitive study of biosurfactant produced by microorganisms
isolated from formation water of petroleum reservoir. Colloid Surf., A. 392: 124 –130. http://
dx.doi.org/10.1016/j.colsurfa.2011.09.044
Xiao, M., Zhang, Z., Wang, J. et al, 2013. Bacterial community diversity in a low-permeability oil
reservoir and its potential for enhancing oil recovery. Bioresour. Technol. 147: 110 –116. http://
dx.doi.org/10.1016/j.biortech.2013.08.031

You might also like