ArticlePDF Available

Experimental design methodology as a tool to optimize the adsorption of new surfactant on the Algerian rock reservoir: cEOR applications

Authors:

Abstract and Figures

In this research work, a new surfactant called surf EOR ASP 5100 used in the SWCTT (single well chemical tracer test) in the Algerian oilfield and sodium dodecyl sulfate (SDS) were used for static adsorption tests. The Algerian rock reservoir has been characterized by different techniques such as SEM, XRD, XRF, BET analysis. The equilibrium was successfully verified by Langmuir isotherm and second-order ( \(R^2>95\%\) models for all concentrations and temperatures to predict the adsorption process. Furthermore, the adsorption process was found to be exothermic ( \(\Delta G^{\circ}<0\) . To quantify the minimal adsorbed quantity, a full factorial design of 2³ (8 experiments) was applied to analyze the individual effects and interactions of operational parameters using variance analysis (ANOVA), desirability method and response surface methodology. The optimal conditions obtained are as follows: the Qe value was 2.3291mg/g for the SDS surfactant at a concentration of 200ppm and temperature of \( 25 {}^{\circ}\) C, and Qe was 3.894513mg/g for EOR ASP 5100 for the concentration of 200ppm and temperature of \( 80 {}^{\circ}\) C.
Content may be subject to copyright.
DOI 10.1140/epjp/i2019-12821-9
Regular Article
Eur. Phys. J. Plus (2019) 134: 436 THE EUROPEAN
PHYSICAL JOURNAL PLUS
Experimental design methodology as a tool to optimize the
adsorption of new surfactant on the Algerian rock reservoir:
cEOR applications
Seif El Islam Lebouachera1,3, Rachida Chemini1, Mohamed Khodja2, Bruno Grassl3, Mohammed Abdelfetah Ghriga3,4,
Djilali Tassalit5, and Nadjib Drouiche6,a
1Laboratoire des Sciences du enie des Proc´ed´es Industriels, Facult´edeG´enie ecanique et du enie des Proc´ed´es Universit´e
des Sciences et de la Technologie Houari Boumediene, Bab-Ezzouar, 16111 Alger, Algeria
2SONATRACH, Direction Centrale Recherche et eveloppement, Avenue 1 Novembre, 35000, Boumerdes, Algeria
3Institut des Sciences Analytiques et de Physico-chimie pour l’Environnement et les Mat´eriaux, IPREM, UMR 5254, CNRS
Universit´e de Pau et des Pays de l’Adour, 2 avenue P. Angot, Technopˆole elioparc, 64000 Pau, France
4Laboratoire de enie Physique des Hydrocarbures, Facult´e des Hydrocarbures et de la Chimie, Universit´e M’Hamed Bougara
de Boumerdes, Avenue de l’Ind´ependance, 35000 - Boumerdes, Algeria
5Unit´edeD´eveloppement des Equipements Solaires, UDES/EPST, Centre de D´eveloppement des Energies Renouvelables,
Route Nationale no. 11, BP386, Bou Isma¨ıl, 42400, Tipaza, Algeria
6CRTSE-Division CCPM- No. 2, Bd Dr. Frantz FANON- p.o.box 140, Alger sept merveilles, 16038, Algeria
Received: 24 January 2019 / Revised: 6 June 2019
Published online: 10 September 2019
c
Societ`a Italiana di Fisica / Springer-Verlag GmbH Germany, part of Springer Nature, 2019
Abstract. In this research work, a new surfactant called surf EOR ASP 5100 used in the SWCTT (single
well chemical tracer test) in the Algerian oilfield and sodium dodecyl sulfate (SDS) were used for static
adsorption tests. The Algerian rock reservoir has been characterized by different techniques such as SEM,
XRD, XRF, BET analysis. The equilibrium was successfully verified by Langmuir isotherm and second-
order (R2>95%) models for all concentrations and temperatures to predict the adsorption process.
Furthermore, the adsorption process was found to be exothermic (ΔG<0). To quantify the minimal
adsorbed quantity, a full factorial design of 23(8 experiments) was applied to analyze the individual
effects and interactions of operational parameters using variance analysis (ANOVA), desirability method
and response surface methodology. The optimal conditions obtained are as follows: the Qevalue was
2.3291 mg/g for the SDS surfactant at a concentration of 200 ppm and temperature of 25 C, and Qewas
3.894513 mg/g for EOR ASP 5100 for the concentration of 200 ppm and temperature of 80 C.
1 Introduction
Oil production from mature reservoirs is declining nowadays and is often low compared to the growing oil demands
around the world, leading to an increased interest toward Enhanced Oil Recovery (EOR) technologies such as surfactant
flooding, polymer flooding, etc. Surfactants with their complex chemical structures consisted of hydrophobic and
hydrophilic groups are used in various applications such as foaming, detergency, dispersion and solubilization of the
oils [1–5]. They are injected in oil reservoirs followed by a mixture of viscous polymer pumped into the porous medium
with an adequate pressure [6–8]. Surfactants are mainly injected to reduce the oil-water interfacial tension (IFT) to
values around 0.001mN/m, which consequently reduces the saturation of the residual oil. The oil trapped by the
capillary forces can be then displaced [9]. In litterature, it was reported that a quarter of the world’s reservoirs are
sandstone type, where the industrialists have tried to apply surfactant polymer formulation methods to recover more
oil from these reservoirs [10,11].
Currently, surfactant injection triggered an increased interest due to rising oil prices [12]. In the literature, many
influencing parameters on EOR injection were taken into account by researchers to quantify the retention quantities
of surfactants during its flow in the consolidated and unconsolidated reservoirs [13–17], as is the case for media
heterogeneities, surfactant types and reservoirs parameters effects. The adsorption could involve many mechanisms,
ae-mail: nadjibdrouiche@yahoo.fr (corresponding author)
Page 2 of 15 Eur. Phys. J. Plus (2019) 134: 436
such as electrostatic and non-polar chain-solid interactions, salvation and desolvation of adsorbates and adsorbents
materials, hydrogen bonding or van der Waals interactions between hydrocarbon chains of the adsorbed surfactant [18].
The surfactant retention on porous media can be related to many factors, such as the mineralogical composition of
rocks, the concentration, the dosage of the surfactant [19,20], the temperature [21, 22], the pH, the adsorption dose, the
ionic strength and the electrolyte concentration [4,23]. On the other hand, in order to optimize an adequate process
for EOR injection, adsorption equilibrium and kinetics are very practical laboratory tests for studying the surfactant
adsorption on a rock surface and to understand the interactions between them. Equilibrium results could facilitate the
determination of adsorption rate and predict the fate of surfactant in order to establish beneficial injection strategies.
Owing to the use of different surfactants with different compositions during the laboratory studies, a systematic
comparison among the adsorption of commercial surfactant is fairly impossible. The type of the employed surfactant,
the mineralogical and morphological properties of the rock and the types of the electrolytes existing in solution are the
most effecting parameters on the adsorption isotherm nature [20,24,25]. It was also found that the rock surface charge
and the fluid-solid interfaces have their own importance in the case of surfactant adsorption. According to studies done
by Wayman (1963) and Scamehorn (1980) about the adsorption of a dilute solution of sodium alkyl benzene sulfonate
on a variety of clay minerals, it was emerged that the Langmuir equation fits well with adsorption isotherms which
are capable of expanding with the usage of Truber diagram [24, 26–36].
Several recent studies have been conducted in order to investigate the retention of surfactants. Recently Rahul
Saha et al. [19] studied the effects of mineralogy and temperature on adsorption characteristics of SDS surfactant,
the results indicated that Langmuir isotherm and pseudo-second order were the models for describing the adsorption
phenomena which follows the order of illite >feldspar >montmorillonite >kaolinite. Ali Ahmadi analyzed the
influence of nanosilica particles on both adsorption and ultimate oil recovery when he used sodium dodecyl sulfate
as surfactant. The author showed that the incorporation of nanoparticle in both hydrophilic and hydrophobic states
could reduce the SDS loss due to the adsorption phenomena [37]. Scamehorn and Wayman studied the adsorption of
a dilute solution of sodium alkyl benzene sulfonate (SABS) as a surfactant on various clay minerals. They concluded
that the adsorption isotherms correspond to the Langmuir equation [24]. Ahmadi and Shadizadeh [18,19] reported
the adsorption kinetics behavior of the natural surfactant Zyziphus Spina Christi on real samples of shale-sandstone
reservoir, where the effect of temperature on the kinetics and equilibrium of adsorption process was experimentally
done. The results indicated that the Freundlich model was more acceptable for the adsorption equilibrium of this
natural surfactant after analyzing the related data for four models [11]. Bera et al. modeled the adsorption of three
different surfactants namely anionic (SDS), cationic (CTAB), and nonionic (Tergitol 15-S-7) on clean sand particles,
which were investigated with a variation of some parameters such as salinity, adsorbent dose, temperature, and pH.
Pseudo-second order and Langmuir were the most appropriate ones for predicting the adsorption process [4].
Other studies using optimization techniques have also been investigated. The statistical design of experiments in
which each factor was involved and simultaneously varied. It allows together aplenty information with a minimum
number of experiments [38–40]. Zhang et al. [41] studied the kinetics and equilibrium of acid dyes on chitosan/surfactant
adsorption. The main and interaction effects between different parameters, such as surfactant concentration, initial
dye concentration and temperature on adsorptions of two adsorbents were analyzed by 23full factorial designs. Cheng
et al. [42] reported four processes variables (i.e. surfactant concentration, temperature, and initial dye concentration)
of adsorption in which Congo red and direct red 80 AC/surfactant were optimized by response surface methodology
(RSM). Radnia et al. have evaluated the adsorption behavior of graphene oxide which represents an important factor
of chemical enhanced oil recovery methods onto sandstone surface at various levels of GO concentration. In this study,
the salinity and pH were assessed by surface response methodology [43]. Lebouachera et al. investigated experimentally
the effect of concentration, salinity and temperature of SDS adsorption onto an Algerian sandstone reservoir rock to
quantify the adsorbed quantity [44].
To our knowledge, no research in the literature has examined the adsorption of EOR ASP 5100 and SDS surfactants
in an Algerian sandstone reservoir using experimental design methodology. In this article, the adsorption behavior of
the new surfactant ASP EOR 5100 on the Algerian rock under the effect of different parameters, such as surfactant
concentration, contact time and temperature, were evaluated. The Adsorption isotherms models of Freundlich, Lang-
muir and Temkin have been analyzed. The kinetics were verified by pseudo first and pseudo second orders. Different
characterizations, such as XRD, SEM, XRF, BET analysis, were used to identify the mineralogy of this rock. The in-
troduction of the novel concept of full factorial design was used to optimize the adsorption quantity for an appropriate
surfactant selection in EOR processes.
2 Materials and methods
2.1 Surfactants
An anionic SDS surfactant with a purity of 98% and molecular weight equal to 288 g/mol purchased from Merck
Company alongside a commercial EOR ASP 5100 surfactant, purchased from Solvay company, were used during the
experiments.
Eur. Phys. J. Plus (2019) 134: 436 Page 3 of 15
2.2 Rock sample
Rock samples were obtained from Hassi Messaoud sandstone reservoir, which is located in south of Algeria. Crushed
rock samples were sieved into fractions of 4 μm to obtain a uniform geometric size for the experiment. The grains were
washed with distilled water and dried in an oven at 150 C for 24 h.
2.3 XRD, XRF, SEM and BET analysis
PANalytical X-ray diffractogram was used to evaluate the semi-quantitative composition of the Hassi Messaoud crushed
cores. The X-ray was applied in a wide range of Bragg angles (22θ70). The data were analyzed using a data
collector and treated with High score Plus PANalytical measuring instrument with CuKα(λ=1.54186 ˚
A). The
minerals composition results are evaluated based on the sequential spectrometer Bruker-Axs: S8 TIGER as shown in
table 1. Our results have been treated using the Spectra Plus software. In addition, we examined the sample by SEM,
looking for crystal morphology with Thermo Fisher Scientific under the name FEI-Quanta 650. The BET surface of
rock was measured using a Quanta chrome Autosorb-3bsurface analyzer based on nitrogen adsorption. The sandstone
powder was dried in an electric convection oven at 120 C overnight to eliminate water and further adsorbed substances.
2.4 Preparation of surfactant solutions
The surfactant solutions were prepared with distilled water and the concentration of SDS and EOR ASP 5100 was in
the range of 200–800 ppm.
The first solutions of SDS and EOR ASP 5100 were prepared with concentrations varying between 200 and 800 ppm
by dissolving the adequate amounts in 100ml distilled water. The surfactant was dissolved slowly in distilled water
under stirring using a magnetic stirrer to completely homogenize the solution.
2.5 Adsorption procedure
Batch adsorption tests were carried out in the laboratory by contacting a certain volume of surfactant solution
with Hassi Messaoud rock under stirring at 130 rpm for few hours to ensure equilibrium. Knowing the equilibrium
concentration and the initial concentration of surfactant, the amount of surfactant adsorption can be calculated using
Adsorbed surfactant =msolution (C0C)
msandstone ×103,(1)
where qis the surfactant adsorption on rock surface (mg/g of rock); msolution represents the total mass of solution in
original bulk solution (g); C0indicates the surfactant concentration in the initial solution before equilibrium (mg/L);
Cdenotes the surfactant concentration in the aqueous solution after equilibrium (mg/L); msandstone is the total mass
of the crushed sandstone. This method represents accurate results with an error of 4% for all experiments [45].
2.6 Evaluation of thermodynamic parameters
The fundamental thermodynamic parameters, such as enthalpy, entropy and Gibbs free energy can be calculated using
Langmuir isotherm constant, which depends on the system temperature [18, 44,46,47].
The Gibbs free energy change (ΔG) is a key parameter for determining the spontaneity of a process which can
be calculated according to the following equation:
ΔG=RT ln(KL),(2)
where Ris the universal gas constant (8.314Jmol/K), Tis the temperature (K) and KLis the Langmuir constant.
Other parameters related to ΔG, which are important in the feasibility and spontaneity of a process, are entropy
(ΔS) and enthalpy (ΔH) obtained according to
ΔG=ΔHS.(3)
By combining the equations, we get
ln(KL)=ΔS
RΔH
RT .(4)
By plotting ln(KL)versus 1/T , the values of ΔHand ΔScoincide with the slope and the intercept, respectively [48,
49].
Page 4 of 15 Eur. Phys. J. Plus (2019) 134: 436
Fig. 1. XRD model of the Hassi Messaoud rock.
Table 1. Constituents of the Hassi Messaoud rock.
Constituent (%)
Na2OMgO Al2O3SiO2P2O5SO3K2OCaO TiO2MnO Fe2O3PAF
0.53 0.24 9.13 84.64 0.04 0.03 1.44 0.11 0.08 0.02 1.25 2.44
2.7 Full factorial design experiments
In this work, a full 23factorial design requiring 8 experiments was used to study the importance of the individual
effects and interactions of SDS and ASP EOR 5100 surfactants on the Hassi Messaoud rock [50,51]. This design
was used to reduce the number of experiments and the time to reach the appropriate model for the optimization
process [52, 53], where 2krepresents the number of experiments, kthe number of factors [54]. Three input variables,
such as surfactant X1”, the concentration of each surfactant X2 and temperature X3 have been varied and one
response (experimental adsorption capacity in equilibrium (Qe)) was evaluated.
Note that the factors used in this work have never been investigated simultaneously using a full factorial design,
and were chosen for their important influence on the predicted response (Qe). The results of the factorial design
were studied and interpreted using the JMP8 software by analysis of variance (ANOVA), and the determination of
coefficient R2[50]. The significance of the regression coefficients parameters was discussed by the student’s test [55].
3 Results and discussion
3.1 Characterization of Hassi Messaoud rock
The Hassi Messaoud rock contains different quantities of quartz, kaolinite, illite, siderite and hallite, which are illus-
trated in fig. 1.
The composition of chemical compounds is indicated in table 1
The major elements of the constituent of the Hassi Messaoud rock were determined by XRF analysis, which
confirms the presence of Quartz with a composition of 84.64%. The specific BET surface area of the Hassi Messaoud
rock was measured under a nitrogen environment and was found to be equal to 5.71 m2/g.
The micro structural morphology of Hassi Messaoud rock was observed using SEM technique with secondary
electron option as shown in fig. 2.
The SEM image indicates that the surface is a mutual mixture of groove and pores, these spots could be an effective
adsorption site of SDS and EOR ASP 5100 surfactants.
Eur. Phys. J. Plus (2019) 134: 436 Page 5 of 15
Fig. 2. SEM image of the Hassi Messaoud rock.
Fig. 3. Concentration and contact time effect on EOR ASP 5100 adsorption kinetics (T=25C, speed = 130 rpm).
3.2 Effect of surf EOR ASP 5100 concentration and contact time
Figure 3 illustrates the effect of concentration and contact time for ASP 5100 surfactant on the Hassi Messaoud rock.
The adsorption amount increases with time and concentration. The kinetics of adsorption was rapid initially due
to the possible contacts of the surfactant with the available sites in the free surface. The adsorption reaches gradually
equilibrium due to the uptakes of ASP EOR 5100 molecules in the pores of adsorbents. The adsorbed quantities for
the surfactant were 1.75, 3.12, 5.52 and 5.85mg/g for 200, 400, 600 and 800 ppm respectively, where equilibrium was
reached after 45 minutes for all initial concentrations.
3.3 Effect of temperature on surf EOR ASP 5100 adsorption
The other important parameter that affects adsorption phenomena is the temperature, which is shown in fig. 4.
The quantity adsorbed for the surfactant at 80C was 1.56, 4.53, 7.00 and 8.98 mg/g for 200, 400, 600 and 800 ppm,
respectively, indicating a higher adsorption.
Page 6 of 15 Eur. Phys. J. Plus (2019) 134: 436
Fig. 4. Effect of temperature on surf EOR ASP 5100 adsorption kinetic (temperature 80 C, speed 130 rpm).
Table 2. Isotherms and their linear forms [41,56].
Isotherm Linear form Plot
Freundlich ln qe=lnKF+(1/n)lnCeln qevs. ln Ce
Langmuir Ce
qe=Ce
qmax +1
bqmax
Ce
qevs. Ce
Temkin qe=RT
BTln(KT)+ RT
BTln(Ce)qevs. ln(Ce)
Several previous works have proved that the increase in the temperature can cause the decrease in the quantity
adsorbed. However, in our experiments, the amount of adsorbed EOR ASP 5100 increased with a temperature (from
25 Cto80C) for all concentrations. The equilibrium is reached after only 12 minutes at 80C compared to 45 minutes
at 25 C. One can explain this result by temperature increasing the mobility of organic compound sand decreased the
energy activation that facilitated the contact between rock sample and the surfactant and cause a possible rapid
penetration into the Hassi Messaoud rock.
3.4 Equilibrium modeling
Once the equilibrium state is reached, the adsorption isotherms indicate the specific relationship between the equilib-
rium concentration of the bulk adsorbate and the amount adsorbed at the surface. The analysis of different data is an
important step to find the most suitable model for experimentation.
Several mathematical models can be used to describe the experimental data of adsorption isotherms. Three ad-
sorption isotherms, Langmuir, Freundlich and Temkin [19] were used to describe the equilibrium behavior of EOR
ASP 5100 on the Hassi Messaoud rock for the different concentrations and temperature. These isotherms are presented
with linear forms on table 2, where qe(mg/g) is the adsorption capacity at equilibrium; Ce(mg/l) is the concentration
of surfactant EOR ASP 5100 at equilibrium; KL(l/mg) is the Langmuir constant associated with the energy of ad-
sorption; qmax (mg/g) denotes the theoretical monolayer adsorption capacity; KFis the Freundlich constant (mg/g);
(l/mg)1/n while 1/n is a measure of the adsorption intensity or surface heterogeneity, BTis the Temkin constant re-
lated to the adsorption heat (kJ/mol); KTis the equilibrium binding constant (l/mol) corresponding to the maximum
binding energy; Ris the gas constant (8.314 J/mol K); Tis the absolute temperature (K).
The isotherm adsorption data was plotted in fig. 5 to obtain the Freundlich, Langmuir and Temkin models param-
eters that were summarized in table 3.
As can be seen from fig. 5, when comparing these results, the high R2values (>0.95) obtained using the Langmuir
linear isotherm indicate that this model describes very well the adsorption of EOR ASP 5100 at different temperatures
and concentrations. In the literature, the Langmuir isotherm model that was initially used to describe gas-solid
adsorption on activated carbon, is now usually preferred in the area of surfactant adsorption.
Eur. Phys. J. Plus (2019) 134: 436 Page 7 of 15
Fig. 5. Adsorption isotherms of EOR ASP 5100 on the Hassi Messaoud rock ((A) Langmuir, (B) Freundlich, (C) Temkin) at
different temperatures.
Table 3. Parameters of models at different temperatures.
Isotherm model Parameters Temperature
Freundlich KF(L/mg) T=25CT=80C
1/nf0.4566 0.0151
R20.3551 0.7389
0.912 0.868
KL(L/mg) 0.03300.0395
Langmuir qmax (mg/g) 2.1724 5.483
R20.985 0.995
KT(L/g) 0.0879 0.0136
Temkin BT0.086 0.401
R20.7667 0.819
The maximum adsorption capacity, qmax of the Langmuir model, was found to be 4.1754 mg/g and 7.9113 mg/g
for the temperature of 25 Cand80
C, respectively. The low KLvalues (0.0051 L/mg and 0.0092 L/mg) in both
temperatures could indicate that probably the rock of Hassi Messaoud has low surface energy, indicating a possible
stronger bond between EOR ASP 5100 and Hassi Messaoud rock. Also,the first adsorbed layer leaves the hydrophobe
tails towards the solution (i.e. the tail is away from the rock surface). In that case, hemimicelle or bilayer structure
are more likely [57].
Page 8 of 15 Eur. Phys. J. Plus (2019) 134: 436
Table 4. Results of pseudo-first order application for different concentrations and temperatures.
Pseudo-first order Correlation R2K1(min1)qe(mg/g)
Temperature 25 C
200 ppm ln(qeqt)=0.0037t+1.4242 0.7582 0.0055 2.0063
400 ppm ln(qeqt)=0.0038t+1.5705 0.8963 0.0111 6.0465
600 ppm ln(qeqt)=0.0042t+1.7165 0.8679 0.0108 8.0892
800 ppm ln(qeqt)=0.0053t+1.7389 0.9063 0.0097 10.7871
Temperature 80 C
200 ppm ln(qeqt)=0.2594t+17.1895 0.7371 0.035 3.0585
400 ppm ln(qeqt)=0.2351t+4.5505 0.9023 0.0121 4.2535
600 ppm ln(qeqt)=0.1557t+3.2325 0.7111 0.0075 8.1472
800 ppm ln(qeqt)=0.1162t+2.9237 0.859 0.0046 7.4973
Table 5. Results of pseudo-second order application for different concentrations and temperatures.
Pseudo-second order Correlation R2K2(g/mg min) qe(mg/g)
Temperature 25 C
200 ppm t/qt=0.4858t+ 195.9879 0.951 0.0011 2.7225
400 ppm t/qt=0.2055t+16.2955 0.9953 0.0026 4.8662
600 ppm t/qt=0.1670t+18,4355 0.992 0.0045 5.9880
800 ppm t/qt=0.1240t+21.0058 0.9860 0.0070 7.0926
Temperature 80 C
200 ppm t/qt=0.2594t+17.1895 0.951 0.035 3.895
400 ppm t/qt=0.2351t+4.5505 0.9757 0.0121 4.2535
600 ppm t/qt=0.1557t+3.2325 0.9897 0.0075 6.3171
800 ppm t/qt=0.1162t+2.9237 0.9840 0.0046 9.085
Also, Freundlich constants related to the adsorption potential, KF, and heterogeneity factor, 1/nf, were calculated
and found that KF(0.4566 and 0.0151) for 25 Cand80
C, respectively, were higher than other two models. The
low values of the heterogeneity factor (1/nf<1.0) given in the table revealed the heterogeneity of the surface of the
Hassi Messaoud rock. On the other hand, it is clear that Temkin model is the least adapted among the isotherms used
for these experiments because it provides a low R2(0.7767 and 0.819) for 25 Cand80
C, respectively. The Hassi
Messaoud field is known to be very heterogeneous and complex, so the adsorption plateau can vary under the same
conditions.
In addition, it can be concluded that the comparison of the models estimated is as follows: Langmuir >Freundlich >
Temkin.
3.5 Adsorption kinetics of ASP EOR 5100 on the Hassi Messaoud rock
The second important evaluation of data was the surfactant adsorption rate on the crushed sandstone which depends on
the interaction between adsorbate and adsorbent and the process conditions. The kinetics of ASP EOR 5100 adsorption
on the sorbent were analyzed by two models: pseudo-first- and second-order model. The forms and linearization of
these two models are cited in previous studies [19,58–60].
First, the parameters and R2values for the pseudo-first order model were calculated by plotting ln(qeqt)versus
taccording to slope and intercept of straight lines. The calculated values for the rate constant (K1) and equilibrium
adsorption rate (qe), were given in table 4. Furthermore, the values of K2and (qe) for the second-order model were
calculated based on the slope and intercept of a plot curve t/qtversus tand are summarized in table 5. The correlation
coefficients of the first model were low (R2<92%) and the intercept indicate that the straight lines do not pass through
the origin. This deviation from the origin probably occurs because of the mass transfer rate in the initial and final
stages of adsorption were different. On the other hand, the high R2values (>95%) suggest the pseudo-second order
model describes well the adsorption phenomena and had a good fitness with the experimental data for surfactant EOR
ASP 5100 adsorption on the Hassi Messaoud rock for each concentration and temperature.
Eur. Phys. J. Plus (2019) 134: 436 Page 9 of 15
Table 6. Thermodynamics properties of ASP EOR 5100 adsorption on the Hassi Messaoud rock.
T(K) KL(L/g) ΔG(kJ/mol) ΔH(kJ/mol) ΔS (kJ/mol)
298 330 18.70 17.50.00659
353 395 12.65
Table 7. Factors and levels used in the factorial design 23.
Factors Low level (1) High level (+1)
X1(surfactant) SDS EOR ASP 5100
X2(concentration in ppm) 200 800
X3(temperature in C) 25 80
Table 8. Experimental design matrix with the results of 23full factorial design.
Run X1X2X3Qe(mg/g) Qepredicted (mg/g)
11112.3294 2.3289125
2 1 112.7225 2.7229875
31 1 14.1762 4.1766875
4 1 1 17.0926 7.0921125
511 1 3.4734 3.4738875
6 1 1 1 3.895 3.8945125
71 1 1 6.144 6.1435125
8 1 1 1 9.085 9.0854875
3.6 Thermodynamic parameters of adsorption phenomena
Table 6 shows the values of ΔHand ΔSobtained by plotting the ln(KL)versus (1/T ) using eq. (4).
The Gibbs free energy is negative (18.70 kJ/mol)at25C which can demonstrate the spontaneity and the nature
of adsorption phenomena. Furthermore, it is remarkable that when temperature increases from 298 to 353 K, the
values of ΔGdecrease, which can be explained by the weaker adsorption due to the spontaneity and feasibility
of the process. In the literature, physisorption and chemisorption could be within the range of the following values
[20 kJ/mol; 0 kJ/mol] or [400 kJ/mol; 80 kJ/mol]. For the low temperature, adsorption is very high [4,44].
It is worth mentioning that the negative adsorption standard free energy changes (ΔG) and positive standard
entropy changes (ΔS) at all temperatures showed the spontaneous happening of the adsorption reactions. A negative
value of enthalpy reveals that this process is exothermic [61,62]; however, the very low value of entropy shows the negli-
gible increase in a randomness of a system. The same tendency was observed recently by Ahmadi et al. when he studied
the thermodynamic analysis of adsorption of a naturally derived surfactant onto shale sandstone reservoirs [19,63].
3.7 Factorial design of experiments
The choice of surfactant X1”, concentration X2 and temperature X3 are among the most important parameters
that affect the retention of surfactant on porous media, were used for the analysis of EOR ASP 5100 and SDS
adsorption on the Hassi Messaoud rock. The higher level is designated by +1 and the lower level by 1 as depicted in
table 7 for the parameters studied.
The response Ywas the experimental adsorption capacity equilibrium Qe(mg/g).
The run order of experiments carried out from right to left to correct the possible experimental errors [55] based
on the Yates’ algorithm, was applied in the case of all 2kfactorial experiments [64].
The design matrix of coded values for factors and response are shown in table 8 and the codified mathematical
model for the 23is given as
Y=a0+a1X1+a2X2+a3X3+a12X1X2+a13 X1X3+a23X2X3,
where Yis the experimental adsorption capacity at equilibrium Qe(mg/g) for SDS and EOR ASP 5100. While, the
a0,ai,aij represents the global mean. The regression coefficients assigned, respectively, to the principal factor effects
and interactions.
The student’s test t was used to confirm the validity of the regression coefficients of the studied parameters, when
t-value is higher than t-critic which is equal to 12.70, the regression is statistically significant [55]. After verification
of the results, R2was equal to 0.99 indicating the model to satisfy well the response [65].
Page 10 of 15 Eur. Phys. J. Plus (2019) 134: 436
Table 9. Estimated regression coefficients for adsorption capacity of SDS and EOR ASP 5100 on the Hassi Messaoud rock.
Parameter constant Estimate Standard error t-ratio p-value
4.8647625 0.000487 9979 <.0001
X10.8340125 0.000487 1710.8 0.0004(a)
X21.7596875 0.000487 3609.6 0.0002(a)
X30.7846875 0.000487 1609.4 0.0004(a)
X1X20.6303375 0.000487 1293 0.0005(a)
X1X30.0066375 0.000487 13.62 0.0467(a)
X2X30.2054625 0.000487 421.46 0.0015(a)
(a)p<0.05.
Fig. 6. The plot of predicted versus experimental adsorption capacity of ASP EOR 5100.
Table 10. ANOVA analysis of experimental adsorption capacity equilibrium Qe(mg/g).
Source Sum of squares DfMean square F-ratio p-value
Model 38.777910 6 6.46299 3399334 0.0004(a)
Residual 1.90125e-6 11.90e-6
C.Total 38.777912 7
(a)p<0.05.
From the data given in table 8, replacing the values coded by the estimates, the model can be written in the form
Qe(mg/g) = 4.8647625 + 0.8340125X1+1.7596875X2+0.7846875X3
+0.6303375X1X2+0.0066375X1X3+0.2054625X2X3.
All coefficients values have a positive sign in the polynomial model, indicating their positive influences on SDS and
EOR ASP 5100 adsorption.
Table 9 and fig. 6 show the predicted values versus the experimental values of the adsorption capacity in equilibrium,
the high value of R2=99.58% and R2
adjusted =99.28% indicate that the model was successful in correlating the response
to the studied parameters [65].
It appears from the table that the linear effects of surfactant, concentration, and temperature are significant. The
same trend was observed for the interactions effects between these factors, which confirms that the model is highly
significant with the p-value of 0.0004 <0.05 [50,66, 67] and F-value of 3399334 cited in ANOVA analysis in table 10.
By analyzing the data obtained and shown in table 10, it can be noted that the concentration (X2) is the most
important parameter for the overall adsorption process with a high (t-ratio = 3609.6), so the concentration exerts
a stronger influence on adsorption. After that, the surfactant type can be the second important parameter that
influenced this phenomenon with a (t-ratio = 1710.8). Temperature was the least significant parameter among the
main parameters with a (t-ratio=1609.4).
The interactions of (X1and X2) (with a high t-ratio=1293) was the most significant compared to the other
combinations (X2and X3) and (X1and X3). The latter was the least significant (with a t-ratio = 13.62).
Eur. Phys. J. Plus (2019) 134: 436 Page 11 of 15
Fig. 7. InteractioneectplotsoftheQefunction of the studied parameters.
Fig. 8. Prediction profiler of Qe(mg/g) function for SDS surfactant.
3.7.1 Interactions plots
Figure 7 illustrates the differences between the interactions of two parameters on the adsorption capacity of surf ASP
EOR 5100 and SDS surfactants. The non-parallel lines indicate the presence of interaction that can be estimated
between surfactant X1 and concentration X2 equal to 0.6303375, which means that the higher surfactant ASP
EOR 5100 affects the adsorption capacity when concentration is low and equal to 200 ppm.
In addition, the effects of interactions, such as surfactant X1 and temperature X3”, concentration X2”and
temperature X3”, are negligible due to parallel lines that indicate the temperature is the least significant or non-
significant in the presence of other parameters simultaneously. These results confirm the previous finding obtained
from table 9, related to each parameter of influence on the adsorption process.
3.7.2 Optimal design conditions using the desirability method
One of the important reasons for this work was to find the optimal conditions at which the adsorption capacity (Qe)
will be minimized. For this, the desirability function was used by Derringer and Suich in 1980 to solve the problems
related to the optimization of multiple responses related to industry [68,69], applied by Derringer and Suichin in many
studies [38,70].
From fig. 8, which shows the prediction profiler function of the studied parameter, it can be concluded that the
optimized conditions were the surfactant SDS concentration equal to 200 ppm and temperature 25 C for a predicted
response Qeequal to 2.3289 mg/g with a desirability value of 0.942118.
Page 12 of 15 Eur. Phys. J. Plus (2019) 134: 436
Fig. 9. Prediction profiler of Qe(mg/g) function for ASP EOR 5100 surfactant.
Fig. 10. Iso-response for Qe(mg/g) of the Hassi Messaoud rock on surfactant EOR 5100.
Likewise, the optimal conditions are obtained for EOR ASP 5100 (fig. 9) at a low concentration equal to 200 ppm
for 0.750683 of desirability, which gives the adsorption capacity equilibrium Qeequal to 3.894513 mg/g, corresponding
to a temperature of reservoir of 80 C. This result can be used in the industrial field.
The results obtained in figs. 8 and 9 show that the proposed mathematical model satisfactorily represents the
experimental results in the studied domain, in terms of desirability and adsorbed quantity illustrated in the previous
sections.
3.7.3 Iso-response of Qefunction
In order to evaluate the influence of concentration and temperature on the response Qe(mg/g) fixed at their optimum
(Qe=3.8945125 mg/g) with surf EOR 5100. The iso-response is shown in fig. 10.
In our case, the minimized adsorption is desired in the range of experimental study, which must be between 200 ppm
and 353 ppm for concentration and temperature scale between (25–80 C).
4 Conclusions
A detailed research on the adsorption of EOR ASP 5100 surfactant onto an Algerian rock reservoir was established.
The XRD and BET characterizations hepled in determing the mineralogy and the specific surface area of the Hassi
Messaoud rock. The adsorption parameters for the Langmuir, Freundlich and Temkin isotherms were determined, and
Eur. Phys. J. Plus (2019) 134: 436 Page 13 of 15
the adsorption kinetics of this new surfactant has been evaluated using two of the well-known models. The influence
of parameters, such as surfactant type, temperature, and surfactant concentration on adsorption capacity, has been
studied to improve oil recovery. These important parameters were evaluated using a statistical mathematical model
namely the 23full factorial design methodology. The following important conclusions can be drawn based on the
results obtained:
1) The main constituent element of the Hassi Messaoud rock (with a specific area equal to 5.71 m2/g) are quartz with
84.64%.
2) The kinetics and adsorption models were adequately modeled by Langmuir and pseudo-second order, respectively,
for all concentration and temperature variations, suggesting that the rate-limiting step may be the chemical ad-
sorption.
3) Higher surfactant concentrations and applied temperatures lead to a greater surface adsorption on the Algerian
rock until reaching a steady state which refers to the point of saturation.
4) From this study, a full factorial design of experiment can be an efficient method for the predicted val-
ues of surfactant adsorption which is reported for the newly statistical model with R2=0.99, F-ratio =
3399334 and p-value = 0.0004; Qe(mg/g) = 4.8647625 + 0.8340125 surfactant + 1.7596875 concentration +
0.7846875temperature + 0.6303375 (surfactantconcentration) + 0.0066375 (surfactant temperature)+0.2054625
(concentration temperature).
5) All factors and interactions considered in the experimental design were statistically significant at the 95% confidence
level.
6) The adsorption capacity values minimized for the SDS surfactant were Qe=2.3291 mg/g for SDS surfactant,
concentration = 200ppm and temperature 25 C. Likewise, minimized adsorption capacity values for EOR ASP
5100 was Qe=3.894513 mg/g for concentration 200 ppm and temperature of reservoir 80 C.
Nomenclature
EOR: Enhanced Oil Recovery X2: Second factor: concentration of
IFT: Interfacial Tension surfactant (ppm)
SDS: Sodium Dodecyl Sulfate X3: Third factor: Temperature (C)
P: Probability ANOVA: Analysis of variance
R2: Coefficient correlation a0: Intercept
R2
adjusted: Coefficient correlation adjusted ai: Regression coefficients attributed
t: student value to the principal factor effects
Y: Response aij : Regression coefficients attributed
C: Initial concentration of surfactant in solution to interactions effects
Ce: Equilibrium concentration of adsorbate RL: Separation factor
in solution DF: Degree of freedom
KT: Temkin isotherm constant 2: Number of levels
BT: Constant in Temkin adsorption isotherm k: Number of factors
KF: Freundlich isotherm constant q: adsorption of surfactant
K1: Rate constant of pseudo-first order kinetic model ASP: Alkaline Surfactant Polymer
K2: Rate constant of the second-order kinetic model t-ratio: The estimate to its standard error
n: Exponent in Freundlich isotherm F-ratio: The ratio of the mean square
Qe: Amount of the adsorption at equilibrium for the effect divided by the mean
qt: Amount of the adsorption at any time tsquare for error
RSM: Response Surface Methodology CTAB: CetylTrimethyl Ammonium Bromide
SEM: Scanning Electron Microscopy GO: Graphene oxide
XRD: X-Ray Diffraction ΔG: Change in Gibbs energy (kJ/mol)
XRF: X-Ray Fluorescence ΔH: change in enthalpy (kJ/mol)
BET: Brunauer, Emmett et Teller ΔS: change in entropy (kJ/mol)
PPM: Part per million R: gas constant
X1: First factor, surfactant T: Temperature
Page 14 of 15 Eur. Phys. J. Plus (2019) 134: 436
This work is part of a project supported by the petroleum company, SONATRACH - Algeria, that we gratefully acknowledge
for its contribution in providing additives and reservoirs rocks as well as the SOLVAY Company is gratefully acknowledged.
Our gratitude goes to Dr. Mohammed Abdelfetah Ghriga and Dr. Azzeddine Mazouzi for their help during the redaction of the
manuscript.
Publisher’s Note The EPJ Publishers remain neutral with regard to jurisdictional claims in published maps and institutional
affiliations.
References
1. E. Sabah, M. Turan, M. Celik, Water Res. 36, 3957 (2002).
2. S.H. Lin, H.G. Leu, Water Res. 33, 1735 (1999).
3. C. Czapla, H.J. Bart, Chem. Eng. Technol. 23, 1058 (2000).
4. A. Bera et al., Appl. Surf. Sci. 284, 87 (2013).
5. G. Cheraghian, Petrol. Sci. Technol. 33, 1580 (2015).
6. X. Xie et al.,SPEJ.10, 276 (2005).
7. A. Seethepalli, B. Adibhatla, K.K. Mohanty, SPE J. 9, 411 (2004).
8. S. Zendehboudi et al.,Can.J.Chem.Eng.91, 1439 (2013).
9. G.A. Pope, J. Petrol. Technol. 63, 65 (2011).
10. E. Worldwide, Oil Gas J. 108, 1 (2010).
11. M.A. Ahmadi, S.R. Shadizadeh, Fuel 159, 15 (2015).
12. S. Iglauer et al., J. Petrol. Sci. Eng. 71, 23 (2010).
13. R. Tabary et al.,Improved oil recovery with chemicals in fractured carbonate formations,inSPE International Symposium
on Oilfield Chemistry (Society of Petroleum Engineers, 2009).
14. K.J. Webb, C.J.J. Black, G. Tjetland. A laboratory study investigating methods for improving oil recovery in carbonates,in
International Petroleum Technology Conference, 2005 (International Petroleum Technology Conference, 2005).
15. G. Moritis, Oil Gas J. 102, 49 (2004).
16. T. Hassenkam et al., Colloids Surf. A 390, 179 (2011).
17. A.A. Al-Taq et al., SPE Reserv. Eval. Eng. 11, 882 (2008).
18. J. Sheng, Modern Chemical Enhanced Oil Recovery: Theory and Practice (Gulf Professional Publishing, 2010).
19. R. Saha, R.V. Uppaluri, P. Tiwari, Colloids Surf. A 531, 121 (2017).
20. R. Zhang, P. Somasundaran, Adv. Colloid Interface Sci. 123, 213 (2006).
21. B. Ball, D. Fuerstenau, Discuss. Faraday Soc. 52, 361 (1971).
22. S. Paria, K.C. Khilar, Adv. Colloid Interface Sci. 110, 75 (2004).
23. H. Shamsi Jazeyi, R. Verduzco, G.J. Hirasaki, Colloids Surf. A 453, 168 (2014).
24. J.F. Scamehorn, Equilibrium Adsorption of Surfactants on Mineral Oxide Surfaces from Aqueous Solutions (1981).
25. C.H. Wayman, Surfactant sorption on heteroionic clay minerals,inInternational Clay Conference Stockholm (1963).
26. K. Hu, A.J. Bard, Langmuir 13, 5418 (1997).
27. P. Chandar, P. Somasundaran, N.J. Turro, J. Colloid Interface Sci. 117, 31 (1987).
28. M.R. Bohmer, L.K. Koopal, Langmuir 8, 2649 (1992).
29. R.F. Tabor, J. Eastoe, P. Dowding, Langmuir 25, 9785 (2009).
30. M. Ishiguro, L.K. Koopal, Adv. Colloid Interface Sci. 231, 59 (2016).
31. B.-Y. Zhu, T. Gu, Adv. Colloid Interface Sci. 37, 1 (1991).
32. P. Somasundaran, S. Krishnakumar, Colloids Surf. A 123, 491 (1997).
33. D. Levitt, M. Bourrel, Adsorption of EOR chemicals under laboratory and reservoir conditions, part III: Chemical treatment
methods,inSPE Improved Oil Recovery Conference (Society of Petroleum Engineers, 2016).
34. K. Ma et al., J. Colloid Interface Sci. 408, 164 (2013).
35. K.M.K. Yu et al.,J.Mater.Chem.13, 130 (2003).
36. W. Zhong, J.P. Claverie, Carbon 51, 72 (2013).
37. M.A. Ahmadi, Eur. Phys. J. Plus 131, 435 (2016).
38. A. Regti et al.,Microchem.J.130, 129 (2017).
39. D. Rocak, M. Kosec, A. Degen, J. Eur. Ceram. Soc. 22, 391 (2002).
40. E. S¸ayan, M. Bayramoˇglu, Hydrometallurgy 57, 181 (2000).
41. L. Zhang et al., J. Mol. Liq. 197, 353 (2014).
42. Z. Cheng et al., Spectrochim. Acta Part A 137, 1126 (2015).
43. H. Radnia, A.R.S. Nazar, A. Rashidi, J. Taiwan Inst. Chem. Eng. 80, 34 (2017).
44. S.E.I. Lebouachera et al.,Res.Chem.Intermediates44, 7665 (2018).
45. A. Barati et al., Chem. Eng. Res. Design 105, 55 (2016).
46. D.M. Ruthven, Principles of Adsorption and Adsorption Processes (John Wiley & Sons, 1984).
47. AL. Weiss, Ber. Bunsengesell. Phys. Chem. 94, 796 (1990).
48. M. Wi´sniewska, Appl. Surf. Sci. 258, 3094 (2012).
Eur. Phys. J. Plus (2019) 134: 436 Page 15 of 15
49. L.-C. Juang, C.-C. Wang, C.-K. Lee, Chemosphere 64, 1920 (2006).
50. D. Hank et al., J. Ind. Eng. Chem. 20, 2256 (2014).
51. S. Abdul-Wahab, J. Abdo, Appl. Therm. Eng. 27, 413 (2007).
52. F. Geyik¸ci, Prog. Org. Coat. 98, 28 (2016).
53. L.S. De Lima et al.,Chem.Eng.J.166, 881 (2011).
54. F. Geyikci, H. uy¨ukg¨ung¨or, Acta Geodyn. Geomater. 10, 363 (2013).
55. J. Goupy, L. Creighton, Introduction to Design of Experiments with JMP Examples (SAS Publishing, 2007).
56. P. Sampranpiboon, P. Charnkeitkong, X. Feng, WSEAS Trans. Environ. Develop. 10, 35 (2014).
57. M. Tagavifar et al., Colloids Surf. A 538, 549 (2018).
58. M. Arabloo, M.H. Ghazanfari, D. Rashtchian, J. Taiwan Inst. Chem. Eng. 50, 12 (2015).
59. S. Lagergren, K. Sven. Vetenskapsakad. Handl. 24, 1 (1898).
60. Y.-S. Ho, Water Res. 40, 119 (2006).
61. C. Qi et al., Res. Chem. Intermed. 44, 2889 (2018).
62. A. Mehrizad, Res. Chem. Intermed. 43, 4295 (2017).
63. M.A. Ahmadi, S.R. Shadizadeh, Z. Chen, Eur. Phys. J. Plus 133, 420 (2018).
64. H. Spliid, Desing and Analysis af Experiments with kFactors Having pLevels (2002).
65. N.A. Jarrah, Chem. Eng. J. 151, 367 (2009).
66. D.N. Gujarati, Econom´etrie, Traduction de la 4i`eme ´edition am´ericaine par Bernard Bernier (De Boeck, Bruxelles, 2004).
67. M. Ghaedi et al., Res. Chem. Intermed. 44, 2929 (2018).
68. G. Derringer, J. Qual. Technol. 12, 214 (1980).
69. L.V. Candioti et al., Talanta 124, 123 (2014).
70. L. Hamdi et al., Desalin. Water Treat. 57, 6098 (2016).
... It was found that the initial dye concentration has a significant influence on the amount of solute molecules that can be pushed to overcome the solid-to-liquid mass transfer barrier through adsorption [95]. Thus, increasing the surfactant concentration results in an increase in surface adsorption [102]. ...
... Similarly, Bera et al. [74] tested PFO, PSO and intraparticle diffusion kinetics. They revealed that q e,cal significantly close to q e,exp estimated from the PSO equation, with a significant R 2 and a low MSE for the 3 models, suggesting that the data on the adsorption kinetics of surfactants on surfaces are represented using a second-order kinetic model [102]. In accordance to the kinetics discussed above, a finding was reached in another research by Saha et al. [81], who used the Elovic model to illustrate the kinetics of SDS adsorption onto the Assam oil field (EM). ...
... Recent studies underscore DoE's value [84] in surfactant science, particularly for cEOR, by rigorously evaluating novel surfactant compositions and their interplay with reservoir conditions. This focus on DoE not only mirrors the ongoing innovation in the field but also heralds the development of cost-effective and environmentally sustainable solutions for surfactant adsorption challenges [84,102,168,170,174,175]. In summary, optimizing surfactant adsorption transcends mere technical adjustments, demanding a strategic approach that embraces recent scientific advancements and innovative methodologies like DoE. ...
Article
This comprehensive review delves into the recent advancements in the study of surfactant adsorption, a crucial factor influencing the efficiency of chemical enhanced oil recovery (cEOR) applications across various oil types. By integrating findings from the latest literature, we shed light on the dynamic interactions between surfactants and diverse adsorbents including sandstone, carbonate, and shale sandstone. This review highlights the nuanced understanding of how operational variables such as initial surfactant concentration, pH, adsorbent dose, contact time, temperature, along with adsorption kinetics, isotherm models, thermodynamics, and the presence of competing ions, contribute to the adsorption process. A novel synthesis of recent studies reveals that the Langmuir and pseudo-second-order models continue to accurately describe the equilibrium and kinetics of adsorption in most scenarios, with the adsorption process predominantly exothermic. Additionally, this review introduces cutting-edge insights into the molecular-level mechanisms underpinning surfactant adsorption, emphasizing the role of crude oil components, initial wettability of reservoir rocks, and the intrinsic properties of the rock itself. By summarizing these contemporary findings, the review aims to provide a deeper understanding of the key parameters affecting the retention of surface-active agents on various adsorbents, thereby proposing new avenues for optimizing surfactant flooding in cEOR. This enhanced focus on recent contributions to the field distinguishes our review from existing literature, offering fresh perspectives on the optimization of surfactant use in oil recovery processes.
... As a result, the experimental design dramatically impacts the size of the generated surface response [12]. In recent years, RSM plays a significant part in a variety of research domains and industry, which has been applied in many optimization studies [13][14][15][16][17][18][19][20][21][22]. It offers numerous advantages, including the ability to investigate the interaction between independent variables and the answer while simultaneously reducing time and money via the use of fewer trials. ...
... The final response surface model was further lowered by removing statistically unimportant features; the fitted second-order model for the standard deviation of nanofiber diameter is given in Eqs. (12) and (13) before and after the insignificant components were removed [78]: GO, SC, and FR encode the GO content, solution concentration, and feed rate, respectively. According to ANOVA results, GO concentration does not affect the standard deviation of nanofibers. ...
Article
Full-text available
Composites, blends, and polymer membranes are all examples of processes that may benefit from response surface methodology (RSM) to optimize material characteristics. RSM is a convenient technique for determining the effect of several variables and their interactions on single or multiple response variables. The theoretical foundations and practical applications of RSM including the advantages and disadvantages of some design models are thoroughly explained. The RSM technique is shown to be relevant for the first time in polymer science and engineering research. We'll divide it into three sections: RSM may be used in some scenarios for polymerization, manufacturing, and evaluating the characteristics of polymers. The paper also compares RSM and other optimization techniques to an artificial neutral network. This review should be beneficial to anyone involved in polymer RSM and optimization. Graphical abstract
... To determine the minimum adsorbed quantity, a full factorial design consisting of 8 experiments derived from 2 3 factors was employed. This design allowed for the examination of individual effects and interactions of operational parameters using methods such as variance analysis (ANOVA), the desirability method and response surface methodology [20]. ...
Article
Full-text available
This research article concentrates on process conditions in addition to improving color selections in polymer compounders and developing more accurate simulation models. The feed rate (FR), temperature (T) and screw speed (SS) are three processing variables that the research investigates using general trends (GTs) and Box-Behnken design (BBD) response surface methodology. The identical set of processing settings was tweaked at three separate phases independently of one another. This study uses the experimental design to investigate process parameters' optimization while holding all other parameters constant. This design was given the name GT. To develop this design and its statistical optimization, this study used the software of the design expert method. A regression model was run in this design, which displayed collective as well as individual effects of the parameters on color images. The values of tri-stimulus color with the best optimization had the smallest proper color variance (dE*). To obtain information on pigment characteristics, an SEM image analysis was conducted, which aids in improving future designs and overcoming manufacturing issues that affect color fluctuation properties and waste reduction for various chemical grades, both of which enhance environmentally friendly processes.
... The specified range is primarily determined by the economic components of the project [75]. At higher values, the costs of injecting the sacrificial agent cannot be offset by additional oil production, as shown in the following works [76,77]. Therefore, studies related to the mechanisms and methods of reducing the adsorption of surfactants on the rock are of particular relevance. ...
Article
Full-text available
Despite the development of alternative energy sources, oil and gas still remain the predominant energy sources in most countries in the world. Due to the gradual hydrocarbon reserve depletion and the existing downward trend in the production level, there is a need to search for methods and technical approaches to level off the falling rates. Chemical enhanced oil recovery methods (EOR) by surfactant solution injections are one of the possible approaches for addressing this issue in already developed fields. Most often, surfactants are injected together with polymers or alkalis. These technologies are called surfactant-polymer (SP) and alkali-surfactant-polymer (ASP) flooding. Basically, SP and ASP have been distributed in China and Canada. In this article, in addition to these countries, we paid attention to the results of pilot and full-scale tests of SP and ASP in Russia, Hungary, and Oman. This study is a comprehensive overview of laboratory and field tests of surfactant solutions used for oil displacement in SP and ASP technologies. The first part of the article discusses the physical fundamentals of the interaction of oil with surfactants. The second part presents the main chemi-cal reagents used to increase oil recovery. In the third part, we describe the main facilities used for the preparation and injection of surfactants. Further, the results of field tests of SP and ASP in the above mentioned countries were considered. In the discussion part, based on the considered results, the main issues and uncertainties were identified, on which some recommendations were proposed for improving the process of preparation and injection of surfactants to increase oil re-covery. In particular, we have identified the area of additional laboratory and scientific-practical research. The outcomes of this work will provide a clearer picture of SP and ASP, as well as information about their limitations, current challenges, and potential paths forward for the development of these technologies from an economic and technological point of view.
... The first one depends on a single factor, changing one condition and fixing the others, while the second relies on the statistical method [19]. Several researchers used a design of experiments approach (DOE) in different applications [20][21][22][23], including a response surface methodology (RSM), when many parameters influence response or more responses of interest. ...
Article
Full-text available
In the present work, the dissimilar joints between C45 carbon steel and nickel-chromium 16NiCr6 steel rods were produced using rotary friction welding process. Statistical analysis based on response surface methodology (RSM), microstructural examination using scanning electron microscopy with backscattered electron diffraction (EBSD) and mechanical tests were performed to investigate the friction weld joints. The results showed that friction time and rotation speed were the most effective parameters on the weld joint quality with the highest t-ration of − 4.27, where the maximum bending strength of 1406.9 MPa was obtained at 2000 rpm for 13 s friction time. Increasing friction time to 13 s resulted in remarkable decrease in grain size (about 35%) at the weld interface, which increased the hardness (350HV0.1) and elastic modulus (260 GPa). Graphical abstract
... One of the main goals of this work was to find the optimal conditions in which the zero-shear viscosity (η 0 ) will be maximized. For this, the desirability function was used by Suich and Derringer in 1980 to solve the problems related to optimizing multiple responses related to the industry that has been applied in many studies [42] [43,44]. Figure 9 shows the prediction profile of the zero shear viscosity η 0 as a function of the studied parameters. ...
Article
In this work, the reinforcing effects of polystyrene microspheres in a hydrolyzed polyacrylamide (HPAM) matrix was studied. The experimental design approach was used to optimize the rheological behavior of the HPAM polymer. The effects of three important parameters, including the size of the polystyrene microspheres, their concentration, and the test temperature on the zero-shear viscosity of HPAM polymer solutions were considered. The results showed that the size of the microspheres and test temperature were the most influential factors simultaneously. Moreover, the optimal zero-shear viscosity, at 20 °C, was predicted to be at microspheres concentration of 50 ppm and microspheres size of 1000 nm. The results showed close agreement between the predicted results by the RSM method and those obtained experimentally.
Article
Full-text available
Hypersaline reservoirs are characterized by high salinity and high calcium and magnesium concentration. In order to enhance oil recovery of the hypersaline reservoirs, a specialized ternary mixed surfactant system composed of nonionic alkanolamide surfactants and anionic surfactant was developed in this study. Through careful analysis and optimization, lauric acid diethanolamide (LDEA), octanoic acid diethanolamide (ODEA), and sodium dodecyl sulfonate (SDS) were identified as promising candidates for the surfactant compounding system, and formed a ternary surfactant system composed of LDEA, ODEA, and SDS with the mass ratio of 4.64 : 0.66 : 1.00. Experimental results revealed that the interfacial tension of the system was consistently below 10⁻² mN m⁻¹ and could even reach ultra-low levels (10⁻³ mN m⁻¹) under conditions of calcium and magnesium ion content of 2000 mg L⁻¹, surfactant concentrations ranging from 0.05 to 0.3 wt%, temperature ranging from 50 to 80 °C, and salinity ranging from 20 000 to 50 000 mg L⁻¹. Furthermore, the mixed surfactant system exhibited favorable wetting capacity and emulsifying power. The static adsorption capacities of the mixed surfactant on oil sands were less than 2 mg g⁻¹. This study offered a novel strategy for the actual exploitation of reservoirs with high calcium-magnesium and high salinity.
Article
In this work, significant interest has been focused on removing pharmaceutical pollutants, i.e., oxytetracycline (OTC), which was studied through adsorption onto activated carbon (ACT) produced from households or agricultural wastes under various experimental conditions. The physicochemical properties of the produced carbon were investigated using multiple techniques. The characterization analysis highlighted an essential concentration of surface functional groups and a very developed porous structure. The adsorption process onto activated carbon occurs with a high yield at a large range of pH values varying between 4 and 8 at an optimum contact time of 2 h. It was established that the pseudo-second-order kinetic model fitted well OTC adsorption onto ACT. Moreover, the adsorption isotherm data showed that the pollutant removal process followed the Langmuir model with high regression coefficients (R 2 ) values and important adsorption efficiency of 80% for 100mg·L -1 of OTC. The displayed results consequently, the use of ACT as an adsorbent could be considered an efficient, sustainable, and low-cost alternative for pharmaceuticals-loaded wastewater treatment.
Article
Full-text available
This research focuses on the standardization of the blood meal production process at a Colombian rendering plant through a design of experiments. Initially, 108 samples of blood meal were taken where only 23% achieved the moisture target (7.5% to 8.5%). Therefore, an analysis of the measurement system was performed using a repeatability and reproducibility (R&R) study. Results showed that 39.96% of the observed variability was caused by the measurement system that was out of control. So, it was necessary to improve the method of sampling reducing the participation of the measurement system in the variability of the process to only 3.79%. Later, several experiments were accomplished with a 2k factorial design. Each experiment consisted of a response variable (blood meal moisture), two controllable factors (drying chamber temperature and percentage of rotation of the metering screw), and an uncontrollable factor (initial blood meal moisture). Finally, experiments were carried out and validated observing that, with a drying chamber temperature of 160 °C and a percentage of screw rotation of 29%, more than 97% of the blood meal was according to the moisture target. In conclusion, is confirmed that the design of experiments is a tool that allows a clear path towards optimization and standardization of processes.
Thesis
Les réservoirs fracturés naturellement occupent la première place dans les réserves mondiales et forment une classe très particulière des réservoirs d'hydrocarbures. Leur étude est très spécifique et complexe à cause de l’existence de la porosité et de la perméabilité secondaires caractéristiques des réseaux de fractures. Les fluides s'écoulent essentiellement à travers un réseau de fractures tandis que la grande quantité de pétrole est emmagasinée par la matrice rocheuse, ce qui rend la production de pétrole à partir de ces réservoirs difficile et un grand défi pour les ingénieurs pétroliers. Lors de la récupération secondaire ou tertiaire dans les réservoirs fortement fracturés, l'injection de fluides de faible viscosité, tels que le gaz ou l'eau, conduit généralement à des percées précoces de fluides (canalisation de fluide ou channeling) dans les puits de production, entraînant ainsi une mauvaise efficacité, à l’échelle microscopique et macroscopique, de la récupération de pétrole. Le challenge pour les opérations de la récupération chimique améliorée du pétrole (cEOR) dans les réservoirs fracturés, c’est d'empêcher ce phénomène de canalisation des fluides « channeling », afin que les fluides injectés puissent entrer, se mettre en contact et déplacer l'huile piégée dans la matrice. Ces opérations sont connues sous le nom de « contrôle de conformance ». Dans un réservoir fracturé, la conformance peut être améliorée en réduisant la perméabilité secondaire des fractures par l’utilisation des systèmes de gels, de la mousse ou de la croissance microbienne. L’objectif de cette thèse est d'étudier la gélification d'un système de gel polymère formé par les polyacrylamides partiellement hydrolysés (PAPHs) et les polyethylenimines (PEIs). Pour des réactifs bien caractérisés, le temps de gélification, la force du gel et la stabilité thermique du gel sont étudiés en fonction des principaux paramètres physico-chimiques du système (PAPH/PEI). La méthode des surfaces de réponses (MSR) est utilisée pour développer un modèle mathématique qui permet la prédiction du temps de gélification des systèmes (PAPH/PEI) dans l’intervalle de températures compris entre 70 et 90 °C. Les mécanismes de réticulation, du PHPA avec de la PEI, sont ensuite étudiés et quantifiés à l'aide des techniques avancées.
Article
Full-text available
The adsorption of surfactant on reservoir rocks or sands is one of the main factors that can significantly reduce the effectiveness of surfactant flooding for light oil recovery. The present work consists of study of the interaction between the surfactant sodium dodecyl sulfate (SDS) and the characteristics of Hassi Messaoud rock established using scanning electron microscopy, Fourier transform infrared analysis, Brunauer–Emmett–Teller surface area analysis and optical microscopy. Parametric study of the adsorption tests was carried out at a temperature between 25 and 80 °C. The conductivity method was used to measure the critical micellar concentration of the SDS surfactant in aqueous solutions. The adsorption data were adjusted using five isotherms and two kinetics of the anionic surfactant. The equilibrium conditions for crushed rock samples were obtained after 11 h of adsorption. According to the experimental results, the equilibrium data were well fitted by the Langmuir isotherm for sandstone sample at all temperatures. The pseudo-second-order model best describes the kinetics of the adsorption process. Furthermore, the adsorption process is exothermic. Several statistical error analyses were applied to evaluate the performance and accuracy of the proposed experiment. A systematic investigation is very useful for selecting an appropriate surfactant for enhanced oil recovery application and reservoir stimulation in the petroleum industry.
Article
Full-text available
The inclusion complex (sulfated-β-CD@NAA) of sulfated β-cyclodextrin and α-naphthylacetic acid was synthesized for adsorption of low concentrations of thorium from aqueous solutions. The structure and property of sulfated-β-CD@NAA were characterized by Fourier transform infrared spectroscopy, thermogravimetric, X-ray diffraction, proton nuclear magnetic resonance, specific surface area measurement and scanning electron microscopy. The characterization analysis confirmed the formation of new solid phase. The results indicated that α-naphthylacetic acid embedded into the cavity of sulfated β-cyclodextrin partly from the secondary face and the thermal stability of inclusion complex was significantly increased. The experiments were designed by Box–Behnken design combined with response surface methodology. The adsorption parameters of pH value, contact time, initial concentration were used as the independent variables and their effects were investigated on the Th(IV) adsorption capacity. Second-order polynomial regression models were derived and analysis of variance was utilized to judge the adequacy of the chosen models. The maximum Th(IV) adsorption capacity (12.75 mg g−1) was achieved under the following conditions: pH value 2.5, contact time 35 min, initial concentration 30 mg L−1. Prediction of models was in good agreement with experimental results. The experimental adsorption kinetic data followed a pseudo-second-order equation with the correlation of 0.9996, indicating the chemical adsorption. The obtained equilibrium isotherm data were fitted well with Langmuir model in the concentration range considered. The thermodynamic parameters (∆H0 < 0, ∆S0 < 0, ∆G0 < 0) indicated that the adsorption process was exothermic and spontaneous. Regeneration of Th(IV)-loaded sulfated-β-CD@NAA with 0.1 M HNO3 solution, and the regenerated sulfated-β-CD@NAA can be reused for adsorption of Th(IV).
Article
Full-text available
This study mainly deals with the adsorption behavior of the graphene oxide (GO) onto sandstone which is an important factor for applying a material in chemical enhanced oil recovery methods. The combined effects of initial GO concentration, salinity and pH of the solution are assessed by adopting response surface methodology. The results show that the GO concentration has a stronger influence on the GO adsorption than those of the initial pH and salinity. The effect of pH on the GO adsorption becomes significant at high GO concentrations and low salinities. The Derjaguin–Landau–Verwey–Overbeek (DLVO) theory is applied to explain the observed trend of GO adsorption under various salinity and pH conditions. Reduction in the height of energy barrier and formation of a secondary minimum are responsible for increasing the GO adsorption at lower pH values and moderate salinities (1 wt.%). The contact angle measurement of the rock surface treated in GO solution at optimum conditions (GO concentration of 0.89 mg/mL, salinity of 5 wt.% and pH of 6.74) shows that the adsorbed GO can alter the wettability of sandstone from strongly oil wet (150°) to intermediate conditions (90°).
Article
Full-text available
A novel nanobiocomposite in the role of adsorbent was prepared by mixing silver (Ag) nanoparticles with cuttlebone (CB) and was characterized by UV–Vis, DLS, XRF, EDX, FE-SEM and BET analysis. Modeling and optimization of simultaneous removal of 4-nitrophenol and 4-chloro-2-nitrophenol in binary solution by adsorption onto prepared adsorbent (Ag–CB) was studied using response surface methodology (RSM). The adsorption of phenols onto Ag–CB reached equilibrium within 25 min. The removal efficiency increased with increment of Ag–CB dosage (1–5 g L−1) and contact time (5–25 min), whereas it decreased with increasing the solution pH (3–11) and the initial concentration of phenols (5–25 mg L−1). The removal efficiency of phenols under optimum conditions (initial phenols concentration of 10 mg L−1, Ag–CB dosage of 4 g L−1, pH of 5 and contact time of 20 min) was 86.41%. Polymath software was used to draw nonlinear isotherm and kinetics curves. The nonlinear regression methods revealed that the adsorption data can be well interpreted by Freundlich isotherm model (KF = 0.30 [(mg g−1)(L mg−1)]1/0.66; n = 0.66) and Ho’s pseudo-second order kinetics equation (k2 = 0.17 g mg−1 min−1). Thermodynamic parameters declared that adsorption process was exothermic and nonspontaneous in the temperature range of 25–45 °C.
Article
Increasing oil recovery from depleted sandstone reservoirs with a low degree of pressure can become possible through performing different chemical methods, particularly polymer-surfactant flooding. Nevertheless, high prices of chemical additives and losing surfactants because of their adsorption on reservoir rocks have always been matters of consideration in this modern EOR solution. In this paper, the adsorption behavior in a system including two phases of gathered porous samples, mainly composed of shale-sandstone and the aqueous solution of an extracted surfactant from Zizyphus spina-christi, has been studied to examine the adsorption thermodynamic and isotherm models of the proposed surfactant. Effects of different surfactant concentration in different temperature on surfactant adsorption density have been analyzed in a batch system. Based on the established conductivity method, the adsorption of the introduced surfactant in a variety of conditions was measured and assessed. It was clear that a higher temperature results in a lower adsorption of the addressed derived natural surfactant from Zyziphus spina-christi on the porous samples. Also, an equilibrium time of ten days was roughly recorded for the crushed rocks. Both Freundlich and Langmuir isotherms performed a high level of compatibility with data belonging to sub-micelle concentration. All in all, the outputs of the performed investigation can be taken as guidelines to select the most appropriate surfactant efficiently for enhanced oil recovery (EOR).
Article
We investigate surfactant adsorption on a model carbonate rock over a wide range of pH and surfactant-to-solid ratios, by both an experimental and a theoretical approach, to obtain a quantitative understanding of how mineral constituents affect the adsorption equilibrium and dynamics. A combination of bulk mineralogy, elemental composition analysis, dissolution behavior, and water ionic composition data were used to characterize the samples’ surface heterogeneity. The adsorption of anionic surfactants is 2.4 mg/m² at pH of ∼8 and decreases approximately linearly with pH values above 9. To constrain and compare the relative adsorption affinity and the likely modes of attachment on mineral constituents as pH changes, we performed surface complexation calculations using a two-surface multisite diffuse layer model. We propose the formation of two surface species on both the major (i.e., calcite) and trace (i.e., the oxide-like sites on the edges of clay platelets) minerals: a monodentate inner-sphere complex and a weak or hydrogen bonding complex. Our modeling results suggest that charge-regulated inner-sphere complexation is the dominant adsorption mechanism on the calcite and oxide-like sites at low pH values regardless of the surface loading. We found weak or hydrogen bond adsorption to be significant on the calcite surface, and this became the dominant adsorption mode at pH ∼10. While the adsorption on calcite increases with surface loading, adsorption on the oxide-like sites remains independent of surface loading. These results suggest that surfactant adsorption can be comparable on the abundant low-surface-area calcite and trace high-surface-area oxide-like sites.
Conference Paper
This is the final installment in a series of three papers examining iron mineralogy and its effect on surfactant adsorption in reservoir and outcrop rock samples. The goal of these studies is to establish best practices for obtaining surfactant adsorption values representative of those in a reduced oil reservoir, despite performing experiments in an oxidizing laboratory atmosphere. This article follows two others examining the abundance and form of iron in the reservoir and in core samples (Part I: Levitt et al., 2015), and a proposed core restoration technique utilizing iron-reducing bacteria (Part II: Harris et al., 2015). In this Part III, chemical reduction methods are examined. Surfactant retention is a leading uncertainty in economic forecasting of chemical EOR, in large part due to the order-of-magnitude effects of artifacts such as improper core preservation. The industry standard is to (a) limit atmospheric contact of cores to the extent feasible, and (b) when necessary, reduce oxidized cores using strong reducing agents such as sodium dithionite, along with buffering and chelating agents such as sodium bicarbonate and EDTA or sodium citrate. However few studies have been performed to determine whether such invasive treatments are necessary, or what unintended effects the use of such reactive chemicals may have. The most striking conclusion from these studies is the lack of clear evidence of any advantage of electrochemical reduction versus a simpler treatment with chelators such as sodium citrate or EDTA. Wang (1993) suggests that oxidation of reservoir cores leads to higher surfactant adsorption due to the reduction of clays, which yields a more negative surface charge. Static experiments with montmorillonite clay, as well as an oxidized outcrop containing significant clay and iron content, illustrate that rinsing with non-reducing agents such as sodium bicarbonate, EDTA, or sodium citrate can lower adsorption as much as a strong reducing agent such as sodium dithionite. In the case of montmorillonite, cation exchange appears to be the mechanism by which adsorption is lowered, and so NaCl alone is sufficient to lower adsorption to near-zero values. For the iron- and clay-containing outcrop material, initial measurements indicating "adsorption" far in excess of a dense bilayer were due in fact to the precipitation of sulfonate surfactant with calcium, which eluded from the dissolution of small amounts of anhydrite. An alkyl alkoxy sulfonate surfactant showed higher calcium tolerance, and did not yield "multilayer" adsorption when equilibrated with the anhydrite-containing core sample. While treatment with a citrate-bicarbonate-dithionite solution does indeed lower adsorption several-fold further, solutions of either sodium bicarbonate or EDTA are at least as effective, and sodium citrate is almost as effective. These non-reductive treatments remove small amounts (~0.1% – ~0.2% of rock mass) of Fe and Al, and fines are invariably apparent in treatment fluids, both of which suggest removal of small amounts of trivalent Fe/Al colloids. Wang (1993) suggests reduction or removal of trivalent iron from clay surfaces as a possible mechanism of lowered adsorption under electrochemically reducing conditions. These results suggest that removal of trivalent cations, with concomitant lowering of anionic surfactant adsorption, is possible with non-reductive chelators such as sodium citrate or sodium EDTA. Sodium bicarbonate is equally effective at lowering adsorption, but does not result in elution of Fe or Al, indicating that these are likely reprecipitated. PIPES buffer, which is used in biological applications for its low propensity to form ligands, lowers adsorption as much and no more than a 10% NaCl rinse, suggesting only anhydrite removal and possibly cation exchange with clays occurs. While these results suggest that non-reductive means may be used to remove artifacts introduced by core oxidation, they come with an important caveat: even rinsing with a brine solution can result in significant alteration of mineralogy. The use of chelating agents will invariably result in dissolution of any soluble minerals present such as gypsum or anhydrite, which can be an important contributor to surfactant (in particular ABS) consumption. In cases where iron removal is necessary due to polymer degradation issues, PIPES buffer is proposed for use as an alternative to bicarbonate, the latter having a greater tendency for ligand formation. The combination of borohydride and bisulfite is suggested as an alternative to dithionite as a reducing agent, resulting in more complete iron removal under some conditions, and anecdotally less tendency for polymer degradation upon subsequent oxidation, though both of these claims should be verified.
Conference Paper
Previous studies have shown that waterflood recovery is dependent on the composition of injection brine in clastic reservoirs. Some researchers have also shown that oil recovery from carbonates is dependent on the ionic composition of the injection water. These studies have however been generated at laboratory conditions which are not representative of the reservoir, and therefore it is uncertain whether these IOR benefits are applicable to actual reservoir waterflood oil recovery. A reservoir condition coreflood study was therefore performed on core from a North Sea carbonate field (Valhall) to determine whether the recovery benefits seen in reduced condition experiments, were also obtained from full reservoir condition tests, using live crude oil and brine.In these reservoir condition tests, two reservoir core plugs were selected from the same reservoir layer and were similar in reservoir properties so that comparisons could be drawn between the experiments. Samples were prepared to give initial water saturations which were uniformly distributed and volumetrically matched to the height above the oil water contact of the samples in the reservoirs.The initial water saturation composition was based upon the simulated formation brine composition of the field. The plugs were then aged in live crude oil to restore wettability.Imbibition capillary pressure tests were then performed at full reservoir conditions, with live oil and brine, using the semi dynamic method.The first experiment utilised a simulated formation water and the second test utilised a simulated sea water, respectively, as the displacing water. The resultant data showed that the sea water used in the capillary pressure test modified the wettability of the carbonate system, changing the wettability of the rock to a more water wet state.This was indicated by comparing the saturation change in the spontaneous imbibition phase of the test between simulated formation and sea waters.
Article
One of the prominent enhanced oil recovery (EOR) methods in oil reservoirs is surfactant flooding. The purpose of this research is to study the effect of nanoparticles on the surfactant adsorption. Real reservoir sandstone rock samples were implemented in adsorption tests. The ranges of the initial surfactant and nano silica concentrations were from 500 to 5000 ppm and 500 ppm to 2000 ppm, respectively. The commercial surfactant used is sodium dodecyl sulfate (SDS) as an ionic surfactant and two different types of nano silica were employed. The rate of surfactant losses extremely depends on the concentration of surfactant in the system, and it was found that the adsorption of surfactant decreased with increasing the concentration of nano silica. Also, it was found that hydrophobic nano silica is more effective than hydrophilic nanoparticles.