Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

Accepted Manuscript

Adsorption of Silica Nanoparticles onto Calcite: Equilibrium, Kinetic, Ther-


modynamic and DLVO Analysis

A. Dehghan Monfared, M.H. Ghazanfari, M. Jamialahmadi, A. Helalizadeh

PII: S1385-8947(15)00944-4
DOI: http://dx.doi.org/10.1016/j.cej.2015.06.104
Reference: CEJ 13876

To appear in: Chemical Engineering Journal

Received Date: 28 April 2015


Revised Date: 23 June 2015
Accepted Date: 24 June 2015

Please cite this article as: A. Dehghan Monfared, M.H. Ghazanfari, M. Jamialahmadi, A. Helalizadeh, Adsorption
of Silica Nanoparticles onto Calcite: Equilibrium, Kinetic, Thermodynamic and DLVO Analysis, Chemical
Engineering Journal (2015), doi: http://dx.doi.org/10.1016/j.cej.2015.06.104

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Adsorption of Silica Nanoparticles onto Calcite: Equilibrium, Kinetic,

Thermodynamic and DLVO Analysis

A. Dehghan Monfared a,b, M. H. Ghazanfari a,∗, M. Jamialahmadib, A. Helalizadehb

a
Chemical and Petroleum Engineering Department, Sharif University of Technology, Tehran,

Iran
b
Petroleum Engineering Department, Petroleum University of Technology, Ahwaz, Iran

Abstract

Recently, application of silica nanoparticles (SNP) for enhancing oil recovery during water

flooding has been much attended. However, understanding how rock and nanoparticles (NP)

interacts through adsorption onto the carbonate reservoir rocks is not well discussed. In this

work, adsorption behavior of SNP onto the calcite had been characterized, through kinetic,

equilibrium, thermodynamics and electrokinetic studies as well as interaction energy analysis

by DLVO theory. Also, field emission scanning electron microscopy (FESEM) was utilized

to visualize the adsorption process. It had been found that kinetic behavior of SNP-calcite

system followed the pseudo-second order model. Equilibrium study revealed that Langmuir

model yielded a better fit which implied that monolayer coverage of SNP onto the calcite

surfaces was more probable. Analysis of kinetic data indicated that the intraparticle diffusion

mechanism was not the sole rate-controlling mechanism and the boundary layer diffusion

affected the adsorption to some extent. Thermodynamic study showed that physical/

electrostatic adsorption was expected to be the prevailing mechanism for the adsorption with

a spontaneous and endothermic nature. Also, adsorption process was enhanced by lowering


Corresponding author, email:ghazanfari@sharif.edu, Sharif University of Technology, Chemical and
Petroleum Engineering Department, Azadi Ave., Tel: +98-21-66166413, Fax: +98-21-6622853.

1
the pH and increasing ionic strength of the solutions. These findings were argued based on

the equilibrium reactions of the SNP and calcite in water and DLVO theory. Results revealed

that repulsive DLVO interaction energies predominated from which an unfavorable

attachment condition was deduced. In such situation, ionic strength of solution controlled the

extent of the SNP-calcite interaction. Reduction in the height of energy barrier and the

formation of a secondary minimum were responsible for enhancing the adsorption process at

higher ionic strengths. Furthermore, FESEM-observations depicted a spherical morphology

for the adsorbed SNP that uniformly distributed on the calcite surface.

Keywords: silica nanoparticles, calcite, adsorption, thermodynamics, DLVO, FESEM

1. Introduction

Since its introduction in 1960 [1], NP have been applied in several branches of science, and

studies on emerging nanotechnologies is still a topic of research. Some examples of

application include medical and biological science, electronics, environmental, agriculture,

etc. Over the past decade, experimental investigation showed that the petroleum industry

could also benefit from the great potential of nanotechnology for development in a variety of

fields. In this area, NP have been utilized in different forms such as: an additive for

improving properties of the drilling fluid [2, 3], well completion slurries [4, 5], a sensor for

reservoir characterization [6, 7] and an agent to control the fine migration[8]. Moreover,

nano-fluid (i.e. suspension of NP in a fluid) has been proved to have potential application for

enhanced oil recovery (EOR) [9, 10]. Among the different types of NP, SNP have been

mentioned to be suitable for EOR because of their ability to be easily surface functionalized

and some other advantages as noted by Miranda [11] . However, most of silica based NP

applied for EOR in sandstone formations [9, 11] and their application to carbonate systems

are not well discussed.

2
SNP have been used either individually or as part of a mixture to exploit their ability in

controlling some factors for EOR purposes. The use of surface-modified SNP for generating

a stable oil/water emulsion [12] and stabilizing supercritical CO2 foam with SNP whose

surface modified with PEG [13] are examples of latter. For the individual applications, the

first proposed mechanism is wettability alteration of reservoir rock. Based on wettability of

surface, three types of SNP are experimentally utilized for EOR: lipophobic and hydrophilic

polysilicon NP, hydrophobic and lipophilic polysilicon NP and neutrally wet polysilicon NP

[9, 14]. In conjunction with wettability alteration, some authors mentioned that interfacial

tension reduction can be another mechanism responsible for the increase in oil recovery [15].

Regardless of underlying mechanism, the interaction between the NP and reservoir rock plays

an important role in the performance of nano-fluids for EOR. Adsorption of NP onto the rock

surfaces not only can beneficially change the rock surface properties but also affects the

transport behavior and mobility of nano-fluid through porous media. Thus, a deep

understanding of this aspect may help to better design a successful NP-assisted EOR process.

Here, a mechanistic study on SNP-carbonate rock interaction is presented. For this purpose,

kinetics and equilibrium of adsorption process at different levels of ionic strength, pH,

temperature and SNP concentration are investigated. Also, thermodynamic and electrokinetic

studies as well as interaction energy analysis by Derjaguin-Landau-Verwey-Overbeek

(DLVO) energy profiles are utilized.

2. Materials and Methods

2.1. Materials

Nanoparticles: The NP used in this study were a type of ultrapure (99.999%) unmodified

surface SNP purchased from TECNAN (Navarrean Nanoproducts Technology, Spain). The

NP received in the form of white powder and had spherical morphology with the nominal

3
average size of 10-15 nm. Specific surface area and true density of particles were 180-270

m2/g and 2.2 g/cm3 respectively. A TEM image of SNP is shown in Fig. 1 (provided by

TECNAN).

Electrolyte: The electrolyte solution was prepared by the addition of sodium chloride (NaCl)

to deionized water. The purity of sodium chloride supplied by Merk Company was 99.5%.

Carbonate Rock Sample: The utilized carbonate rock sample was provided from

Aligoodarz Mine located in south west of Iran. The main constituent of the carbonate rock

was calcite. The purity of the sample is known to be as high as approximately 99% (based on

reported analysis by Zagros Powder Company). X-Ray Diffraction (XRD) technique was

used to check the purity. The resulted XRD pattern of the provided sample, depicted in Fig. 2,

revealed that all peaks were coincided with calcite and no additional peak was observed.

2.2 Methods

2.2.1. Nanoparticle dispersion preparation

Nano-fluid suspensions were prepared by addition of the NP powder to deionized water.

Dispersing was performed using a magnetic stirrer for 15 min. To provide a uniform colloidal

suspension and breaking the particle agglomerate, magnetic stirring was followed by

sonication process. Based on stability analysis, the optimum time for the sonication was

found to be 40 min. A stock solution of 0.2 % by weight was made. The prepared stock

solution was then diluted to the desired concentrations using suitable amount of deionized

water or electrolyte.

2.2.2. Zeta potential measurement

Zeta potential of both SNP and calcite was measured in the presence/absence of salt over a

range of pH (3.7-10 for SNP and 7.6-10 for calcite) by Malvern ZEN 3600 instrument. To do

so, a suspension of 400 mg/L and 2.5g/L for the SNP and crushed rock were prepared,

4
respectively. The salt concentrations were 0.001, 0.05 and 0.1 M. The pH of solutions was

adjusted by droplets of 0.1 M NaOH and 0.1 M HCl; both were supplied from Merk

Company.

2.2.3. Adsorption Experiment

The adsorption rate and capacity of SNP onto the calcite surface were determined via a series

of batch experiments. For this purpose, crushed calcite mineral (particles size was in the

range of 500-700 µm) was repeatedly washed with deionized water and dried. Then, 4 gr of

this cleaned crushed calcite was added to 32 ml of the SNP dispersion at different

concentrations and desired electrolyte or pH ranges. To investigate the equilibrium of

adsorption, a range of initial concentration for the SNP solution was selected (200-1200

mg/L). The experiments were conducted at three different temperatures: 293, 303 and 313 K.

The adsorption kinetic of the SNP was evaluated by analyzing the concentration of samples

taken at suitable time intervals. Having reached the equilibrium condition, the adsorption

isotherm was determined. The preliminary test revealed that the equilibrium condition was

reached after 30 hrs for 293 K. This time was found to be reduced to 24 hrs and 10 hrs for

temperatures of 303 K and 313 K, respectively. The concentration of the SNP in solutions

were back calculated using a calibration curve obtained by UV-visible spectrophotometer, as

utilized by other researchers[16].

2.2.4. Scanning Electron Microscopy

The adsorption of SNP onto the calcite surface was characterized by scanning electron

microscopy (SEM). Analysis was performed using a field emission scanning electron

microscope (FESEM; S-4160 Hitachi instrument).

5
3. Theoretical Background

3.1. Kinetics and Equilibrium Models of Adsorption

3.1.1. Adsorption Kinetics

A peer look at the kinetics of adsorption gives insight to explain the mechanism of interaction

between adsorbate and adsorbent. Three most common kinetic models are described here:

Pseudo-first order model: Linear formulation of pseudo-first order is shown by [17]:

ln −  =  −  (1)

where q(t) and qe are the amount of adsorption at time t and equilibrium per unit mass of

adsorbent (mg/g), respectively and k1 is the pseudo-first order rate constant (1/hr).

Pseudo-second order model: The linear form of pseudo-second order equation can be

expressed as [18]:

 1 
= +
  
(2)

where k2 is the pseudo-second order rate constant (g/mg.hr).

Intraparticle diffusion model: Intraparticle diffusion model is usually utilized to understand

the mechanism of adsorption. The mathematical form of this model is as follow [19, 20]:

 =  / +  (3)

where ki is the intraparticle diffusion rate constant (mg/g hr1/2) and Ci (mg/g)is the intercept

that indicate the boundary layer effect; i.e. the larger intercept is related to the greater

boundary layer effect.

3.1.2. Adsorption Isotherm

Adsorption isotherm models are used to evaluate the equilibrium of adsorption process.

Furthermore, they also resulted in some useful finding guiding to understand the inherent

adsorption behavior. Two common adsorption isotherms are discussed below.

6
Langmuir isotherm: Ervin Langmuir [21] developed an isotherm that originally proposed to

describe the adsorption characteristics of gas-solid system. The model assumes a physical and

monolayer adsorption onto a homogeneous surface [22] and usually formulated as:

  
 =
1 +   (4)

where qeq is the amount of the NP adsorption at the equilibrium per unit mass of adsorbent

(mg/gr), Qm is the maximum uptake capacity (mg/g), Ceq is the solution equilibrium

concentration (mg/L) and kL is Langmuir constant related to the adsorption site (L/mg).

Freundlich isotherm: Freundlich model [23] has empirical nature and suggests that a

multilayer sorption occurs onto a heterogeneous surface [22]. It can be expressed by:

 =  
/ (5)

where kF and n F are Freundlich constants. kF shows the adsorption capacity of adsorbent and

nF is related to the adsorption intensity.

3.2. Adsorption Thermodynamics

Thermodynamics parameters such as free energy change (∆Go), enthalpy change (∆Ho) and

entropy change (∆So) are estimated from change in the thermodynamic equilibrium constant,

KT, with temperature [24, 25]; defined as:

"! !
 = =
!
" 
(6)


where a s is the activity of the adsorbed adsorbate, ae is the activity of adsorbate in the

equilibrium solution, qs is the amount of adsorption in mmol of adsorbate per gram of

adsorbent, Ce is the adsorbate concentration in the equilibrium solution (mmol/mL), vs is the

activity coefficient of the adsorbed adsorbate and ve is the activity of adsorbate in the

equilibrium solution.

At infinite dilute solution; i.e. adsorbate concentration in the solution tends to zero, the values

of activity coefficients approach unity and the definition of KT is simplified to:

7
!
lim = = 
!
%& →( 
(7)


Having calculated the KT, the thermodynamics parameters; ∆Go (J/mol), ∆Ho (J/mol) and ∆So

(J/mol.K) can be estimated as [26, 27]:

∆* ( = −+,  (8)

∆-( ∆.(
 = −
+ +,
(9)

where R is the universal gas constant (8.314 J/mol.K) and T is the absolute temperature (K).

3.3. DLVO Theory

Developed by Derjaguin, Landau [28], Verwey and Overbeek [29] in 1940s, DLVO theory

was employed to quantitatively evaluate the stability of colloid in aqueous solution. This

theory states that the colloid stability is related to the total interaction potential, which is the

sum of attractive London-van der Waals energy (EvdW) and the electrostatic repulsion energy

(EEDL) for charged colloidal particles:

/012 = /345 + /60 (10)

However, DLVO theory can be used to describe the adsorption of colloids onto a surface. If

the colloidal particles are considered to be approximately spherical and the surface of the

adsorbent is assumed to be flat, the interaction between the particles and collectors can be

described using sphere-plate DLVO energy profile[30]. For a spherical particle interacting

with a flat plate, the van der Waals energy is related to the separation distance as [31, 32]:

78 + + ;
/345 = − : + + = >?
6 ; ; + 2+ ; + 2+
(11)

where AH is the Hamaker constant, R is the radius of the sphere, and x is the distance of

closest point on the sphere surface from the flat plate. The value of the Hamaker constant for

two media interacting across a third medium can be approximated based on their physical

properties. The detail of calculation can be found in [33].

8
Based on linearized Poisson-Boltzman equation, analytical expressions describing the

electrostatic repulsion energy of two infinite flat plates were developed. This expressions

were utilized to derive the sphere-flat plate interaction energy [31, 32]:

−coth [C; + + − +O1 − P/+  ]


I +coth [C; + + + +O1 − P/+  ] W
Y H 2D D V
/60 = @A( AB CD! + DE F H+   RSRℎ[C; + + − +O1 − P/+ ] V PXP
! E 
( H D! + DE V
H 2D V
! DE
−  RSRℎ[C; + + + +O1 − P/+  ]
G D! + DE
(12)
U

where ɛ0 is the vacuum permittivity(F/m), ɛr is the relative permittivity of medium, κ is the

inverse of Debye length (1/nm), ψp and ψs are the surface potentials of the plate and sphere,

respectively and r is the variable of integration. Usually, zeta potential is used as an

approximation for the surface potential [34].

4. Results and Discussions

4.1. Kinetics and Equilibrium of Adsorption

In this section, adsorption of SNP onto the calcite surface is analyzed by fitting the data to

kinetic and isotherm models.

Kinetics: Adsorption kinetic experiments were performed at two initial solution

concentrations of 800 and 1000 mg/L and the temperature of 293 K. To investigate the effect

of temperature, the experiments were repeated in 303 K, as well. Description of the kinetic

data using the pseudo-first order model was shown in Fig. 3. The slopes and intercepts of the

linear plots (ln [qe-q(t)] versus t ) were used to obtain k1 and qe. The parameters of pseudo-

second order model, i.e. k2 and qe, were determined from linear plots of t/q(t) versus t (Fig.

4). The estimated parameters are given in Table 1.

The correlation coefficient (R2) was used to evaluate the quality of fit for each case.

Moreover, the difference between experimental and calculated values of equilibrium

adsorption capacity was selected as another criterion.

9
Pseudo-first order fit resulted in relatively high R2 values. However, analysis revealed that

the pseudo-second order is likely to be the more appropriate model. Higher correlation

coefficient (R2>0.99) and closeness of the calculated and experimental equilibrium adsorption

capacity suggested that the adsorption kinetics of SNP onto calcite surfaces followed the

pseudo-second order kinetic. Inspection of the data showed that the rate constant (k2) and

equilibrium adsorption capacity, both, increased with the increase in nano-solution

concentration and temperature. This implied that at higher temperature and nano-solution

concentrations, the adsorption of SNP onto surfaces of the calcite rock was somewhat

boosted and had greater capacity.

Equilibrium: Fig. 5 shows the adsorption isotherm for SNP adsorption onto the calcite

surface at different temperatures. The equilibrium data were analyzed by Langmuir and

Freundlich isotherms. Linearizing Langmuir model, a straight line plot of Ceq/q eq versus Ceq

was constructed (Fig. 6(a)). Then, the slope and intercept were used to determine the

maximum uptake and Langmuir constant.

A similar linear regression scheme was applied to calculate the Freundlich constants using

the straight line plot of ln(qeq) versus ln(Ceq) as illustrated in Fig. 6(b). The quantitative

results of curve fitting at three distinct temperatures are listed in Table 2. Comparing the R2

values in this table showed that the Langmuir isotherm could better describe the equilibrium

data. Observation indicated that the maximum adsorption capacity increased with the increase

in temperature. This reminded the endothermic nature of adsorption that will be discussed in

the following section.

4.2. Adsorption Mechanisms

The following steps are commonly assumed to describe the mechanisms of the adsorption

process [20]: (i) transport of adsorbate from solution to the external surface of adsorbent (film

diffusion); (ii) pore diffusion or solid surface diffusion involving the diffusion of adsorbate

10
molecules to an adsorption site (intraparticle diffusion); (iii) adsorption of adsobate at a site

on the internal/external surface of adsorbent. The slowest step controls the adsorption rate.

The last step is often assumed to be relatively fast, and thus, it cannot be the rate-controlling

mechanism [20, 35]. Usually, either film or intraparticle diffusion or a combination of both is

considered to be the rate-controlling mechanism [20, 35].

Intraparticle diffusion model (Equation (3)) was applied to explore the rate-controlling

mechanism in the adsorption of SNP onto calcite. To achieve this, the plot of q(t) versus t1/2

was constructed (Fig. 7). Based on this model, if the plot was linear, the intraprticle diffusion

mechanism contributed in the adsorption. However, the observed multi-linearity in the

intraparticle diffusion plot (Fig. 7) implied that more than one mechanism had contributed in

the adsorption process. The three linear portions illustrated in Fig. 7 corresponded to three

steps: The first, steeper portion described the transport of SNP from the solution to the

external surface of calcite. The second portion was attributed to the intraparticle diffusion

step and the last portion is the final equilibrium. Intraparticle diffusion would be the only

rate-controlling step if the plot passed through the origin [36]. This was not the case here, as

shown in Fig. 7. Therefore, the intraparticle diffusion was not the sole rate-controlling

mechanism and the boundary layer (film) diffusion affected the adsorption to some extent.

4.3. Thermodynamics Parameters

Thermodynamics parameters estimation is a practical tool to understand the possibility and

nature of adsorption process. These parameters can be obtained from equilibrium adsorption

data at different temperatures. To do this, the equilibrium constant was determined by

plotting ln(q s/Ce) versus qs at different temperatures [37, 38] as illustrated in Fig. 8(a). The

temperatures were selected in the range of the utilized values for adsorption thermodynamics

11
studies [20, 39]. The value of KT for corresponding temperature can be found from the

intercept of the resulting straight line. The linear plots shown in Fig. 8(a) indicated a

relatively good correlation coefficient for the fitted data. The values of KT and the calculated

free energy change (∆Go) from Equation (8) are listed in Table 3. The negative values of ∆Go

suggested that the adsorption of SNP onto calcite is feasible and spontaneous. In addition, the

absolute value of ∆Go was increasing with the temperature. This demonstrated that the

adsorption was more spontaneous at higher temperatures. For a physical sorption the change

in the value of free energy is supposed to be in the range of −20 to 0 kJ/mol [40]. Here, the

free energy changes were less than −20 kJ/mol where physical/electrostatic adsorption was

expected to be the prevailing mechanism. Regarding Equation (9), ∆Ho and ∆So of the

adsorption equilibrium can be obtained from the slope and intercept in the plot of ln(KT)

versus 1/T, respectively [41, 42] (Fig. 8(b) and Table 3). The value of ∆Ho was calculated to

be 3.5892 kJ/mol. The positive value of ∆Ho demonstrated that the adsorption had an

endothermic nature. It was supported by previous observation which indicated that the value

of adsorption capacity rose at higher temperatures. The small value of ∆Ho was further

confirming the physical characteristic of the adsorption process [43]. The value of ∆S o was

found to be 23.4505 J/mol. The positive value of ∆So can be an indication of NP tendency to

adsorb onto the calcite surface. It also showed that the degree of disordering increased during

the adsorption process.

4.4. Surface Charge of SNP and Calcite

Surface charge of particles control the amount and extent of adsorption process. A peer look

on the adsorbate/adsorbent-water equilibria can give some insights about the surface charge

of materials. Additionally, electrokinetic study (i.e. zeta potential measurement) is a practical

method for determining the average surface charge of the particles. Based on the two

mentioned evaluations, the surface charge of both SNP and calcite are discussed below.

12
Zeta potential values of the SNP and calcite as a function of pH are presented in Fig. 9. The

values of zeta potential for three electrolyte concentrations (0.001, 0.05 and 0.1 M NaCl) at a

constant pH=8 are shown, as well. The data indicate that both SNP and calcite are negatively

charged in the studied range of pH and salinity. The iso-electric point (IEP) ,defined as the

pH at which the net charge of a surface is zero,can be estimated by extrapolating the curve to

the lower pH values . In this study, the IEP of SNP suspension is approximated to be about

2.8 which falls in the range of reported values in literature. [44]. The negative surface charge

of silica pertains to dissociation of the water molecules onto the silica surface and formation

of silanol groups (SiOH) as schematically illustrated in Fig. 10 [45, 46]. The silanol groups

are further dissociated through the following mechanism [47]:

−SiOH ⇋ −SiO^ + H_ (13)

The formation of −SiO^ sites causes the surface of silica to be negatively charged.

However, the negative zeta potential of calcite cannot be argued as simple as silica. Literature

contains relatively inconsistent information about the surface charge. Some researchers have

only reported negative zeta potential [48, 49] while others have obtained positive values [50-

52]. However, the IEP values in the range of 5.4-10.8 were also presented in some published

works [53-56]. Many authors argued this behavior. The nature of the calcite sample, pH of

the solution, the type and concentration of so called potential determining ions were some

commonly explained reasons [51, 55, 57, 58]. Furthermore, this inconsistency in the reported

values for calcite surface charge can be due to the relatively complex dissolution of calcite in

water. The governing equilibria which controls the system can be briefly described by the

following reactions[52, 59]:

CaCObs ⇋ Ca_ aq + Ceb^   k =10-8.48 (14)

Ceb^   + H el ⇋ HCeb^ aq + OH ^   k =10-3.7 (15)

13
HCeb^ aq + Hel ⇋ Heb aq + OH ^   k =10-7.7 (16)

H eb aq ⇋ H el + CO f k =101.47 (17)

Ca_ aq + HCeb^ aq ⇋ CaHCeb_   k =101.1 (18)

Ca_ aq + H el ⇋ CaOH _


 + ._ k =10-12.9 (19)

Ca_ aq + 2H el ⇋ CaOH  S + 2._   k =10-22.8 (20)

The associated values are equilibrium constants. According to k values for the last two

equations, the concentrations of Ca(OH)+ and Ca(OH)2 are very low and therefore can be

neglected. Presence of different ions of opposite charge gives rise to a difficult condition for

surface charge prediction. However, Cygan et al. [60] have studied the simulation of bulk and

surface structures of calcite and related carbonate phases using interatomic potential model

[61]. Based on their evaluation, the vacancy defect energy (i.e. the energy associated with

removing of an ion from the crystal to an infinite distance to form a charged and isolated

vacancy) for Ca_ was smaller than Ceb^ . Thus, Ca_ vacancies are more probable than

Ceb^ which implies that the surface of calcite is negatively charged. In this work, the

observed negative value for zeta potential can be due to the formation of prevalent Ca_

vacancies (i.e. negative sites) on the surface of calcite. IEP was approximated to be about 7.1

where located in the range of the reported values for calcite.

Despite the negative charges of both SNP and calcite surface in the tested pH range, the

adsorption process is still feasible. Attachment of particles to the surfaces under such a

condition where repulsive colloidal interactions predominate is termed as unfavorable [62]. In

many aquatic environments, colloids and solid surfaces (i.e. adsorbate and adsorbent) have

net negative charge and adsorption process occurs under the unfavorable attachment

condition [63, 64]. In such a condition, DLVO interaction energy profile is used to describe

the situation. However, this description is qualitative and it only gives some valuable

14
conceptual insights about the process. When the attachment is unfavorable, the energy profile

shows an energy barrier at some separation distance and it may also include a secondary

minimum (Fig. 13). Unfavorable attachment can pertain to the following phenomena: (i)

Fraction of colloidal particles may be retained in the secondary energy minimum [65-67]. (ii)

Some colloids may have enough energy to overcome the energy barriers to attachment in the

primary minimum [67].(iii) The intrinsic charge heterogeneity of colloids /solid surface can

provide some favorable site for attachment [68-70]. Here, the surface charge heterogeneity on

calcite surface can be due to contribution of both Ca_ and Ceb^ in the calcite-water

equilibria.

4.5. Effect of pH on adsorption

The effect of pH was investigated in the pH of 7.5, 8.5 and 10, in absence of electrolyte and

at a fixed temperature of 303 K. SNP concentration was selected to be 600 mg/L. The

conducted tests in acidic pH range did not yield reliable results and are not reported here,

because the pH values changed during the test based on following equilibrium reaction

between acid and calcite [71]:

CaCObs + 2. _ ⇋ Ca_ aq + Ce g + . e (21)

According to this reaction, the concentration of H+ ions decreased due to their consumption

over time. Therefore the pH of system shifted from acid to base. In this way, an adsorption

test with initial pH of 4 attained a pH of approximately 7.5 at the end of process. This

difficulty for working with calcite in acidic pH was also mentioned by other authors [72].

Fig. 11 shows how the adsorption of SNP onto the calcite surface decreases when pH

increases. Change in adsorption with the change in pH is due to the fact that pH can affect the

surface charge and electrostatic interaction between the NP and calcite surface. As the pH

increases, the formation of negative sites from the silanol groups on the silica surfaces are

enhanced based on the following reaction:

15
−SiOH + OH ^ ⇋ −SiO^ + HO (22)

As a result, the surface charge of SNP becomes more negative at basic pH and the zeta

potential decreases. As for calcite, according to the reactions (15) and (16) the excess of

eH ^ shifts the equilibrium to the left and increases the concentration of negative ions; Ceb^

and HCeb^ . Thus, the formation of negative sites on the surface is enhanced and a lower zeta

potential is resulted. Because both SNP and calcite surfaces will be more negative at higher

pH, the electrostatic repulsion between the surfaces becomes more prevalent and the

adsorption decreases.

4.6. Salinity Effect

To investigate the effect of salinity (or ionic strength), adsorption tests were conducted in the

ionic strength range of 0-0.2 M while the pH and temperature were kept constant at the

values of 8 and 303 K, respectively. The tests were also performed at a different SNP

concentration (400 mg/L) and temperature (293 K). The range of ionic strength was selected

based on the literature [73]. it is also worth mentioning that in high saline environments the

electrostatic repulsion vanishes and the pure attractive interaction is achieved. Thus, a

moderately low ionic strength environment was simulated to better investigate the salinity

effect. The results of this study can also be utilized to incorporate SNP into low salinity water

flooding for EOR.

The role of salinity on the adsorption of SNP onto the calcite surface is illustrated by plotting

the equilibrium adsorption versus ionic strength for different amount of NaCl (Fig. 12). The

conducted tests revealed the clear dependence of adsorption on ionic strength such that

increasing ionic strength resulted in higher amount of adsorption.

Similar to pH, the salinity can influence the surface charges and changes the state of the

electrostatic interaction. Fig. 9 shows that increase in ionic strength lowers the value of the

zeta potential for both silica and calcite to some extent. This reduction in zeta potential is

16
supposed to be due to compression of the double layer [34, 74]. Also, it can be ascribed to the

interaction of counter ion (Na_ ) with the surface or some anions in the solution. Na+ ions can

come into close contact with silica surfaces and screen their negative charge. Actually, Na_

ions make a complex with some −SiO^ site and neutralize their negative charge:

−SiO^ i _
⇋ −SiO^ + Na_ (23)

Therefore, the zeta potential of SNP becomes less negative in the presence of the opposite

charge of Na_ .

Also, Na_ is able to interact with the calcite surface. These ions can come closer to the

calcite surface relative to Ca_ and Cl^ . This was attributed to smaller hydration shells

around the Na_ ions. Compared to the Ca_ and Cl^ , when Na_ approaches the hydrated

surface of calcite, it pays a lower energy penalty by the removal or restructuring of its

hydration sell [75]. Thus, the Na_ ions neutralize some negative sites on the calcite surface

and reduce the net negative charge. In other words, the extent of the repulsive forces between

SNP and calcite decrease and the attractive forces dominate and the adsorption is enhanced.

Furthermore, the salinity effect can be described by DLVO interaction. Fig. 13 shows the

interaction energy profiles for SNP-calcite surface at different ionic strength. The value of

Hamaker constant for SNP interacting with calcite in the water was calculated to be 4.4×10 -21

using appropriate equations [33]. The properties of SNP, calcite and water were taken from

the literature [76]. When the zeta potential of both SNP and calcite were decreased by the

addition of electrolyte, the height of repulsive energy barrier on corresponding interaction

energy profile was reduced (Fig. 13). The predicted value of energy barrier by DLVO was

decreased from 29.66 kT at 0.001 M to 1.9 kT at 0.1 M. As a result, the chance increased for

some particles to overcome the barrier and fall in the primary minimum. On the other hand,

Fig. 13 demonstrated that a secondary minimum was formed in interaction energy profile for

some ionic strength and became deeper at higher ionic strengths. As mentioned before,

17
adsorption in secondary minimum was a proposed mechanism under unfavorable condition

and the amount of adsorption increased for a higher depth secondary minimum.

4.7. FESEM Visualization

The adsorbed SNP onto the calcite surface was characterized by FESEM observation. To

reduce the charging, imaging was performed on gold coated samples. FESEM image of fresh

calcite was shown in Figs. 14 (a) and (b). The picture depicted a rough surface with some

pores and cracks. In Figs. 14(c) and (d), FESEM images visualized the adsorption of SNP

onto the calcite surface. A spherical morphology for the SNP was observed clearly.

Additionally, FESEM in Fig. 14(c) and (d) confirmed that the adsorbed SNP had a uniform

spatial distribution on the calcite surface.

5. Conclusions

In this work, the adsorption behavior of SNP onto the calcite surfaces was investigated at

various conditions of the SNP concentration, ionic strength, pH and temperature. Adsorption

isotherms, adsorption kinetic, thermodynamic study, and elecktrokinetic data as well as

DLVO theory were utilized to conceptually understand some aspects of this adsorption

process. Based on the results obtained in this work, the following conclusions can be drawn:

• Kinetic of SNP adsorption onto the calcite surfaces can be better described by the

pseudo-second order model. Both rate constant and equilibrium adsorption capacity

increased at higher temperatures.

• For the equilibrium of adsorption, Langmuir isotherm presented a better fit than

Freundlich isotherm. Therefore, monolayer coverage onto a homogenous surface can

be proposed. The trend of the maximum adsorption capacity changes with

temperature reminds the endothermic nature of the adsorption process.

18
• Evaluation of kinetic data based on the intraprticle diffusion model indicated that the

intraparticle diffusion mechanism was not the sole rate-controlling mechanism and the

boundary layer diffusion affected the adsorption to some extent.

• Thermodynamic study showed that a physical/electrostatic adsorption was expected to

be the prevailing mechanism for adsorption process. The estimated values of the

thermodynamics parameters showed that the adsorption of SNP onto the calcite

surfaces is feasible, spontaneous and endothermic. The latter supported the pervious

finding for the equilibrium analysis at different temperatures.

• Electrokinetic study revealed that the adsorption process occurred under unfavorable

attachment condition. Because both silica and calcite surfaces had negative charge

under the range of tested pH and salinity.

• For the studied range of pH, the amount of equilibrium adsorption was decreasing

with the pH. This finding was argued based on the change in surface charge and

electrostatic interaction between the SNP and the calcite surface at different pH. It

was also supported by the inspection of the equilibrium reaction in calcite/water and

silica/water systems.

• Ionic strength of solution played an important role on the adsorption of SNP onto the

calcite; i.e. adsorption was enhanced at higher ionic strength. DLVO interaction

energy profile was applied for interpretation. It was revealed that repulsive DLVO

interaction energies predominated from which an unfavorable attachment condition

was deduced. In such situation, ionic strength of the solution controls the extent of

SNP-calcite interaction. Reduction in the height of energy barrier and the formation of

a secondary minimum are thought to be responsible for enhancing the adsorption

process at higher ionic strengths.

19
• FESEM observations depicted a spherical morphology for the adsorbed SNP that

uniformly distributed on the calcite surface.

References

[1] R.P. Feynman, There's plenty of room at the bottom, Engineering and science 23 (1960)
22-36.
[2] J.K.M. William, S. Ponmani, R. Samuel, R. Nagarajan, J.S. Sangwai, Effect of CuO and
ZnO nanofluids in xanthan gum on thermal, electrical and high pressure rheology of water-
based drilling fluids, Journal of Petroleum Science and Engineering 117 (2014) 15-27.
[3] R. Krishnamoorti, Extracting the benefits of nanotechnology for the oil industry, Journal
of petroleum technology 58 (2006).
[4] H. Li, H.-g. Xiao, J. Yuan, J. Ou, Microstructure of cement mortar with nano-particles,
Composites Part B: Engineering 35 (2004) 185-189.
[5] J.N. de Paula, J.M. Calixto, L.O. Ladeira, P. Ludvig, T.C.C. Souza, J.M. Rocha, A.A.V.
de Melo, Mechanical and rheological behavior of oil-well cement slurries produced with
clinker containing carbon nanotubes, Journal of Petroleum Science and Engineering 122
(2014) 274-279.
[6] M. Alaskar, K. Li, R. Horne, Transport of Temperature Nanosensors Through Fractured
Tight Rock: An Experimental Study, SPE Saudi Arabia Section Technical Symposium and
Exhibition, Society of Petroleum Engineers, 2014.
[7] M. Prodanovic, S. Ryoo, A.R. Rahmani, R.V. Kuranov, C. Kotsmar, T.E. Milner, K.P.
Johnston, S.L. Bryant, C. Huh, Effects of magnetic field on the motion of multiphase fluids
containing paramagnetic nanoparticles in porous media, SPE Improved Oil Recovery
Symposium, Society of Petroleum Engineers, 2010.
[8] A. Habibi, M. Ahmadi, P. Pourafshary, Y. Al-Wahaibi, Reduction of Fines Migration by
Nanofluids Injection: An Experimental Study, SPE Journal 18 (2012) 309-318.
[9] B. Ju, T. Fan, Z. Li, Improving water injectivity and enhancing oil recovery by wettability
control using nanopowders, Journal of Petroleum Science and Engineering 86–87 (2012)
206-216.
[10] B.A. Suleimanov, F.S. Ismailov, E.F. Veliyev, Nanofluid for enhanced oil recovery,
Journal of Petroleum Science and Engineering 78 (2011) 431-437.
[11] C.R. Miranda, L.S. De Lara, B.C. Tonetto, Stability and mobility of functionalized silica
nanoparticles for enhanced oil recovery application, Proceedings of the SPE International
Oilfield Nanotechnology Conference, 2012, pp. 12-14.
[12] T. Zhang, M. Roberts, S.L. Bryant, C. Huh, Foams and emulsions stabilized with
nanoparticles for potential conformance control applications, SPE International Symposium
on Oilfield Chemistry, Society of Petroleum Engineers, 2009.
[13] D.A. Espinoza, F.M. Caldelas, K.P. Johnston, S.L. Bryant, C. Huh, Nanoparticle-
stabilized supercritical CO2 foams for potential mobility control applications, SPE Improved
Oil Recovery Symposium, Society of Petroleum Engineers, 2010.
[14] M.O. Onyekonwu, N.A. Ogolo, Investigating the Use of Nanoparticles in Enhancing Oil
Recovery, Nigeria Annual International Conference and Exhibition, Society of Petroleum
Engineers, Tinapa - Calabar, Nigeria, 2010.
[15] L. Hendraningrat, S. Li, O. Torsæter, A coreflood investigation of nanofluid enhanced
oil recovery, Journal of Petroleum Science and Engineering 111 (2013) 128-138.
[16] H.F. Lecoanet, J.-Y. Bottero, M.R. Wiesner, Laboratory Assessment of the Mobility of
Nanomaterials in Porous Media, Environmental Science & Technology 38 (2004) 5164-5169.

20
[17] S.Y. Lagergren, Zur Theorie der sogenannten Adsorption gelöster Stoffe, 1898.
[18] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process
Biochemistry 34 (1999) 451-465.
[19] W.J. Weber, J.C. Morris, Kinetics of adsorption on carbon from solution, Journal of the
Sanitary Engineering Division 89 (1963) 31-60.
[20] N. Suriyanon, P. Punyapalakul, C. Ngamcharussrivichai, Mechanistic study of
diclofenac and carbamazepine adsorption on functionalized silica-based porous materials,
Chemical Engineering Journal 214 (2013) 208-218.
[21] I. Langmuir, THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS,
MICA AND PLATINUM, Journal of the American Chemical Society 40 (1918) 1361-1403.
[22] K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm systems,
Chemical Engineering Journal 156 (2010) 2-10.
[23] H. Freundlich, Over the adsorption in solution, Journal of Physical Chemistry 57 (1906)
385.
[24] Y.-H. Li, Z. Di, J. Ding, D. Wu, Z. Luan, Y. Zhu, Adsorption thermodynamic, kinetic
and desorption studies of Pb2+ on carbon nanotubes, Water Research 39 (2005) 605-609.
[25] Y.-J. Tu, C.-F. You, C.-K. Chang, Kinetics and thermodynamics of adsorption for Cd on
green manufactured nano-particles, Journal of Hazardous Materials 235–236 (2012) 116-122.
[26] Manasi, V. Rajesh, N. Rajesh, Adsorption isotherms, kinetics and thermodynamic
studies towards understanding the interaction between a microbe immobilized polysaccharide
matrix and lead, Chemical Engineering Journal 248 (2014) 342-351.
[27] G. Zhao, J. Li, X. Wang, Kinetic and thermodynamic study of 1-naphthol adsorption
from aqueous solution to sulfonated graphene nanosheets, Chemical Engineering Journal 173
(2011) 185-190.
[28] B.V. Derjaguin, L. Landau, Theory of the Stability of Strongly Charged Lyophobic Sols
and of the Adhesion of Strongly Charged Particles in Solutions of Electrolytes, Acta Phys.
Chim. URSS 14 (1941) 633-662.
[29] J.O. E. Verwey, Theory of the Stability of Lyophobic Colloids, Elsevier, Amsterdam.
(1948).
[30] B. Derjaguin, Friction and adhesion. IV. The theory of adhesion of small particles,
Kolloid Zeits 69 (1934) 155-164.
[31] S. Bhattacharjee, M. Elimelech, Surface element integration: a novel technique for
evaluation of DLVO interaction between a particle and a flat plate, Journal of colloid and
interface science 193 (1997) 273-285.
[32] K.A. Dunphy Guzman, M.P. Finnegan, J.F. Banfield, Influence of Surface Potential on
Aggregation and Transport of Titania Nanoparticles, Environmental Science & Technology
40 (2006) 7688-7693.
[33] J.N. Israelachvili, Intermolecular and Surface Forces, Elsevier Science2010.
[34] M. Elimelech, J. Gregory, X. Jia, R. Williams, Particle Deposition and Aggregation:
Measurement, Modelling and Simulation 1995, Butterworth-Heinemann: Oxford, UK.
[35] A.B. Albadarin, C. Mangwandi, A.a.H. Al-Muhtaseb, G.M. Walker, S.J. Allen, M.N.M.
Ahmad, Kinetic and thermodynamics of chromium ions adsorption onto low-cost dolomite
adsorbent, Chemical Engineering Journal 179 (2012) 193-202.
[36] Y.-S. Ho, Removal of copper ions from aqueous solution by tree fern, Water Research
37 (2003) 2323-2330.
[37] J. Biggar, M. Cheung, Adsorption of picloram (4-amino-3, 5, 6-trichloropicolinic acid)
on Panoche, Ephrata, and Palouse soils: a thermodynamic approach to the adsorption
mechanism, Soil Science Society of America Journal 37 (1973) 863-868.
[38] A.A. Khan, R.P. Singh, Adsorption thermodynamics of carbofuran on Sn (IV)
arsenosilicate in H+, Na+ and Ca2+ forms, Colloids and Surfaces 24 (1987) 33-42.

21
[39] T.S. Anirudhan, P. Senan, Adsorption characteristics of cytochrome C onto cationic
Langmuir monolayers of sulfonated poly(glycidylmethacrylate)-grafted cellulose: Mass
transfer analysis, isotherm modeling and thermodynamics, Chemical Engineering Journal 168
(2011) 678-690.
[40] M.J. Jaycock, G.D. Parfitt, Chemistry of interfaces, E. Horwood; Halsted Press1981.
[41] M. Toor, B. Jin, Adsorption characteristics, isotherm, kinetics, and diffusion of modified
natural bentonite for removing diazo dye, Chemical Engineering Journal 187 (2012) 79-88.
[42] J. Fu, Z. Chen, M. Wang, S. Liu, J. Zhang, J. Zhang, R. Han, Q. Xu, Adsorption of
methylene blue by a high-efficiency adsorbent (polydopamine microspheres): Kinetics,
isotherm, thermodynamics and mechanism analysis, Chemical Engineering Journal 259
(2015) 53-61.
[43] S. Glasstone, Textbook of physical chemistry, (1951).
[44] S.V. Patwardhan, F.S. Emami, R.J. Berry, S.E. Jones, R.R. Naik, O. Deschaume, H.
Heinz, C.C. Perry, Chemistry of aqueous silica nanoparticle surfaces and the mechanism of
selective peptide adsorption, Journal of the American Chemical Society 134 (2012) 6244-
6256.
[45] T.A. Michalske, S.W. Freiman, A molecular interpretation of stress corrosion in silica,
(1982).
[46] T. Mahadevan, S. Garofalini, Dissociative chemisorption of water onto silica surfaces
and formation of hydronium ions, The Journal of Physical Chemistry C 112 (2008) 1507-
1515.
[47] R.K. Iler, The chemistry of silica, Wiley, New York, 1979.
[48] H. Douglas, R. Walker, The electrokinetic behaviour of Iceland Spar against aqueous
electrolyte solutions, Trans. Faraday Soc. 46 (1950) 559-568.
[49] M. Smani, P. Blazy, J. Cases, Beneficiation of sedimentary Moroccan phosphate ores,
AIME Trans 258 (1975) 168-182.
[50] Y.C. Huang, F.M. Fowkes, T.B. Lloyd, N.D. Sanders, Adsorption of calcium ions from
calcium chloride solutions onto calcium carbonate particles, Langmuir 7 (1991) 1742-1748.
[51] T. Berlin, A. Khabakov, Differences in the electrokinetic potentials of carbonate
sedimentary rocks of different origin and composition, Geochemistry 3 (1961) 217-230.
[52] R. Eriksson, J. Merta, J.B. Rosenholm, The calcite/water interface: I. Surface charge in
indifferent electrolyte media and the influence of low-molecular-weight polyelectrolyte,
Journal of colloid and interface science 313 (2007) 184-193.
[53] P. Somasundaran, G. Agar, The zero point of charge of calcite, Journal of Colloid and
Interface Science 24 (1967) 433-440.
[54] M. Fuerstenau, G. Gutierrez, D. Elgillani, The influence of sodium silicate in non-
metallic flotation systems, Trans. AIME 241 (1968) 319-323.
[55] J.O. Amankonah, P. Somasundaran, Effects of dissolved mineral species on the
electrokinetic behavior of calcite and apatite, Colloids and Surfaces 15 (1985) 335-353.
[56] A. Martinez-Luevanos, A. Uribe-Salas, A. Lopez-Valdivieso, Mechanism of adsorption
of sodium dodecylsulfonate on celestite and calcite, Minerals engineering 12 (1999) 919-936.
[57] D. Siffert, P. Fimbel, Parameters affecting the sign and magnitude of the eletrokinetic
potential of calcite, Colloids and surfaces 11 (1984) 377-389.
[58] P. Smallwood, Some aspects of the surface chemistry of calcite and aragonite Part I: An
electrokinetic study, Colloid and Polymer Science 255 (1977) 881-886.
[59] P. Somasundaran, J.O. Amankonah, K. Ananthapadmabhan, Mineral—solution
equilibria in sparingly soluble mineral systems, Colloids and Surfaces 15 (1985) 309-333.
[60] R.T. Cygan, K. Wright, D.K. Fisler, J.D. Gale, B. Slater, Atomistic models of carbonate
minerals: Bulk and surface structures, defects, and diffusion, Molecular Simulation 28 (2002)
475-495.

22
[61] D.K. Fisler, J.D. Gale, R.T. Cygan, A shell model for the simulation of rhombohedral
carbonate minerals and their point defects, American Mineralogist 85 (2000) 217-224.
[62] M. Elimelech, C.R. O'Melia, Kinetics of deposition of colloidal particles in porous
media, Environmental Science & Technology 24 (1990) 1528-1536.
[63] C.R. O'melia, Particle—particle interactions in aquatic systems, Colloids and surfaces 39
(1989) 255-271.
[64] W. Stumm, Chemistry of the Solid-Water Interface: Process at the Mineral-Water and
Particle-Water Interface in Natural Systems, John Wily & Sons, New York 347 (1992).
[65] A. Franchi, C.R. O'Melia, Effects of natural organic matter and solution chemistry on the
deposition and reentrainment of colloids in porous media, Environmental science &
technology 37 (2003) 1122-1129.
[66] S.L. Walker, J.A. Redman, M. Elimelech, Role of Cell Surface Lipopolysaccharides in
Escherichia c oli K12 Adhesion and Transport, Langmuir 20 (2004) 7736-7746.
[67] N. Tufenkji, M. Elimelech, Deviation from the Classical Colloid Filtration Theory in the
Presence of Repulsive DLVO Interactions, Langmuir 20 (2004) 10818-10828.
[68] B.D. Bowen, N. Epstein, Fine particle deposition in smooth parallel-plate channels,
Journal of colloid and interface science 72 (1979) 81-97.
[69] H. Kihira, E. Matijevic, Kinetics of heterocoagulation. 3. Analysis of effects causing the
discrepancy between the theory and experiment, Langmuir 8 (1992) 2855-2862.
[70] N. Tufenkji, M. Elimelech, Breakdown of colloid filtration theory: Role of the secondary
energy minimum and surface charge heterogeneities, Langmuir 21 (2005) 841-852.
[71] W. Stumm, L. Sigg, B. Sulzberger, Chemistry of the Solid-Water Interface: Processes at
the Mineral-Water and Particle-Water Interface in Natural Systems, Wiley1992.
[72] A. Martínez-Luévanos, A. Uribe-Salas, A. Lopez-Valdivieso, Mechanism of adsorption
of sodium dodecylsulfonate on celestite and calcite, Minerals Engineering 12 (1999) 919-
936.
[73] I.G. Godinez, C.J.G. Darnault, A.P. Khodadoust, D. Bogdan, Deposition and release
kinetics of nano-TiO2 in saturated porous media: Effects of solution ionic strength and
surfactants, Environmental Pollution 174 (2013) 106-113.
[74] D. Liu, P.R. Johnson, M. Elimelech, Colloid deposition dynamics in flow-through
porous media: Role of electrolyte concentration, Environmental science & technology 29
(1995) 2963-2973.
[75] M. Ricci, P. Spijker, F. Stellacci, J.-F. Molinari, K. Voïtchovsky, Direct visualization of
single ions in the stern layer of calcite, Langmuir 29 (2013) 2207-2216.
[76] H.-J. Butt, M. Kappl, Surface and interfacial forces, John Wiley & Sons2009.

23
Table 1. Pseudo-first order and Pseudo-second order adsorption kinetic parameters
Initial Con. Temperature Pseudo-first order model Pseudo-second order model
(mg/L) (K)
qe, exp (mg/g) k1 (1/hr) qe (mg/g) R2 1 k2 (g/mg.hr) qe (mg/g) R22

800 293 1.461984 0.2017 0.975212 0.9182 0.323972 1.562256 0.9995


800 303 1.595136 0.2266 0.921272 0.9848 0.412838 1.688904 0.9996
1000 293 1.583816 0.3362 1.230598 0.9274 0.404946 1.672800 0.9969
1000 303 1.728064 0.3532 1.162415 0.9668 0.563169 1.807011 0.9990

24
Table 2. Equilibrium adsorption parameters for Langmuir and Freundlich isotherms.
Isotherms Parameters Temperature (K)

293 303 313


Langmuir Qm (mg/g) 1.857355 2.034588 2.617801
kL (L/mg) 0.00702 0.007317 0.007913
R2L 0.9983 0.9946 0.9978

Freundlich kF 0.228459 0.279627 0.306144


nF 3.445899 3.646973 3.299241
R2F 0.9761 0.9725 0.9777

25
Table 3. Obtained thermodynamics parameters for adsorption of SNP onto calcite
Temperature (K) Thermodynamics Constant

KT ∆Go (kJ/mol)
293 3.8367 -3.27548
303 4.0605 -3.53009
313 4.2147 -3.74359

26
Figure Caption

Fig. 1. TEM image of SNP (provided by TECNAN).

Fig. 2. XRD pattern of calcite sample.

Fig. 3. Pseudo-first order kinetic model and experimental data for adsorption of SNP onto the

calcite surface; (a) T=293 K, (b) T=303 K.

Fig. 4. Pseudo-second order kinetic model and experimental data for adsorption of SNP onto

the calcite surface; (a) T=293 K, (b) T=303 K.

Fig. 5. Adsorption isotherm data for SNP adsorption onto calcite.

Fig. 6. Linear plots of isotherm models fitted to adsorption data (a) Langmuir isotherm,

(b) Freundlich isotherm.

Fig. 7. Inraparticle diffusion plots for SNP adsorption onto calcite.

Fig. 8. (a) Linear plot of ln(qs/Ce) versus q s for estimating the equilibrium constant at

different temperatures. (b) Linear plot of ln(KT) versus (1/T) for estimation of the

thermodynamics parameters.

Fig. 9. Zeta potential for the SNP and calcite in deionized water, 0.001 M, 0.05 M and 0.1 M

NaCl.

Fig. 10. Dissociation of water molecules onto the silica surface and formation of silanol

groups:(a) Adsorption of water molecules to the siloxane (Si−O−Si) bound. (b) and (c)

reaction of water with siloxane. (d) Rupture of the Si−O bonds in siloxane and formation of

SiOH

Fig. 11. Effect of pH on adsorption of SNP onto the calcite surface.

Fig. 12. Effect of IS on the adsorption of SNP onto the calcite surface.

27
Fig. 13. Interaction energy profile for SNP-Calcite in 0.001, 0.05 and 0.1 M NaCl and T=293

K. The small figure shows an enlargement for secondary energy minima.

Fig. 14. FESEM images of (a, b) fresh calcite surface (c, d) adsorbed SNP onto the calcite

surface.

28
Fig. 1.

29
Fig. 2.

30
Fig. 3.

31
Fig. 4.

32
Fig. 5.

33
Fig. 6.

34
Fig. 7.

35
Fig. 8.

36
Fig. 9.

37
Fig. 10.

38
Fig. 11.

39
Fig. 12.

40
Fig. 13.

41
Fig. 14.

42
Research Highlights

• Pseudo-second order model can better describe the kinetic data.

• Monolayer coverage of SNP onto calcite surfaces is more probable.

• Both intraparticle diffusion and film diffusion mechanisms are involved.

• The adsorption process is spontaneous and endothermic.

• DLVO interaction energy profile is applied for interpretation of adsorption process.

You might also like