Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6....

11
Contents lists available at ScienceDirect Colloids and Surfaces A journal homepage: www.elsevier.com/locate/colsurfa Microscopic and physicochemical studies of polymer-modied kaolinite suspensions Cao J. a , Kang X. b , Bate B. c, a Department of Civil, Architectural, and Environmental Engineering, Missouri University of Science and Technology, Rolla, MO, United States b College of Civil Engineering, Hunan University, PR China c Institute of Geotechnical Engineering, College of Civil Engineering and Architecture, Zhejiang University, PR China GRAPHICAL ABSTRACT ARTICLE INFO Keywords: Biopolymers Zeta potential Microscopic particle size Time-lapse ABSTRACT Sedimentation tests on kaolinite in biopolymers (xanthan gum, chitosan, polyacrylic acid and polyacrylamide) solutions were conducted and recorded with time-lapsed technique up to 35 days. Microscopic particle size, zeta potential, nal volume and solid content were measured along the elevation of the graduated cylinder at the end of tests, and settling velocity and intensity calculated. Test results suggested that positive charged edges of kaolinite particles attached to negatively-charged xanthan gum (0.001 and 0.5 g/l) long chain via electrostatic force, exhibiting edge-to-edge (EE) fabric structure, and that charge neutralization was the primary interaction mechanism between kaolinite and either chitosan (0.05 and 5 g/l) or PAA(0.05 and 1 g/l),and that steric sta- bilization dominates the interaction between kaolinite and PAM molecules at 0.1 and 0.5 g/l PAM solutions. 1. Introduction Grain size distribution (GSD) of the aggregates (or ocs) of a ne- grained soil in marine or lacustrine environments has been studied primarily in mesoscale (0.25 μm to 1 mm) due to its relevance to the macroscale mechanical or uid dynamic behaviors [1]. At microscopic (sub-micron) scale, however, GSD and the physicochemical properties of ne-grained soils, which govern the fabrics and characteristics of https://doi.org/10.1016/j.colsurfa.2018.06.019 Received 14 December 2017; Received in revised form 26 May 2018; Accepted 8 June 2018 Corresponding author at: Institute of Geotechnical Engineering, College of Civil Engineering and Architecture, Zhejiang University, PR China. E-mail addresses: [email protected] (J. Cao), [email protected] (X. Kang), [email protected] (B. Bate). Colloids and Surfaces A 554 (2018) 16–26 0927-7757/ © 2018 Elsevier B.V. All rights reserved. T

Transcript of Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6....

Page 1: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

Contents lists available at ScienceDirect

Colloids and Surfaces A

journal homepage: www.elsevier.com/locate/colsurfa

Microscopic and physicochemical studies of polymer-modified kaolinitesuspensions

Cao J.a, Kang X.b, Bate B.c,⁎

a Department of Civil, Architectural, and Environmental Engineering, Missouri University of Science and Technology, Rolla, MO, United Statesb College of Civil Engineering, Hunan University, PR Chinac Institute of Geotechnical Engineering, College of Civil Engineering and Architecture, Zhejiang University, PR China

G R A P H I C A L A B S T R A C T

A R T I C L E I N F O

Keywords:BiopolymersZeta potentialMicroscopic particle sizeTime-lapse

A B S T R A C T

Sedimentation tests on kaolinite in biopolymers (xanthan gum, chitosan, polyacrylic acid and polyacrylamide)solutions were conducted and recorded with time-lapsed technique up to 35 days. Microscopic particle size, zetapotential, final volume and solid content were measured along the elevation of the graduated cylinder at the endof tests, and settling velocity and intensity calculated. Test results suggested that positive charged edges ofkaolinite particles attached to negatively-charged xanthan gum (0.001 and 0.5 g/l) long chain via electrostaticforce, exhibiting edge-to-edge (EE) fabric structure, and that charge neutralization was the primary interactionmechanism between kaolinite and either chitosan (0.05 and 5 g/l) or PAA(0.05 and 1 g/l),and that steric sta-bilization dominates the interaction between kaolinite and PAM molecules at 0.1 and 0.5 g/l PAM solutions.

1. Introduction

Grain size distribution (GSD) of the aggregates (or flocs) of a fine-grained soil in marine or lacustrine environments has been studied

primarily in mesoscale (0.25 μm to 1mm) due to its relevance to themacroscale mechanical or fluid dynamic behaviors [1]. At microscopic(sub-micron) scale, however, GSD and the physicochemical propertiesof fine-grained soils, which govern the fabrics and characteristics of

https://doi.org/10.1016/j.colsurfa.2018.06.019Received 14 December 2017; Received in revised form 26 May 2018; Accepted 8 June 2018

⁎ Corresponding author at: Institute of Geotechnical Engineering, College of Civil Engineering and Architecture, Zhejiang University, PR China.E-mail addresses: [email protected] (J. Cao), [email protected] (X. Kang), [email protected] (B. Bate).

Colloids and Surfaces A 554 (2018) 16–26

0927-7757/ © 2018 Elsevier B.V. All rights reserved.

T

Page 2: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

flocs, are still not well understood. On the other hand, Stoke’s law (Eq.(1)) predicted the terminal velocity of small particles in a suspension:

=v Fπμr6

d

(1)

Where Fd is the frictional force (unit: N) acting on the interface betweenfluid and particle, also known as Stokes’ drag, μ the dynamic viscosity(unit: Pa), r (m) the radius of the spherical object, and v (m/s) the flowvelocity relative to the object. However, the real sedimentation processis complicated by particle colliding [2,3], Coulombic interactions [4],van der Waals interactions [5–7], and Brownian motions [1,2], whichrender significant difficulties in simulating the sedimentation process offine grain soils. Zeta potential, which is a good indicator of the inter-particle forces, yields insight into the interaction between the clayparticles and the fluid phase [8,9].

Polymers emerged as new engineering materials in recent decadesdue to their environmental-friendly nature, minimal carbon footprint,and high efficiency. Polymer modification has been widely used inseveral disciplines, including chemical enhanced oil recovery [10–13],soil erosion control [14–18], dewatering of mine tailings [19–21],dredging of sediments [22–24], waste water treatment [25,26] and soilstabilizations [27–29]. Despite these applications, the interaction me-chanisms between polymers and soil particles have not been fully ex-plored. Polymer bridging and charge neutralization are the two majorinteraction mechanisms [30–32].Polymer bridging occurs when soilparticles are brought together, sometimes forming flocculation, viahydrogen bonding between the end functional groups on the polymerchain and the sites on clay surfaces [7,21,33,34]. Charge neutralization,on the other hand, occurs when polymer chain with cationic functionalgroups adsorbed onto clay mineral surfaces and modified the particlefabrics [21]. Besides, depletion flocculation and stabilization are theless common mechanisms which are generated by dissolving nonionicpolymer into the dispersions. The depletion interactions are induced bythe unbalanced osmotic force caused by the exclusion of the free non-adsorbing polymer molecules from the space between the two ap-proaching particles [35]. The effects of polymers on the clay particlesbehaviors are complicated, and the acting mechanisms often depend onsolid concentration [36], cation exchange capacity [37], polymer do-sage and molecular weight [38], adsorption density of polymer [7],types of polymer functional groups [30] and pH [39,40].

Time-lapse visualization of sedimentation process has been used foryears in the study of glaciers [41], soil movement and erosions [42–44]and earth flow movement [45]. Time-lapse technique in monitoring thesedimentation process saves labor and provides turbidity informationthat linked to final volume, settling rate and maybe even solid content,even though the aqueous turbidity could be directly detected by Ana-lytic Jena Specord S 600 BU machine [46].

This study aims at elucidating the fundamental mechanisms on theeffects of sodium chloride (NaCl) and four polymers (xanthan gum,chitosan, polyacrylic acid and polyacrylamide) with kaolinite. Toachieve this goal, the following tasks will be performed: First, a series ofsedimentation tests for kaolinite and polymer mixtures with time-lapsed recording were performed. When final volume was reached,particle size, zeta potential and solid content were measured. Finally,the relationship between the microscopic/physicochemical propertiesand the macroscopic sedimentation behaviors (final volume, settlingvelocity and intensity) were proposed.

2. Materials and methodology

Georgia kaolinite (RP-2, properties are shown in Table 1) washomoionized with sodium cations before use. The homoionizationprocess is similar to that in Bate and Burns [8], and a brief descriptionwas given below. 2 kg of kaolinite was added to 14 L of 2.0 mol/l NaClsolution. The resulting suspension was mechanically stirred for 15minand allowed to stand for at least 24 h for gravity separation. The

supernatant was then siphoned off, and the solids were rinsed with de-ionized water to remove any loosely bound cations. This process wasrepeated until the electrical conductivity of the supernatant was be-low350μS/cm.

Traditional method of oven-drying a slurry and mechanicalgrounding to obtain the solid powder will break kaolinite plates [47],and therefore was not adopted in this study. Instead, after Na-homo-ionization, the slurry was manually stirred for 5min to obtain a uniformsuspension, which was siphoned into graduated cylinders for the sedi-mentation tests. In order to obtain consistent solid content, suspensionvolume calibration tests were performed. It was found out that 85ml ofuniform suspension contained consistently 10 ± 0.1 g of solids.Therefore, 85ml of uniform suspension was extracted for each 100mlgraduated cylinder, and approximately 15ml of solutions with pre-scribed chemical concentrations was added to reach the full volume of100ml. The resulting chemical concentrations of different chemicals(NaCl, Xanthan gum, Chitosan, polyacrylic acid or polyacrylamide) in100ml kaolinite slurries were summarized in Table 2. The molecularstructures of xanthan gum (Pfaltz&Bauer, CAS#: 11138-66-2), chitosan(Alfa Aesar, CAS#: 9012-76-4. 85% deacetylated), polyacrylic acid(PAA) (Polysciences, Inc., Lot#: 541449) and polyacrylamide (PAM)(Acros Organics, CAS#: 62649-23–4) used in this study were shown inFig. 1. Molecular weights of above polymers are given in Table 3.

A 360° rotational sedimentation panel (Fig. 2) was developed in thisstudy to simultaneously measures lurries in up to 28graduated cylin-ders. Graduated cylinders were locked onto the shelves of the panel.After slowly rotating the sedimentation panel for 5min to ensure uni-form slurries, the panel was locked, graduated cylinders stood upright,and the sedimentation process started. Time-lapse photos were takenfor all the tested graduated cylinders with high resolution digitalcamera (Canon EOS 5d Mark II, Canon, Japan). The time intervals areas follows: 1 min interval in the first 1 h, 30min interval in the fol-lowing 19 h, 2 h interval in the following 11 days, 4 h interval in thefollowing 8 days, and then 24 h interval for another 15 days (total of 35days). No further changes were observed in all the graduated cylindersafter 35 days. The interface between supernatant and suspension, and

Table 1Properties of Geogia kaolinite used in this study.

Source Active Minerals International, Hunt Valley, MD, USA

Trade name ACTI-MIN RP-2Color Creama

Specific gravity 2.60a

d50 0.36 μma

Chemical composition SiO2 45.60%, Al2O3 38.40%, Fe2O3 0.88%, TiO2

1.69%, CaO 0.05%, MgO 0.02%, K2O 0.15%, Na2O0.21%, LOI 13.70%a

Cation exchangecapacity (CEC)

CEC=7.9 meq/100 gb

Isoelectric point (IEP) Face pH ≈ 4, edge pH ≈ 7.2c

Point of zero charge(PZC)

4.6d

a From ACTI-MIN RP-2 data sheet, Active Minerals International [62].b From Hazen Research Inc.c From Palomino and Santamarina [6].d From Stumm [64].

Table 2Chemical concentrations in kaolinite slurries.

Chemical Name NaCl(mol/l)

Xanthangum (g/l)

Chitosan (g/l) PAA (g/l) PAM (g/l)

Concentration 0.003 0.001 0.05 0.05 0.10.006 0.01 0.5 0.5 0.50.015 0.1 1 0.25 –0.15 0.5 5 1 –

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

17

Page 3: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

that between suspension and sediment (“mud line”) as a function oftime were read (accuracy about 0.5 ml) from the time-lapse photos.Then the mud line vs. time readings were plotted, and the initial linearportion of the curve was used to determine the initial settling rateα. Thefinal volume of the sediment was taken from the last time-lapse photoat time of 35 days.

After the final volume was reached, approximate 0.5ml slurry wereextracted by transfer pipettes (Fisherbrand, graduated 3ml, Cat No. 13-711-9AM, Fisher scientific) and pipettor(D64362, Fisherbrand,200–1000 μl) at selected elevations of the graduated cylinder and

transferred into 2.5ml macro disposable cuvettes (759071D, BRANDGMBH+CO KG, Fisher scientific) for zeta potential or grain size dis-tribution measurement. The elevation selection for grain size distribu-tion is as follows. 1–2 elevations in supernatant were selected, with oneelevation near the top. 3–4 elevations were taken from the suspension,with one elevation slightly above the “mud line” if perceived. 2–3elevations were chosen from the sediment, with one elevation at thebottom. The preliminary results in this study indicated that zeta po-tential did not vary significantly with elevation. Therefore, the eleva-tion selection for zeta potential was reduced to one elevation at each ofthe three locations: supernatant, suspension, and sediment. Averageresults were calculated for further discussion. Care was taken to mini-mize the disturbance by low extraction flowrate.

Zeta potential was measured with Zetasizer Nano ZS90 zeta po-tential analyzer (Malvern Instruments, UK) and folded capillary cell(DTS 1061, Malvern Instruments, UK). High-accuracy pipette (D64362,Fisherbrand, 200–1000 μl) was used to transfer 500 μl samples into thefolded capillary cell. Two cycles were used to take zeta potential values,

Fig. 1. Molecular structures of biopolymers.

Table 3Molecular weight (M.W.) of polymers used in this study.

Polymer Name Xanthan gum Chitosan PAA PAM

M. W. (g/mol) 0.9× 105-1.6× 105

1× 105-3× 105

∼1× 106 ∼ 2×105

Fig. 2. Sedimentation panel.

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

18

Page 4: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

and average values were reported. Repeat tests were performed untilthe last 3 readings were within +/- 10mV. Before each measurementthe folded capillary cell was first rinsed with ethanol (Fisher Scientific)then with deionized water 3 times to prevent cross-contamination. Thetest temperature was set at 23.5 °C. The Zetasizer was calibrated priorto measurement by using the standard calibrating solution (Duke3520 A, Hi-Q Nano sensor calibrant). Some samples in the middle sec-tion and all samples in the bottom section exceeded the solid contentmeasurement limit of ZetaSizer. Therefore these samples were diluted100 times in 2.5 ml macro disposable cuvette with deionized water withthe exception of NaCl-kaolinite mixture, where NaCl solution of theoriginal concentration was used for dilution. This is because sodiumwas not as strongly adsorbed on the kaolinite surface as the other 4polymers used in this study [8,21].

The grain size distribution of kaolinite particles were measured bydynamic light scattering technique with Malvern Zetasizer Nano ZS90and 10mm lightpath, 1.5ml capacity semimicro style polystyrenecuvettes (Cat No. 14-955-127, Fisherbrand, Fisher scientific). All NaCl/polymer-kaolinite mixtures were preconditioned to 25 °C before testing.High accuracy pipette(D64362, Fisherbrand, 200–1000 μl) was used totransfer 500 μl samples into 2.5ml macro disposable cuvettes. Similarto the zeta potential test, high solid concentrations can exceed themeasurement limit of ZetaSizer. So samples from suspensions and se-diments were diluted 10 times (20 μl mixture: 200 μl deionized water/NaCl solutions) before the particle size distribution measurement.20 μlmixture was transferred by using high accuracy pipette (5 μl,Eppendorf Research) with the range from 10 μl to 100 μl. Dilution wasperformed in 1.5ml disposable plastic cuvettes with deionized water forpolymer-kaolinite mixtures, or with NaCl original concentration solu-tions for NaCl-kaolinite mixtures. It was observed that clusters of kao-linite aggregates taken from the sediments and some suspensions wouldsettle to the bottom of the cuvettes during dilution, which would not bedetected by Zetasizer. In order to avoid this problem, the cuvette wasgently rotated immediately before the test to agitate the kaolinite

clusters into the entire aqueous space. Care was taken not to destroy thecluster structure. Some samples were re-tested after a few minutes toconfirm/improve the results. At least one repeat test with a freshsample was performed at each elevation. Cumulative density functionof lognormal distribution model was selected to fit the particle sizedistribution curves of kaolinite. The mean values and standard devia-tions of kaolinite particle size were taken into account and calculatedfor the later discussion.

The solid content profile along the elevations was measured afterthe final volume was reached. The supernatant was first siphoned outwith a pipet at 10–20ml intervals, followed by the suspension and se-diment. Care was taken not to disturb the suspension and the sedimentby extracting from the surface and by low extraction flowrate. Thevolume intervals were separated at the interface (mud line) between thesuspension and the sediment. Each volume interval was moved to asmall beaker (25ml), which was then weighed (Ohaus EX224, FisherScientific), oven-dried (104 °C), and weighed again for the water con-tent measurement. Then the solid content at each volume interval wasreported in terms of the solid weight percentage: weight of solid/totalweight. It was noted that some solid particles were attached to the sideof the graduated cylinder. However, over 90% of the solid particleswere extracted into the beaker as oven-dried solid mass indicate, whichwarranted the representativeness of the results.

The intensities of kaolinite particles in each polymer along theelevation were plotted by using Matlab. Firstly, the rgb2gray functionwas taken to convert the true color image to the grayscale intensityimage. After choosing the coordinates of the elevation of each cylinderusing the getpts function, the intensity data on the grayscale imagealong each elevation can then be directly read and plotted by the plotfunction.

3. Results

The final volumes of kaolinite in NaCl, xanthan gum, chitosan, PAA

Fig. 3. Final volumes of kaolinite inNaCl and polymers solutions.

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

19

Page 5: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

and PAM solutions were shown in Fig. 3. The zeta potential curves ofkaolinite in different concentrations of NaCl/polymersolutions areplotted in Fig. 4. The profiles of the mean particle size and solid con-tents of kaolinite along the elevations in the five solutions were pre-sented in Figs. 5 and 6, respectively. And Fig. 7 demonstrated the in-tensity curves of kaolinite along the elevations as time progressed in0.15mol/lNaCl and 0.05 g/l PAA solutions. The intensity curves in0.003mol/lNaCl, 0.001 and 0.5 g/l xanthan gum, 1 g/l PAA, 0.1 and0.5 g/l PAM and 0.05 and 5 g/l chitosan solutions were very similarwith the curve in 0.05 g/l PAA solution. The sedimentation trends ofkaolinite particles in different concentrations of NaCl and polymerssolutions were exhibited in Fig. 8 and the settling velocities,α (ml/min)of kaolinite in NaCl and PAA solutions were calculated and plotted inFig. 9. From these figures, the following observations can be made:

3.1. NaCl

(1) The final volume and mean particle size of kaolinite increasedwith the increment of NaCl concentration (Figs. 3 and 5a). (2) The solidcontents decreased as the concentration of NaCl increased, especiallyunder the elevation of 15ml (Fig. 6a). (3) A clear boundary wasfoundon 0.15mol/l NaCl intensity curve as time went on (Fig. 7a). Thisagreed with the final volume result (41.5 ml) shown in Fig. 3. (4) Thesettling velocity decreased with the increment of NaCl concentration(Fig.9).

3.2. Xanthan gum

(1) The final volume and profiles of both the particle size and solidcontent for kaolinite in 0.001 g/l xanthan gum solution (Figs. 5b and6b) were very close to those for kaolinite in 0.003mol/l NaCl solution(both at around 15ml) (Figs. 3, 5a and 6a). This suggests that thexanthan gum dosage of 0.001 g/l (electrical conductivity= 2.441 u s/cm) is probably insignificant to change the fabric structures of kaoliniteparticles (fresh washed kaolinite suspension: electrical con-ductivity= 328.7 u s/cm). (2) The final volume of kaolinite in 0.5 g/lxanthan gum solution (21.5 ml) was higher than that in 0.001 g/l so-lution (15ml) (Fig. 3). (3) Under the mudline, the kaolinite particlesizes were larger in 0.5 g/l solution than those in 0.001 g/l solution atthe same elevation (Fig. 5b).

3.3. Chitosan

(1) The mean size of kaolinite particles ranged from 100 to 600 nmalong the elevation in both chitosan solutions, and (2) the particle sizewas influenced slightly by the two chitosan concentrations, except forthose at the elevation of 18ml in 5 g/l solution, where the maximumparticle size is around 900 nm (Fig. 5c). (3) The mudlines of final vo-lume of kaolinite were around 14ml in 0.05 g/l and 20ml in 5 g/lchitosan solutions (Fig. 3).

Fig. 4. Zeta potential curves in NaCl/polymers solutions.

Fig. 5. The profiles of the mean particle size of kaolinite along the elevation in different concentrations of NaCl and polymers solutions.

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

20

Page 6: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

3.4. PAA

(1) Profiles of both the mean particle size and solid content ofkaolinite did not vary too much in all different concentrations of PAAsolutions, and (2) most of the kaolinite particles were smaller than600 nm, but few particles are larger than 900 nm in 0.05 and 1 g/l so-lutions (Figs. 5d and 6d). (3) The final volume increased with the

increment of PAA concentration (Table 4). (4) The settling velocitydecreased as the PAA concentration increased (Fig. 9).

3.5. PAM

(1) Both the profiles of the mean particle size and solid content forkaolinite along the elevation in 0.1 and 0.5 g/l polyacrylamide (PAM)

Fig. 6. The profiles of solid contents of kaolinite along the elevation in different concentrations of NaCl and polymers solutions. (Dashed lines represent the mudlinesbetween the sediments and suspensions).

Fig. 7. The intensity curves of kaolinite along the elevation as time progressed in 0.15M NaCland 0.05 g/l PAA solutions.

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

21

Page 7: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

solutions were similar to each other, if the standard deviations of theparticle size were taken into consideration (Figs. 5e and 6e). (2) All theparticle sizes detected were smaller than 900 nm in both PAM solutions.(3) The final volume of kaolinite in 0.1 g/l solution was around 20ml,which was close to that in 0.5 g/l solution at around 21ml (Fig. 3). (4)The sedimentation volumes shown on sedimentation trends (Fig. 8c)decreased firstly and then increased until the values became stable inboth PAM solutions.

4. Discussion

4.1. NaCl

Based on the fabric map developed by Palomino and Santamarina[6], kaolinite particles were primarily deflocculated and dispersed inlow concentration of NaCl solution (< 0.1–0.15mol/l) at pH=5–7.2.With the increment of NaCl concentration, most of them formed edge toface (EF) fabric structures due to the decrement of electronic doublelayer thickness and the repulsive force. Further increasing the con-centration of NaCl, until it reached the threshold concentration (≈0.1-0.15mol/l), kaolinite particles converged into edge to edge (EE) fabricstructure. This conclusion agreed with the observations made in thisstudy. Kaolinite particles were mainly dispersed in 0.003mol/l NaClsolution. Thus, mainly small particles (< 500 nm) were detected. EFand EE fabrics were formed in 0.015 and 0.15mol/l solutions respec-tively. Therefore, large particles (> 900 nm) were measured, especiallythose with the size larger than 1500 nm in 0.15mol/l NaCl solution.Due to the formation of fabric structures, final volumes in 0.015 (30ml)and 0.15mol/l (41ml) solutions are larger than that in 0.003mol/lsolution (15ml).

4.2. Xanthan gum

The dissociation of carboxyl functional groups (−COOH) along thexanthan gum polymeric chain (pKa value of 3.1 [48]) releases protonsin aqueous solution (Fig. 1a): R−COOH → R−COO− + H+ [49]. Thisreaction has two effects on kaolinite suspensions in xanthan gum so-lutions: (1) the negative charges from −COO− caused the zeta poten-tial of the solution to become more negative and the values droppedfrom −38.7 down to −45.2 mV with the increment of xanthan gumconcentration from 0.001 to 0.5 g/l (Fig. 4). In addition, those negativecharges repulsed the faces of kaolinite particles which were also ne-gative charged in pH≈ 4–7 solutions [6]. Therefore, the solutions weredominated by the steric repulsive force, which hindered the bridging

Fig. 8. The sedimentation trends of kaolinite particles.

Fig. 9. Settling velocities of kaolinite in different concentrations of NaCl andPAA solutions.

Table 4Final volumes of kaolinite in PAA solutions.

PAA concentration (g/l) 0.05 0.25 0.5 1Final volume (ml) 25 and 34 36 53 75

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

22

Page 8: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

effects between kaolinite particles and xanthan gum polymers. The finalvolume increased (Fig. 3, 21.5 ml>15ml) and the solid contentslightly decreased (Fig. 6b) in 0.5 g/l xanthan gum solution due to thehigher repulsive force compared to those in 0.001 g/l solution [50]. Theparticle size should not increase with the increment of xanthan gumconcentrations from 0.001 to 0.5 g/l, which disagreed with the testresults shown in Fig. 5b. However, this could be explained by the H+

effect which was discussed as follows. (2) The released H+ reduced thepH of the suspension, which likely renders more positive charges to theedges of kaolinite particles [1,6]. Those positive charges attracted thenegative charged xanthan gum molecule via electrostatic force. Thus, itwas postulated that kaolinite particles attached to the long xanthangum polymer chain exhibited a pattern similar to edge to edge (EE)fabric structure, resulting in the large particles (> 900 nm) in 0.5 g/lxanthan gum solution under the mudline of 21.5ml (Fig. 5b). Abovepostulation was validated by the SEM image (Fig. 6b) where the EEfabric structure was observed.

4.3. Chitosan

The primary interactive mechanisms between chitosan and kaoliniteare charge neutralization and interparticle bridging effect [51]. Firstly,the amine functional groups (-NH2) along the chitosan chain (Fig. 1b)(85% deacetylated) attract protons from water molecule in aqueoussolution, which makes the chitosan chain positively charged. The po-sitive chitosan chains are attracted to the negative faces of kaoliniteparticles by electrostatic force at pH=7.59–8.49, which slightly in-creases the net surface charge of kaolinite (less negative). Zeta poten-tial, an indicator of net surface charge of a solid, of kaolinite suspen-sions increases mildly (from -41 to −40mV) with the increment ofchitosan concentration (from 0.05 to 5 g/l) (Fig. 4). This is probablydue to the small dosage of chitosan used in this study (mass ratio ofkaolinite and 5 g/l chitosan=21:1). The analogous profiles of kaoliniteparticle sizes along the elevation in chitosan and 0.003mol/l NaCl so-lutions suggests that the interactive mechanism between chitosan andkaolinite is probably the same with NaCl and kaolinite. In NaCl solu-tion, it is charge neutralization. Secondly, most of chitosan chains ad-sorb onto kaolinite particle faces in patches (Fig. 10a) at the testedconcentrations in this study due to the high charge density of kaolinite(CEC=7.9meq/100 g, Hazen Research Inc.) and short distances ofdeacetylated sites on chitosan chain [30,37,51]. Under these con-centrations, the interparticle polymer bridging effects (Fig. 10b) play asecondary role to the electrostatic force [37,51,52]. Therefore, most ofsmall particles (< 600 nm) were measured in both chitosan solutions(Fig. 5c). It is postulated that large particles (800 nm–900 nm) detectedin 5 g/l solution are possibly due to the interparticle bridging effect.And the bridging effect probably also caused a higher final volume(Fig. 3) and slightly smaller solid contents (Fig. 6c) in 5 g/l solutioncompared to those in 0.05 g/l solution (20ml> 14ml). Overall, thesedimentation results suggest that charge neutralization dominates theinteraction between kaolinite and chitosan solutions.

4.4. PAA

The primary interactive mechanisms between PAA polymer andkaolinite particle are charge neutralization and bridging effect: (1) PAAis negative charged in aqueous solution due to the dissociation of car-boxyl functional groups (−COOH) along the polymer chain (Fig. 1c,pKa=4.5). The octahedral alumina side of kaolinite face is positivecharged at pH ≈ 4 [6]. Therefore, negative PAA polymers mainly ad-sorb on the positive alumina side of kaolinite face and edge by chargeneutralization in acidic solution [53–55]. This adsorption renders morenegative surface charges, as confirmed by the zeta potential decrementfrom -39.1 to -46.5 mV with the increment of PAA concentration from0.05 to 1 g/l (Fig. 4). (2) PAA polymers can interact with kaoliniteparticles via hydrogen bonding between the hydroxylated alumina side

of kaolinite face and carboxyl groups of PAA polymer [54] or inter-particle bridging effect due to the partially adsorption of PAA polymerson the alumina side of kaolinite face [38,53].

The distinct bridging effect based on previous study [38] occurredin the solution with small PAA concentration (< 0.001 g/l) and largemolecular weight (≥ 2.5×105 g/mol). The bridging effect is wea-kened with the increment of PAA concentration, even though the mo-lecular weight is larger than 2.5×105 g/mol. This is probably becauseof the high repulsive forces from charge neutralization between kaoli-nite particles [56]. When the PAA concentration is larger than∼0.01 g/l, as is the case in this study, the PAA polymer chain primarilycrowds in a coiled form on alumina side of kaolinite surface (Fig. 10c)in acidic condition [32,56,57]. This coiled form hinders the bridgingeffect between PAA and kaolinite, leaving charge neutralization to bethe dominant mechanism. Therefore, small particles (< 600 nm) weremostly detected in both PAA solutions (Fig. 5d). It should be pointedout that few kaolinite particles could still flocculate because of limitedbridging effect as few large particles (≥900 nm) were measured in bothPAA solutions (Fig. 5d). The kaolinite particles tend to disperse in thePAA solution with high concentration (low settling velocity, Fig. 9) dueto the increment of repulsive force, which is mostly generated by theadsorption of PAA polymers on the kaolinite surface (Fig. 10c). Thus,the final volume in 1 g/l PAA solution (75ml) is larger than that in0.5 g/l solution (25ml) (Table 2), whereas the solid content, in turn, issmaller in 1 g/l solution because of the reduction of density of solids.And also the settling velocity decreased with the increment of PAAconcentrations. It is also noticed that solid contents increase slightlyalong the elevation in each PAA solution (Fig. 6d). This is attributed togravity force and the subsequent consolidation of kaolinite particles[58,59].

4.5. PAM

The high carboxyl content PAM used in this study has two func-tional groups: carboxyl and amide (Fig. 1d). Carboxyl groups (eCOOH)hydrolyze into negative carboxylic ions (eCOO−) and positive hy-drogen ions (H+). The former gives negative charges to the polymerchain, while the latter generates acidic condition to the aqueous solu-tion. When pH < 7.2, the edges of kaolinite particles become posi-tively charged [6], and attract the negatively charged PAM chainsthrough electrostatic interaction. Meanwhile, the faces of kaoliniteparticles are still negatively charged and repel the PAM chains. On theother hand, another functional group on the chain, the amide functionalgroup (eCONH2) can be adsorbed on both the silanol and aluminoleOH groups on both the faces and edges of kaolinite particles by hy-drogen bonding [21,60,61]. Because of the poorly hydrated silica faceand because of that hydrogen ions are relative tightly held on thealuminate faces, the amide groups are primarily hydrogen bonded tothe broken edges of kaolinite platelets, which are composed of exposedsilica and alumina [60,63]. Therefore, it can be inferred that highcarboxyl content PAM polymer chain mainly adsorbs to the edge ofkaolinite particles to form a fabric structure similar to edge to edge (EE)(Fig. 10d) by charge neutralization between carboxyl group and theedge of kaolinite particle and hydrogen bonding between the carbonylof amide group and the free silanol at the edge of kaolinite particle.

This adsorption via charge neutralization increases with the incre-ment of PAM concentration, but by bridging effects would be reduced[21]. It is because of the overwhelming repulsive force produced by thenegative PAM polymer chains which are adsorbed onto two differentedges of particle (Fig. 10e). The whole solution becomes steric stabili-zation. Nabzer et al. [63] found that 0.95×10−3 g/l PAM is thethreshold for 2 wt% kaolinite particles changing from EE bridging tosteric stabilization, which is equivalent to≈ 0.05 g/l PAM in this study.Hence, in 0.1 and 0.5 g/l PAM solutions, the interaction mechanismbetween PAM molecule and kaolinite particles is probably steric sta-bilization. This may be the reason that most particles detected in both

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

23

Page 9: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

solutions are under 900 nm (Fig. 5e). Due to the same interaction me-chanism, the final volumes (∼ 20ml in 0.1 g/l and ∼ 21ml in 0.5 g/l,Fig.3) and solid contents curves (Fig. 6e) are also similar in both PAMsolutions.

5. Conclusions

This study examined the interaction mechanisms between polymerand kaolinite particles. Sedimentation tests on kaolinite in NaCl andbiopolymers (xanthan gum, chitosan, polyacrylic acid and poly-acrylamide) solutions were conducted and recorded with time-lapsedtechnique up to 35 days. A comprehensive set of data, including particlesize, zeta potential and solid content along the elevation of the grad-uated cylinders, final volume and settling velocity were obtained. Majorobservations are summarized as follows:

(1) Coulombic interaction dominates the behavior of kaolinite-NaClsuspensions by compressing the electrical double layer surroundingboth the faces and edges of kaolinite particles, and forming eitheredge-to-face (EF) kaolinite fabric in 0.015mol/l NaCl solution or

edge-to-edge (EE) kaolinite fabric in 0.15mol/l NaCl solution, bothof which result in larger particle size (> 900 nm) and higher finalvolumes than a lower NaCl concentration (0.005mol/l) does.

(2) Steric repulsive force between −COO− on xanthan gum chain andthe negatively-charged sites on kaolinite face hinders the bridgingeffect between xanthan gum and kaolinite particles in 0.001 and0.5 g/l xanthan gum solutions. On the other hand, protons releasedfrom the −COOH functional groups of xanthan gum chains accu-mulated on the kaolinite edges, which enhanced the attractionsbetween positively-charged kaolinite edges and negatively-chargedchain and exhibited layered structures with edge-to-edge connec-tions between particles. As a result, large kaolinite particles(> 900 nm) were presented in 0.5 g/l xanthan gum solution.

(3) The primary interactive mechanism between kaolinite and chitosanis charge neutralization. Most positively-charged chitosan chainsadsorbed onto kaolinite faces in patches at 0.05 and 5 g/l solutions,leaving the interparticle bridging effect as a secondary mechanism.As a result, most small particles (< 600 nm) were presented in 0.05and 5 g/l chitosan solutions. Some large particles (800 nm–900 nm)were also detected in 5 g/l solution, which is possibly due to the

Fig. 10. The schemata of interactive mode of kaolinite particles with (a) (b) chitosan [30], (c) PAA (this study) and (d) (e) PAM [63].

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

24

Page 10: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

bridging effect between chitosan chains and kaolinite particles. Thisbridging effect is also believed to cause a higher final volume (20mlvs. 14ml) and slightly smaller solid contents in 5 g/l solution thanin 0.05 g/l solution.

(4) Small kaolinite particles (< 600 nm) were primarily detected inPAA solutions, which is postulated to be due to the adsorption ofthe coiled form of positively-charged PAA chains on the aluminaside of kaolinite faces. This coiled form hinders the bridging effectbetween PAA chains and kaolinite particles, promotes the inter-particle repulsion, and leaves charge neutralization to be thedominant interaction mechanism. As PAA concentration increased,charge neutralization and the subsequent interparticle repulsiveforces lead to increment of final volumes, decrement of solid con-tents in sediment region, and decrement of settling velocity.

(5) In both 0.1 and 0.5 g/l PAM solutions, the interactive mechanismbetween PAM molecule and kaolinite particles is considered tobesteric stabilization in the literature. Because of the interparticlerepulsive force produced by the negative PAM chains adsorbed onadjacent kaolinite edges, the particle size is small (< 600 nm).Stericstabilization is short ranged, as a result, no large aggregates wereformed, the final volume is low (20ml in 0.1 g/l and 21ml in 0.5 g/l solutions), and the solid content is high in the sediment region.

Acknowledgements

This work is sponsored by the National Natural Science Foundationof China (Award No.: 51779219). The financial support by the One-Thousand-Young-Talents Program of the Organization Department ofthe CPC Central Committee to both the second author and the corre-sponding author are deeply appreciated.

References

[1] G. Zhang, H. Yin, Z. Lei, A.H. Reed, Y. Furukawa, Effects of exopolymers on particlesize distributions of suspended cohesive sediments, J. Geophys. Res.: Oceans 118(2013) 3473–3489.

[2] M. Han, D.F. Lawler, The (relative) insignificance of G in flocculation, J. AWWA 84(1992) 79–91.

[3] D.N. Thomas, S.J. Judd, N. Fawcett, Flocculation modelling: a review, Water Res.33 (1999) 1579–1592.

[4] J.C. Santamarina, K.A. Klein, M.A. Fam (Eds.), Soils and Waves, John Wiley andSons Ltd, New York, 2001, pp. 67–71.

[5] J.N. Israelachvili (Ed.), Intermolecular and Surface Forces, third edition, Elsevier,Amsterdam, 2011.

[6] A.M. Palomino, J.C. Santamarina, Fabric map for kaolinite: effects of pH and ionicconcentration on behavior, Clays Clay Miner. 53 (2005) 211–223.

[7] A. Zaman, Effects of polymer bridging and electrostatics on the rheological behaviorof aqueous colloidal dispersions, Part. Part. Syst. Charact. 20 (2003) 342–350.

[8] B. Bate, S.E. Burns, Effect of total organic carbon content and structure on theelectrokinetic behavior of organoclay suspensions, J. Colloid Interface Sci. 343(2010) 58–64.

[9] R.J. Hunter, R.H. Ottewill, R.L. Rowell (Eds.), Zeta Potential in Colloid Science,Academic Press, London, 1981.

[10] E. Pefferkorn, Polyacrylamide at solid/liquid interfaces, J. Colloid Interface Sci. 216(1999) 197–220.

[11] D.O. Shah, R.S. Schechter (Eds.), Improved Oil Recovery by Surfactant and PolymerFlooding, Academic Press, San Diego, 1977.

[12] K.S. Sorbie (Ed.), Polymer-Improved Oil Recovery, CRC Press, Boca Taton, FL,1991.

[13] T. Stutzmann, B. Siffert, Contribution to the adsorption mechanism of acetamideand polyacrylamide onto clays, Clays Clay Miner. 25 (1977) 392–406.

[14] R.D. Lentz, I. Shainberg, R.E. Sojka, D.L. Carter, Preventing irrigation furrow ero-sion with small application of polymers, Soil Sci. Soc. Am. J. 56 (1992) 1926–1932.

[15] R.D. Lentz, R.E. Sojka, Applying polymers to irrigation water: evaluating strategiesfor furrow erosion control, Trans. ASAE 43 (2000) 1561–1568.

[16] R.A. Nugent, G. Zhang, R.P. Gambrell, The effects of exopolymers on the erosionalresistance of cohesive sediments, ICSE (2010) 162–171.

[17] J. Yu, T. Lei, I. Shainberg, A.I. Mamedov, G.J. Levy, Infiltration and erosion in soilstreated with dry PAM and gypsum, Soil Sci. Soc. Am. J. 67 (2003) 630–636.

[18] X.C. Zhang, W.P. Miller, M.A. Nearing, L.D. Norton, Effects of surface treatment onsurface sealing, runoff, and interrill erosion, Trans. ASAE 41 (1998) 989–994.

[19] N. Beier, W. Wilson, A. Dunmola, D. Sego, Impact of flocculation-based dewateringon the shear strength of oil sands fine tailings, Can. Geotech. J. 50 (2013)1001–1007.

[20] A. McFarlane, K. Bremmell, J. Addai-Mensah, Improved dewatering behavior of

clay minerals dispersions via interfacial chemistry and particle inter-actionsoptimization, J. Colloid Interface Sci. 293 (2006) 116–127.

[21] P. Mpofu, A.J. Mensah, J. Ralston, Investigation of the effect of polymer structuretype on flocculation rheology and dewatering behavior of kaolinite dispersions, Int.J. Miner. Process. 71 (2003) 247–268.

[22] D.W. Hunter, The positive impact of polymers on sediment treatment and handling,Proceedings of the Western Dredging Association Twenty-Sixth TechnicalConference, (2006).

[23] R.H. Jones, R.R. Williams, T.K. Moore, Development and application of design andoperation procedures for coagulation of dredged material slurry and containmentarea effluent, Technical Report -U.S. Army Engineer Waterways Experiment Station,(1978) Vicksburg, MS.

[24] C. Wang, K.Y. Chen, Laboratory study of chemical coagulation as a means oftreatment for dredged material, Technical Report - U.S. Army Engineer WaterwaysExperiment Station, (1977) Vicksburg, MS.

[25] J.L. Cleasby, E.R. Baumann, A.H. Dharmarajah, G.L. Sindt, Design and operationguidelines for optimization of the high-rate filtration process: plant survey results,AWWA Res. Found. Rept. Bolton and Menk, Inc., Denver, Colorado, 1989.

[26] R.D. Letterman, R.W. Pero, Contaminants in polyelectrolytes used in water treat-ment, J. AWWA 82 (1990) 87–97.

[27] A. Ates, The effect of polymer-cement stabilization on the unconfined compressivestrength of liquefiable soils, Int. J. Polym. Sci. (2013).

[28] X. Kang, B. Bate, Shear wave velocity and its anisotropy of polymer modified Highvolume class F fly ash-kaolinite mixtures, J. Geotech. Geoenviron. Eng. 142 (2016)04016068.

[29] K. Newman, J. Tingle, Emulsion polymers for soil stabilization, U. S. Army EngineerResearch and Development Center, 2004 FAA Worldwide Airport TechnologyTransfer Conference, New Jersey, USA, 2004.

[30] F. Bergaya, B.K.G. Theng, G. Lagaly, Handbook of Clay Science. Developments inClay Science vol. 1, Elsevier, 2006, pp. 192–200 (chapter 5).

[31] V. Chaplain, M.L. Janex, F. Lafuma, C. Graillat, R. Audebert, Coupling betweenpolymer adsorption and colloidal particle aggregation, Colloid Polym. Sci. 273(1995) 984–993.

[32] B.K.G. Theng (Ed.), Formation and Properties of Clay Polymer Complexes, Elsevier,Amsterda, New York, 1979.

[33] F. Csempesz, S. Rohrsetzer, The effects of polymer bridging on the flocculationkinetics of colloidal dispersions, Colloids Surf. 31 (1988) 215–230.

[34] Y. Otsubo, Effect of surfactant adsorption on the polymer bridging and rheologicalproperties of suspensions, Langmuir 10 (1994) 1018–1022.

[35] R.I. Feigin, D.H. Napper, Depletion stabilization and depletion flocculation, J.Colloid Interface Sci. 75 (1980) 525–541.

[36] L.T. Lee, R. Rahbari, J. Lecourtier, G. Chauveteau, Adsorption of polyacrylamideson the different faces of kaolinites, J. Colloid Interface Sci. 147 (1991) 351–357.

[37] second edition, B.K.G. Theng (Ed.), Formation and Properties of Clay-PolymerComplexes, vol. 4, Elsevier Science, 2012.

[38] K.K. Das, P. Somasundaran, Flocculation-dispersion characteristics of alumina usinga wide molecular weight range of polyacrylic acids, Colloids Surf. A: Physicochem.Eng. Asp. 223 (2003) 17–25.

[39] S. Kim, A.M. Palomino, Polyacrylamide-treated kaolin: a fabric study, Appl. ClaySci. 45 (2009) 270–279.

[40] V. Bertolino, G. Cavallaro, G. Lazzara, S. Milioto, F. Parisi, Biopolymer-targetedadsorption onto halloysite nanotubes in aqueous media, Langmuir 33 (13) (2017)3317–3323.

[41] R.D. Miller, D.R. Crandell, Time-lapse motion picture technique applied to the studyof geologic processes, Sci. AAAS 130 (1959) 795–796.

[42] J.A. Hayward, J.H. Barton, Erosion by frost-heaving: a time-lapse photographicrecord, Soil Water 6 (1969) 3–5.

[43] N. Matsuoka, Combining time-lapse photography and multisensory data logging tomonitor slope dynamics in the southern Japanese Alps, Japan Geoscience UnionMeeting, (2014) 2014.

[44] J.G. Shellberg, A.P. Brooks, C.W. Rose, Sediment production and yield from analluvial gully in northern Queensland, Australia, Earth Surf. Process. Landforms 38(2013) 1765–1778.

[45] D.R. Crandell, D.J. Varnes, Movement of the slumgullion earthflow near Lake City,Colorado, Short Papers in the Geologic and Hydrologic Sciences: U.S. GeologicalSurvey Professional Paper 424 (1961), pp. 136–139.

[46] G. Cavallaro, G. Lazzara, S. Milioto, Exploiting the colloidal stability and solubili-zation ability of clay nanotubes/ionic surfactant hybrid nanomaterials, J. Phys.Chem. C 116 (41) (2012) 21932–21938.

[47] M. Zbik, R.S.C. Smart, Influence of dry grinding on talc and kaolinite morphology:inhibition of nano-bubble formation and improved dispersion, Miner. Eng. 18(2005) 969–976.

[48] A. Oprea, M. Nistor, L. Profire, M.I. Popa, C.E. Lupusoru, C. Vasile, Evaluation of thecontrolled release ability of theophylline from xanthan/chondroitin sulfate hydro-gels, JBNB 4 (2013) 123–131.

[49] K.M. Dontsova, J.M. Bigham, Anionic polysaccharide sorption by clay minerals, SoilSci. Soc. Am. J. 69 (2005) 1026–1035.

[50] G. Cavallaro, G. Lazzara, S. Milioto, G. Palmisano, F. Parisi, Halloysite nanotubewith fluorinated lumen: Non-foaming nanocontainer for storage and controlledrelease of oxygen in aqueous media, J. Colloid Interface Sci. 417 (2014) (2013)66–71.

[51] J. Li, S. Jiao, L. Zhong, J. Pan, Q. Ma, Optimizing coagulation and flocculationprocess for kaolinite suspension with chitosan, Colloids Surf. A: Physicochem. Eng.Asp. 428 (2013) 100–110.

[52] L. Chen, D. Chen, C. Wu, A new approach for the flocculation mechanism of chit-osan, J. Polym. Environ. 11 (2003) 87–92.

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

25

Page 11: Colloids and Surfaces A - Missouri S&Tweb.mst.edu/~bateba/Bate_Js/Bate2018CSA.pdf · 2018. 6. 15. · Colloids and Surfaces A journal homepage: ... Introduction Grain size distribution

[53] Z. Pan, A. Campbell, P. Somasundaran, Polyacrylic acid adsorption and con-formation in concentrated alumina suspensions, Colloids Surf. A: Physicochem. Eng.Asp. 191 (2001) 71–78.

[54] D. Santhiya, S. Subramanian, K.A. Natarajan, S.G. Malghan, Surface chemical stu-dies on the competitive adsorption of poly(acrylic acid) and poly(vinyl alcohol)onto alumina, J. Colloid Interface Sci. 216 (1999) 143–153.

[55] A.A. Zaman, R. Tsuchiya, B.M. Moudgil, Adsorption of a low-molecular-weightpolyacrylic acid on silica, alumina, and kaolin, J. Colloid Interface Sci. 256 (2002)73–78.

[56] K.K. Das, P. Somasundaran, Ultra-low dosage flocculation of alumina using poly-acrylic acid, Colloids Surf. A: Physicochem. Eng. Asp. 182 (2001) 25–33.

[57] K.F. Tjipangandjara, P. Somasundaran, Effects of the conformation of polyacrylicacid on the dispersion-flocculation of alumina and kaolinite fines, Adv. PowderTechnol. 3 (1992) 119–127.

[58] R.E. Gibson, R.L. Schiffman, K.W. Cargill, The theory of one-dimensional

consolidation of saturated clays.II. Finite nonlinear consolidation of thick homo-geneous layers, Can. Geotech. J. 18 (1981) 280–293.

[59] H. Pu, P.J. Fox, Y. Liu, Model for large strain consolidation under constant rate ofstrain, Int. J. Number. Anal. Met. 37 (2013) 1574–1590.

[60] L. Nabzar, A. Carroy, E. Pefferkorn, Formation and properties of the kaolinite-polyacrylamide complex in aqueous media, Soil Sci. 141 (1986) 113–119.

[61] E. Pefferkorn, L. Nabzar, A. Carroy, Adsorption of polyacrylamide to Na kaolinite:correlation between clay structure and surface properties, J. Colloid Interface Sci.106 (1985) 94–103.

[62] Active Minerals International, Data Sheet for ACTI-MIN RP-2 Kaolin Clay, ActiveMinerals International, Hunt Valley, MD, 2007.

[63] L. Nabzar, E. Pefferkorn, R. Varoqui, Stability of polymer-clay suspensions.Thepolyacrylamide-sodium kaolinite system, Colloids Surf. 30 (1988) 345–353.

[64] W. Stumm, Chemistry of the Solid-Water Interface: Processes at the Mineral-Waterand Particle-Water Interface in Natural Systems, Wiley, New York, 1992.

J. Cao et al. Colloids and Surfaces A 554 (2018) 16–26

26