Download as pdf or txt
Download as pdf or txt
You are on page 1of 74

   

The application of atomic force microscopy in mineral flotation

Yaowen Xing, Mengdi Xu, Xiahui Gui, Yijun Cao, Bent Babel, Martin
Rudolph, Stefan Weber, Michael Kappl, Hans-Jürgen Butt

PII: S0001-8686(17)30410-4
DOI: doi:10.1016/j.cis.2018.01.004
Reference: CIS 1858

To appear in: Advances in Colloid and Interface Science

Please cite this article as: Xing Yaowen, Xu Mengdi, Gui Xiahui, Cao Yijun, Babel Bent,
Rudolph Martin, Weber Stefan, Kappl Michael, Butt Hans-Jürgen, The application of
atomic force microscopy in mineral flotation, Advances in Colloid and Interface Science
(2018), doi:10.1016/j.cis.2018.01.004

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.
ACCEPTED MANUSCRIPT

The application of atomic force microscopy in mineral flotation


Yaowen Xing1,2, 3, Mengdi Xu1, Xiahui Gui2,3, Yijun Cao2,4, Bent Babel5, Martin Rudolph5,

Stefan Weber3, Michael Kappl3, Hans-Jürgen Butt3,*

T
1. School of Chemical Engineering and Technology, China University of Mining and Technology,

IP
Xuzhou 221116, China;

R
2. Chinese National Engineering Research Center of Coal Preparation and Purification, Xuzhou

SC
221116, China;

3. Max Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany;

NU
4. Henan Province Industrial Technology Research Institute of Resources and Materials, Zhengzhou

University, Zhengzhou 450001, China


MA
5. Helmholtz-Zentrum Dresden-Rossendorf, Helmholtz Institute Freiberg for Resource Technology,

Chemnitzer Str. 40, 09599 Freiberg, Germany.

*Corresponding author: butt@mpip-mainz.mpg.de (Butt H.); kappl@mpip-mainz.mpg.de (Kappl. M)


D

yijuncao@126.com (Cao Y.); guixiahui1985@163.com (Gui X.)


TE

Abstract
P

During the past years, atomic force microscopy (AFM) has matured to an indispensable tool
CE

to characterize nanomaterials in colloid and interface science. For imaging, a sharp probe
AC

mounted near to the end of a cantilever scans over the sample surface providing a high

resolution three-dimensional topographic image. In addition, the AFM tip can be used as a

force sensor to detect local properties like adhesion, stiffness, charge etc. After the invention

of the colloidal probe technique it has also become a major method to measure surface forces.

In this review, we highlight the advances in the application of AFM in the field of mineral

flotation, such as mineral morphology imaging, water at mineral surface, reagent adsorption,

inter-particle force, and bubble-particle interaction. In the coming years, the complementary

characterization of chemical composition such as using infrared spectroscopy and Raman

spectroscopy for AFM topography imaging and the synchronous measurement of the force
1
ACCEPTED MANUSCRIPT

and distance involving deformable bubble as a force sensor will further assist the fundamental

understanding of flotation mechanism.

Keywords: atomic force microscopy, mineral flotation, surface imaging, inter-particle force,

T
IP
bubble-particle interaction

R
Contents

SC
1. Introduction ............................................................................................................................ 3
2. Imaging of minerals ............................................................................................................... 8
3 Water at mineral surface ........................................................................................................ 12

NU
3.1 Water structure near the interface ................................................................................... 12
3.2 Surface nanobubbles characterization ............................................................................ 15
MA
3 Reagent adsorption on mineral surface ................................................................................. 18
3.1 Structure of adsorbed reagent ......................................................................................... 18
3.2 Single molecule force spectroscopy ............................................................................... 20
D

4 Quantification of inter-particle force ..................................................................................... 24


TE

5 Bubble-particle interaction and thin liquid film drainage ..................................................... 32


6 Other applications ................................................................................................................. 40
P

7. Conclusions and perspectives............................................................................................... 44


CE

Acknowledgments .................................................................................................................... 46
Symbols .................................................................................................................................... 47
AC

References ................................................................................................................................ 47

2
ACCEPTED MANUSCRIPT

1. Introduction

Froth flotation, belonging to the family of heterocoagulation separation techniques, is based

on the difference in surface hydrophobicity of dispersed particles. It has been widely used in

T
IP
mineral processing [1-4], fine coal upgrading [5-7], wastewater treatment [8, 9], oil sands

R
processing [10, 11], fly ash decarburization [12-14] and pulp-deinking [15, 16]. For mineral

SC
flotation, when the raw minerals are collected from the underground, crushing and grinding is

required for liberating the valuable components from the interlocking particles. Then, these

NU
fine particles are mixed with water and conditioned with appropriate reagents (collectors,
MA
frothers, depressants and regulators). Air is finally introduced and hydrophobic particles are

captured by the rising bubbles while hydrophilic particles remain in the pulp.
D

The essential sub-processes in flotation are: wetting, reagents adsorption on minerals,


TE

inter-particle force, and bubble-particle interaction. Each step plays a critical role in the final
P

flotation recovery. A comprehensive understanding of these sub-processes is of great


CE

fundamental and practical importance to design high-efficiency flotation process. Up to now,


AC

a number of studies have been carried out to investigate these sub-processes [17-28]. Contact

angle measurement is the most commonly used method in the study of mineral-water wetting

[29]. Spectroscopic techniques such as infrared spectroscopy, X-ray photoelectron

spectroscopy and sum-frequency generation spectroscopy (SFG) are usually applied to better

understand reagent adsorption on mineral surface and water structure at mineral-water

interface [17-21]. Zeta potentials are measured to characterize the charge on particles which is

essential to predict inter-particle electrostatic interaction. Theory and computational fluid

dynamics are reported to study hydrodynamic inter-particle or bubble-particle interaction

[22-28]. However, the flotation mechanisms at the nanoscale are still not well understood due
3
ACCEPTED MANUSCRIPT

to the difficulty in experimental verification.

Particularly, surface and interfacial forces (van der Waals, electrical double layer, hydrophobic

force, hydration, hydrodynamic, and adhesion forces) are of great importance to understand

T
IP
inter-particle or bubble-particle interaction in flotation cells [1, 30]. According to the classical

R
Derjaguin-Landau-Verwey-Overbeek (DLVO) theory, the stability of a colloidal system is

SC
controlled by two components [30, 31], namely van der Waals dispersion and electrostatic

double layer forces. The van der Waals force between different bodies depends on their

NU
differences in optical properties and can be characterized by the Hamaker constant. For the
MA
interaction of a solid/liquid with a liquid/gas interface, the van der Waals force is repulsive; it

favors thickening of the liquid film. For the electrostatic double layer force, surface charges
D

and the salt concentration are the key parameters for determining its value.
TE

In addition, non-DLVO forces such as hydration and hydrophobic forces are present in
P

flotation system. A short-ranged repulsive hydration force emerges when two hydrophilic
CE

surfaces are brought into contact and the structure of water is required to change upon
AC

approach [32-34]. In contrast, between two hydrophobic surfaces attractive forces are

observed [35-38]. Hydrophobic forces are usually the driving force for the attachment of solid

particles to bubbles. Laskowski and Kitchener [39] first found that the water films present on

methylated silica surfaces were unstable and the film ruptured spontaneously. Their

experiment indicates that in addition to the traditional DLVO forces, an attractive force also

acts between the bubbles and the hydrophobic solid particles. Three years later, Blake and

Kitchener [40] experimentally measured the thickness of the water film when a small gas

bubble was slowly advanced towards a polished silica plate with different hydrophobicity.

They showed that the film rupture thickness on methylated silica ranged from 60 to 220 nm.
4
ACCEPTED MANUSCRIPT

In this case, the thin water film on the methylated silica surface was unstable due to a

long-range attractive force. The first direct evidence for long-range attractive hydrophobic

forces between solid surfaces was provided by Pashley and Israelachvili using the Surface

T
IP
Force Apparatus (SFA) [35, 36]. In the SFA, the force between two hydrophobized mica

R
surfaces was measured. To date, attractive hydrophobic forces have been accepted as the

SC
reason for bubble-particle attachment, although its origin remains under debate. Measuring

the hydrophobic force between a bubble and a particle versus distance is, however,

NU
challenging. Due to its strong attractive characteristics and the softness of the bubble surface
MA
usually an instability accors once a certain distance has been reached [41, 42].

With the invention of Atomic force microscopy (AFM) a new tool became available for
D

measuring the interaction between particles and bubbles and for imaging the surface structure
TE

of minerals. (Fig. 1). AFM was first developed by Binnig et al. [43] in 1986, for imaging the
P

surface topography at nanometer resolution [44]. A sharp probe mounted near the end of a
CE

cantilever is raster scanned along the sample surface. A laser beam was used to monitor the
AC

deflection of the cantilever via a position sensitive diode detector (PSD). Cantilever deflection

is proportional to the force acting on the tip. Usually a feedback loop is used to keep the force

between tip and sample constant by moving the sample up and down during scanning. In that

case the height of the sample is plotted versus the lateral position to obtain a

three-dimensional topographic image both in air or liquid condition [45-47]. Typically, three

common imaging modes, i.e., contact mode, tapping mode, and non-contact mode, are

available in commercial AFMs based on the way in which the tip scans over the sample [47,

48]. In contact mode, the tip remains in contact with the sample surface by applying a

constant load between tip and the sample (Fig. 1. b). Due to the large contact pressure, the tip
5
ACCEPTED MANUSCRIPT

or the sample may get damaged during the imaging, leading to the decrease in resolution [48,

49]. For the soft surface, contact mode imaging will underestimate the height of the sample

due to the deformation effect. For tapping mode, the tip is oscillated with a value close to its

T
IP
resonance frequency. The tip comes into and out of contact with the sample repeatedly at the

R
lower turning point. This intermittent contact between tip and sample leads to a damping of

SC
the tip oscillation, reflected both by a reduction in amplitude and a shift of the resonance

frequency. Depending on the feedback used during intermittent mode, one discriminates

NU
between amplitude modulation (AM) and frequency modulation (FM). In the more common
MA
AM mode, the tip is approached to the sample until a preset amplitude reduction of cantilever

oscillation is reached, and this reduced amplitude is kept constant during scanning using a
D

feedback loop. In FM mode, a preselected shift of the resonance frequency is used for the
TE

feedback during approach and scanning. FM mode was originally developed for AFM in
P

ultra-high vacuum (UHV), where AM mode suffers from a slow response time of the
CE

feedback due to a much higher quality factor of the oscillating cantilever, as damping by a
AC

fluid is missing [50]. Recently it was demonstrated, that FM mode can also be used in liquids

[51], allowing imaging and force measurements with atomic resolution. Compared with

contact mode, intermittent contact mode allows to image soft surface with a higher resolution,

as interaction between tip and sample is minimized and especially lateral shear is eliminated.

a) b)

6
ACCEPTED MANUSCRIPT

Fig. 1 a), The working schematic of an AFM [30]; b), The force regimes governing AFM

imaging. Reprinted with permission from [48], Copyright 2004, Elsevier

Soon after its introduction, AFM was used as a force sensor to measure forces, including

T
IP
surface forces, adhesion force, and hydrodynamic force, between different surfaces or

R
molecules of interest, especially with the aid of the so called “colloidal probe technique” [47,

SC
52-54]. The force is obtained by multiplying the measured probe deflection by the spring

constant of the cantilever. In the colloid probe technique, a micrometer-size particle is

NU
attached to the end of the cantilever and replaces the sharp tip. Usually the particle is spherical
MA
to facilitate the quantitative analysis. Thus, the interaction geometry between the probe and

surface is greatly simplified.


D

AFM has also been widely used by flotation scientists to imaging mineral morphology and
TE

measure the force between single particles. However, it is difficult to predict flotation
P

recovery directly from force measurements because a large number of particles are involved
CE

in flotation practice. Still, force measurements can provide useful input parameters for
AC

computational fluid dynamics simulation to design high efficiency flotation process.

The aim of this review is to demonstrate the versatility, flexibility, and future potential of

AFM in order to shed new light on the flotation mechanisms. We highlight the current state of

the art in the application of AFM in the field of mineral flotation, such as mineral morphology

imaging, water at mineral surface, reagent adsorption, inter-particle force, and bubble-particle

interaction. Limitations and future perspectives are also discussed. This review is presented in

a way that is accessible and valuable to flotation researchers with basic knowledge

background regarding AFM.

7
ACCEPTED MANUSCRIPT

2. Imaging of minerals

AFM was used to image pure mineral surfaces such as mica, galena, molybdenite and

phyllosilicate. All can be easily cleaved [55-61], to prepare clean, atomically flat surfaces. For

T
IP
example, Gupta et al. [59, 62] used AFM for imaging the crystal lattice of silica and alumina

R
faces of phyllosilicate (kaolinite). Negatively charged glass or mica and positively charged

SC
alumina were used as the substrates for kaolinite deposition. Due to the electrostatic

double-layer attraction, the positively charged alumina face of kaolinite particles attached to

NU
the glass or mica surface, thus exposing the silica faces. In contrast, the alumina face of
MA
kaolinite particles was exposed, when alumina was used as a substrate. It was found that

tetrahedral oxygen atoms on the silica face formed a closed hexagonal ring-like network with
D

a vacancy in the centre. On the alumina face, the hexagonal lattice ring of hydroxyls
TE

surrounded a hydroxyl. The atomic spacing between neighboring hydroxyls was determined
P

as 0.36 ± 0.04 nm. Siretanu et al. [61] used AM-AFM to image the basal (001) planes of
CE

various types of phyllosilicates, namely gibbsite, kaolinite, illite, and Na-montmorillonite, in


AC

water (Fig. 2). All images show the characteristic hexagonal structure with a periodicity ≈0.5

nm that arises from the quasi-hexagonal arrangement of silica-tetrahedra and

alumina-octahedra, the classical building units of all clay materials.

8
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA
Fig. 2 Atomic resolution AFM images of clays basal plane: (a) gibbsite, (b)

Na-montmorillonite, (c) illite and (d–f) kaolinite clay particles obtained with super sharp tip at
D

room temperature equilibrated in ultrapure water. Scale bars, 2 nm. Top insets represent
TE

zooms of atomic scale images after processing with Fourier transformation superimposed with
P

X-ray crystal structure of clays along the ab plane. Reprinted with permission from [61],
CE

Copyright 2016, The Royal Society of Chemistry


AC

In sulfide mineral flotation the oxidation state is of great importance. The floatability of

sulfide minerals is directly related to the degree of its surface oxidation [63]. For example,

fresh galena is hydrophilic. However, it was found that slight surface oxidation increased the

hydrophobicity and floatability of galena due to the formation of sulfur rich layers [60, 64].

Therefore, collectorless flotation of sulfide minerals becomes possible by adjusting the

surface oxidation. Xie et al. [57] studied the effect of applied electrochemical potential, with

regard to the Ag/AgCl/3.4 M KCl reference electrode, on the galena surface morphology at

pH 5.6. It was found that surface roughness increased significantly when the potential

9
ACCEPTED MANUSCRIPT

exceeded +300 mV (Fig. 3). This was due to the agglomeration of electrochemical oxidation

products on the galena surface. Hampton et al. [60] studied the effect of applied potential on

electrochemical oxidation of the galena surface at pH 4.5 (Fig. 4). Sulphur domains were

T
IP
observed and surface roughness increased when the electrochemical potential increased to

R
+258 mV. . Hampton et al. [60] concluded that the electrochemical oxidation of galena is a

SC
two-step process: First, sulphur is released into the solution and then the sulphur deposits on

favourable sites; the sulphur domains are distributed heterogeneously (Fig. 4), while in Xie’s

NU
report [57], the sulphur domains were more homogeneously distributed. The reason for this
MA
discrepancy may be due to the increased surface roughness of galena in Hampton’s

experiment. Surface heterogeneity led to a heterogeneous distribution of oxidation products.


D
P TE
CE
AC

Fig. 3 Surface topographies of galena surfaces under different applied potentials in 0.5 M

NaCl: (A) −700 mV, (B) −300 mV, (C) 0 mV, (D) 300 mV, and (E) 450 mV. Adapted with

permission from [57], Copyright 2016, American Chemical Society

A B

10
ACCEPTED MANUSCRIPT

T
IP
C

R
D

SC
NU
MA
Fig. 4 Surface topographies of galena surfaces under different applied potentials: A, +8 mV; B,
D

+258 mV; C, +308 mV; D, +358 mV. Adapted with permission from [60], Copyright 2011,
TE

American Chemical Society


P

For non-pure minerals which do not have a cleavage face, pre-treatment such as careful
CE

polishing is needed to lower the surface roughness so that it can be imaged by AFM. Bruening
AC

and Cohen [65] applied AFM to identify the surface roughness variations before and after coal

oxidation. Morga [66] used AFM to study the effect of heat treatment on the surface

roughness of semifusinite and fusinite. Heating increased the surface roughness of both

semifusinite and fusinite.

AFM also has the ability to characterize the pore structure on coal surfaces [67]. Dun et al.

[68] found that magmatic intrusions have a great impact on the pore structures of coal. The

pore size of low-rank bituminous coal was much larger than high-rank anthracite. As a result,

the small molecular weight alkanes (collectors) are easily lost in the pores due to the capillary

effect [6]. Indeed, a higher concentration of collector is usually required for low-rank coal

11
ACCEPTED MANUSCRIPT

flotation than that of high-rank anthracite.

Both the sharpness and the aspect ratio of the probe influence images obtained with an AFM

due to the convolution and broadening effect. Only probes with a high aspect ratio and small

T
IP
probe radius are suitable to quantitatively characterize the roughness and pore structure on

R
mineral surfaces.

SC
3 Water at mineral surface

NU
3.1 Water structure near the interface
MA
AFM is capable of imaging mineral surfaces in liquid. In flotation, both hydrophilic and

hydrophobic particles are in contact with water. The interaction of the water molecules with
D

the surface atoms of the solid has an influence on the structure of the interfacial water.
TE

Knowing the water structure at the solid-water interface is essential to understand the flotation
P

mechanisms. Currently, the interfacial water structure is attracting great interest from
CE

researchers in the colloid and flotation area, and numerous studies have focused on it [69-76].

Water molecules close to a hydrophilic surface have a strong attractive interaction with
AC

surface atoms, e.g. via hydrogen bonds, increasing the surface free energy. This interaction

also leads to a structuring of the interfacial water, where water molecules on average reside

longer at positions with strong interaction than at positions with lower interaction force. Such

hydration layers typically consist of 3-5 water molecular layers [69-78].

Hydration layers are dynamic structures i.e., the water molecules in hydration film are always

movable. The water molecules of the first hydration layer are typically completely exchanged

within nanoseconds [79]. At the hydrophobic solid-liquid interfaces, in contrast, a water

exclusion zone with several angstroms was observed using high-resolution X-ray reflectivity

12
ACCEPTED MANUSCRIPT

and neutron reflectivity [80-82].

To map hydration structures with the tip of AFM is challenging because the tip-sample forces

induced by the presence of the hydration layers can be of the same order as the thermal

T
IP
fluctuations in standard AFM cantilevers. In 2005, Fukuma et al. [83] demonstrated true

R
atomic resolution imaging on a mica surface in aqueous buffer solution by using an AFM with

SC
very low detector noise. The low noise in the system was important because to obtain atomic

scale resolution, the amplitude of the cantilever oscillation had to be on the order of the size

NU
of a water molecule, i.e. less than one nanometer. Using small amplitude force distance
MA
spectroscopy, they also observed oscillations in the tip-sample force which they ascribed to

the layering of water molecules in the tip-sample gap. This layering is consistent with earlier
D

reports by Israelachvili and Pashley [32] and Zachariah et al. [84]. Israelachvili and Pashley
TE

[32] found that the short-range hydration force between two mica surfaces in water showed an
P

oscillatory profile with a mean periodicity of the water molecule diameter indicating the
CE

existence of ordered water layers. In contrast, Zachariah et al. [84] suggested that the hydrated
AC

ion layering is responsible for the oscillatory hydration force. However, information on the

potential lateral ordering of water molecules at the interface was still missing. Thus, a couple

of years later Fukuma extended his high-resolution AFM to enable three dimensional mapping

of the tip-sample force [85]. In this three-dimensional scanning force microscopy (3D-SFM,

Fig. 5), the tip is scanned up and down like the needle of a sewing machine while the lateral

position is slowly changed in a scanning motion, which enables to visualize 3D distribution of

water at a mica water interface in a region of 2 x 2 µm in 53 sec with an atomic-scale

resolution. In the first 3D-SFM investigation on a mica surface in aqueous buffer solution, a

Si cantilever with a resonance frequency of 123 kHz, a Q factor of 5.8, and 14.2 N/m spring
13
ACCEPTED MANUSCRIPT

constant in liquid was used. Once complete 3D force field was obtained, any 1D profile or 2D

cross sections could be extracted. A hexagonal pattern of force peaks with the same

periodicity as the mica substrate was observed. Above this first hydration layer, a second and

T
IP
a third layer of force peaks was observed with a lateral shift. Later, similar hydration

R
structures were observed on hydrophilic surfaces like calcite, fluorite, and dolomite [73, 78,

SC
86]. Most authors so far have used for 3D SFM in frequency modulation, mostly because the

force reconstruction from the observables frequency and phase shift, and oscillation and drive

NU
amplitude are well established [87, 88]. Recently, reliable force reconstruction methods for
MA
amplitude modulation AFM have also become available [89].

Although 3D-SFM imaging is always carried out on an atomically flat interface, the necessity
D

of this technique becomes more evident when it is applied to the systems having a larger
TE

corrugation and inhomogeneity. Recently, several groups have investigated theoretically by


P

molecular dynamics simulations how the measured force field is connected to the overlap of
CE

hydration layers of tip and surface [79, 90, 91]. It was found that the force maxima correspond
AC

to the undisturbed positions of water molecules at the surface and that the presence of the tip

has only a minor influence on the measured water structure. Further improvements in terms of

the measurement technique could be achieved by using small cantilevers combining high

resonance frequency and low spring constant [92, 93], which provide a better signal-to-noise

ratio. Furthermore, Söngen et al. have improved the 3D scanning method by introducing a tip

protection mechanism that retracts the tip when a threshold force is exceeded at the lower

turning point, providing more stable imaging [94].

14
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
Fig. 5 Schematics of (a) 2D-SFM and (b) 3D-SFM. (c) Experimental setup for the developed
MA
3D-SFM [85]. A phase-locked loop (PLL) circuit is used for the resonance frequency shift
D

detection while an automatic gain control (AGC) circuit is used for keeping the amplitude of
TE

the cantilever oscillation constant. The inset shows a 3D force image obtained at a mica-water

interface (2×2×0.78 nm3). Adapted with permission from [85], Copyright 2010, American
P
CE

Physical Society

3.2 Surface nanobubbles characterization


AC

When hydrophobic mineral surfaces are submerged in water, nanobubbles or air are easily

formed when the dissolved gas becomes oversaturated for example by increasing temperature

or mixing different liquids. Such nanobubbles can be stable for many hours [95, 96].

Nanobubbles at solid-water interface are an important subject since studies show that an

enhanced flotation recovery can be obtained by using nanobubbles [97, 98]. Parker et al. [99]

first suggested that the discontinuities in AFM force curves between two hydrophobic

surfaces were due to the presence of nanobubbles. Experimentally, surface nanobubbles were

firstly imaged by using AFM tapping mode by Ishida et al. [100] and Lou et al. [101].

15
ACCEPTED MANUSCRIPT

Subsequently, force mapping mode AFM [102-104], peak-force tapping mode AFM 1

[105-107], and peak-force quantitative nano-mechanics mode AFM2 [108] were all reported

as being capable to image nanobubbles (Fig. 6). In parallel, non-invasive techniques have

T
IP
confirmed the existence of nanobubbles such as infrared spectroscopy [109, 110], total

R
internal reflection fluorescence microscopy [111], cryo-scanning electron microscopy [112,

SC
113], quartz crystal microbalance [114], time-resolved fluorescence microscopy [115],

scanning synchrotron based scanning transmission soft X-ray microscopy [116] and small

NU
angle X-ray scattering [117]. The height of nanobubbles has been reported in the range of
MA
10-100 nm, while the length of the three-phase contact line is in general between 50 and 500

nm [118]. It is anticipated that nanobubbles should rapidly dissolve into surrounding water
D

due to a high internal Laplace pressure. However, surprisingly long lifetimes of nanobubbles
TE

have been observed experimentally [95, 96]. However, it is out of the scope of this paper to
P

review the stability theory of nanobubbles; an excellent recent review is [95].


CE

a b
AC

1
In Peak-force tapping the cantilever is periodically moved up and down. In contrast to
normal tapping mode, the cantilever is oscillated below resonance frequency
2
In Peak-force quantitative nano-mechanics mode, a series of force-distance curves is
recorded. While keeping the peak force constant, a several properties such as modulus,
adhesion force, and deformation depth can be extracted and quantified from the force-distance
curve at each pixel.
16
ACCEPTED MANUSCRIPT

c d

T
R IP
Fig. 6 Nanobubbles pictures imaged using different AFM working modes: (a) tapping mode,

SC
Reprinted with permission from [101], Copyright 2000, American Vacuum Society; (b) force

NU
mapping mode, Reprinted with permission from [102], Copyright 2013, American Chemical

Society; (c) Peak-force tapping mode, Reprinted with permission from [106], Copyright 2016,
MA
American Chemical Society; (d) Peak-force QNM mode. Reprinted with permission from
D

[108], Copyright 2013, The Royal Society of Chemistry


TE

Using AFM, the effects of production methods, substrate properties, salts, solution pH,
P

dissolved gas, surfactants, and temperature on nanobubbles can be studied. Previous studies
CE

showed that degassing pre-treatment reduces nanobubble formation [119, 120]. The gas type
AC

was also found to have great impact on nanobubble nucleation [121]. Zhang et al. [103, 122]

found that pH, salts, and sodium dodecyl sulfate has little effect on nanobubbles stability, thus

refuting the stability mechanism of the contamination skin hypothesis. Xu et al. [123]

observed that the formation of surface nanobubbles was temperature-dependent. More

nanobubbles were observed at high temperature than that at low temperature due to the lower

solubility of gas saturation at high temperature.

17
ACCEPTED MANUSCRIPT

3 Reagent adsorption on mineral surface

3.1 Structure of adsorbed reagent

T
The interaction between flotation reagents and mineral surfaces are of particular interest. In

IP
flotation practice, various kinds of flotation reagents, i.e., collectors, frothers, dispersants, and

R
flocculants, are used to regulate interfacial properties and thus the interaction between bubble

SC
and particle or inter-particle. Here, AFM can help analyzing the adsorption conformation of

NU
surfactants or polymers on mineral surface [56, 124-131]. A molecular layer or agglomerate

can be identified from AFM imaging. For example, the adsorption of the surfactants sodium
MA
dodecyl sulphate (SDS) and hexadecyltrimethyl ammonium bromide (CTAB) at

graphite–water interfaces has been studied using AFM by Parachuri et al. [132]. For both
D
TE

surfactants with either positively or negatively charged head group, the AFM images show the

adsorbed surfactant structures as linear parallel hemi-cylindrical micelles when the


P

concentration is below the CMC concentration. Parachuri et al. [133] further studied the effect
CE

of cosurfactants (non-ionic 1-dodecanol and cationic dodecyl trimethyl ammonium bromide)


AC

on the SDS surface micelle structures. The orderliness of the SDS hemicylinders was replaced

by a herring-bone pattern for the SDS/C12OH system. However, in the case of the

SDS/C12TAB system, the structures were more similar to the SDS only system. Jaschke et al.

[134] used AFM imaged the linear aggregates of ionic surfactant aggregates at a gold surface

in aqueous solution. The orientation of these aggregates was determined either by monatomic

steps on the gold surface or by the gold lattice itself. Chennakesavulu et al. [128] visualized

the conformation of oleate collector on a fluorite crystal surface. Both monolayer and bilayer

structures were observed even at low oleate concentration of 10-7 M. Paiva et al. [129] studied

18
ACCEPTED MANUSCRIPT

the effect of calcium ions on the adsorption of potassium oleate onto apatite surfaces. They

found that calcium ions play a critical role in potassium oleate adsorption, since the

adsorption was completed by forming calcium dioleate agglomerates. Beaussart et al. [56]

T
IP
explored the effect of three kinds of dextrins, a regular wheat dextrin (TY), carboxymethyl

R
(CM) dextrin, and hydroxypropyl (HP) dextrin, on molybdenite flotation. Topographies of

SC
adsorbed dextrins on molybdenite surface were imaged using AFM (Fig. 7). The surface

coverages of the modified dextrins (CM and HP) were much higher than that of regular TY,

NU
leading to a lower contact angle and flotation recovery.
MA
D
P TE
CE
AC

Fig. 7 Topographies of adsorbed dextrins on molybdenite surface: (a) bare molybdenite, (b)

Dextrin TY, (c) CM Dextrin, (d) HP Dextrin. Reprinted with permission from [56], Copyright

2012, Elsevier

As common practice, guar gum is used as the depressant for molybdenite flotation. The effect

of guar gum on molybdenite flotation was studied by Xie at al. [135]. The presence of

polymer aggregates on molybdenite surfaces could be correlated with molybdenite flotation

results. The flotation recovery of molybdenite decreased sharply from 69% to 11% and 3%

19
ACCEPTED MANUSCRIPT

after guar gum adsorption at concentrations of 1 and 5 ppm. Mierczynska-Vasilev and Beattie

[136] studied the adsorption of three kinds of substituted carboxymethyl cellulose on talc and

chalcopyrite. The wettability depressions of three polymers on talc were always more

T
IP
effective than on chalcopyrite since more polymers adsorbed onto talc surface from AFM

R
images. More specifically, low carboxymethyl cellulose coverage on both minerals was

SC
observed for a high substitution of the carboxymethyls.

NU
3.2 Single molecule force spectroscopy

Polymers are often used as flocculants in ultra-fine particle flotation and filtration. Fine
MA
particles form large-size flocs under the bridging function of high-molecular-weight polymer.

Theoretically, hydrogen bonding, the hydrophobic interaction, electrostatic attraction, and


D
TE

chemical bonding are the possible driving forces for flocculation. In early times it was,

however, difficult to measure the force between flocculant and mineral particle. With the
P

introduction of single molecule force spectroscopy (SMFS) it became possible to measure


CE

inter- and intramolecular interaction forces in polymer and supramolecular system with pN
AC

resolution using AFM [137-140]. Typically in such experiments, a polymer bridges the AFM

probe and substrate. Then the bridge is stretched by slowly retracting the tip while measuring

the force (Fig. 8).

a b

20
ACCEPTED MANUSCRIPT

Fig. 8 (a) Schematic of SMFS; (b) Stretching force curves between a polyprotein molecule

and a copper surface. ∆𝐿𝑐 is the contour length increment. Reprinted with permission from

[138], Copyright 2012, Nature Publishing Group

T
IP
Measuring the force between polymer flocculant molecule and mineral surface is of great

R
importance to understand the selective adsorption mechanism of the flocculant and thus

SC
develop new high efficiency flocculants. Sun et al. [141] carried out SMFS experiments

NU
between a novel hybrid polymer Al(OH)3-polyacrylamide (Al-PAM) and a silica surface. In

this case, the Al(OH)3 served as the core connecting a number of PAM chains. It was found
MA
that the adhesion force between Al(OH)3 and silica was much higher than that between PAM

chain and silica due to electrostatic attraction. A partially hydrolyzed polyacrylamide (HPAM)
D
TE

was found to be beneficial to bitumen extraction not only improving bitumen recovery but

also accelerating fine solids settling in the tailings stream. Long et al. [142] measured the
P
CE

adhesion force between a partially hydrolyzed polyacrylamide (HPAM) molecule and

different kinds of surfaces (silica, mica, and bitumen). To obtain more representative results
AC

for single-molecule desorption, for each test condition, the measurement was performed at a

number of different surface locations, and 2 to 3 substrate-tip pairs were used. More than

1000 force curves were recorded for each test condition in Long’s report. In deionized water,

the adhesion forces were reported to be 40, 200, and 80 pN on silica, mica, and bitumen

surfaces, respectively (Fig. 9). When the process water from oil sand processing plant,

containing various kinds of ions, was used, the forces changed to 50, 100, and 40 pN,

respectively. The adsorption strength between HPAM and mica was much higher than that on

the bitumen surface. This indicated that a selective flocculation between clay particles, such

21
ACCEPTED MANUSCRIPT

as mica, can achieve when HPAM is added to bitumen-clay mixture suspensions.

T
R IP
SC
Fig. 9 Adhesion forces between HPAM and different kinds of surfaces in DI water and

commercial plant process water. Reprinted with permission from [142], Copyright 2006,

NU
American Chemical Society.
MA
Pensini et al. [143] directly coated carboxymethyl cellulose polymer on a fresh mica surface

and a silica/borosilicate/iron oxide particle was attached to a tip-less cantilever. They recorded
D
TE

retraction forces between particles and carboxymethyl cellulose coated mica. Different

rupture events could be identified from the shape of the force curves (Fig. 10). Compared to
P
CE

SMFS experiments, the bridge between the colloidal probe and mica surface is no longer

formed by a single molecule. Instead a number of molecules interact due to the large contact
AC

area. It was found that the adhesion force between iron oxide and carboxymethyl cellulose

coated mica in Milli-Q water was always higher than that between silica/borosilicate. Solution

pH and ions were found to have great impact on the adhesion force.
a) b)

22
ACCEPTED MANUSCRIPT

Fig. 10 a), Schematics of different rupture events during particle pull-off from carboxymethyl

cellulose coated mica. Each minimum indicates a detachment event occurs. The last minimum

is the adhesion force between one polymer molecule and colloid particle; b), AFM force

T
IP
curves between carboxymethyl cellulose and iron oxide particles in milli-Q water buffered

R
with NaHCO3 to pH = 8. Reprinted with permission from [143], Copyright 2013, Elsevier

SC
High molecular-weight flocculants can be attached to the tip and their interaction with a

NU
mineral surface is relatively easy to measure. It is, however, difficult to directly measure the

force between a single surfactant molecule (collector) and a mineral surface. Therefore, some
MA
substitutes were used to model collector-mineral interaction. Fa et al. [144, 145] prepared a

calcium dioleate micro-sphere to model the interaction between the oleate collector and
D
TE

calcite/fluorite. A 10 µm diameter calcium dioleate micro-sphere with a surface roughness Ra

of 36 nm for a given 10 µm2 surface area was used.. More specifically, strong long-range
P
CE

attractive forces in approach curves and adhesion forces in retraction force curves were

observed between calcium dioleate and fluorite; the force between calcium dioleate and
AC

calcite was much weaker. Indeed, such studies indicate that, in the presence of calcium,

calcium dioleate forms firstly in bulk solution and then adsorbs in the form of calcium

dioleate [146].

Xing et al. [147] used solid-state paraffin and stearic acid to represent conventional

hydrocarbon oil and fatty acid collector in fine coal flotation. A coal particle with diameter of

approximately 35 μm was attached to the end of a tip-less cantilever. The interactions between

paraffin/stearic acid and fresh/oxidized coal particles were measured directly using atomic

AFM colloidal probe technique. In this case, the paraffin substrate for AFM experiments was

23
ACCEPTED MANUSCRIPT

prepared using section-cutting while stearic acid substrate was prepared by the pellet method.

More specifically, a significant jump-into contact between oxidized coal particles and stearic

acid was observed due to the hydrogen bonding while a monotonous repulsive force between

T
IP
oxidized coal and paraffin. Consequently, a fatty acid collector is more suitable for oxidized

R
coal flotation than hydrocarbon oil.

SC
4 Quantification of inter-particle force

NU
Inter-particle interaction in flotation pulp is also an important issue affects the final separation

efficiency. For example, coating of micro-fine slimes on targeted mineral particles has a
MA
detrimental effect on flotation selectivity [148, 149]. These slimes follow bubble-particle

aggregates and enter the froth, lowering the concentrate grade. On the other hand, such slime
D
TE

coatings make targeted particles become hydrophilic and prevent bubbles or collectors from

adhering to particles, reducing flotation recovery. Selective flocculation was also widely used
P
CE

for ultra-fine particle flotation to increase bubble-particle collision efficiency or reduce

gangue minerals entrainment [150]. Measuring the driving force responsible for these
AC

inter-particle interactions is of both academic and practical importance in flotation

engineering.

The quantification of inter-particle force in colloid science is one of the most important

applications of AFM [54]. Early work mainly focused on studying surface forces and

verifying the correctness of the Derjaguin-Landau-Verwey-Overbeek (DLVO) theory. The

critical step in AFM force measurement is to update the sensitivity and calibrate the spring

constant of the cantilever. During the data analysis, how to convert force-displacement curve

to force-distance one is also a critical step. However, great attention should be paid to the

24
ACCEPTED MANUSCRIPT

surface roughness effect since it would influence the force results and thus the repeatability of

the experiments. Ducker et al. [52] used AFM to measure the force between a hydrophilic

silica sphere and a silica plane in presence of sodium chloride. The measured force at long

T
IP
range could be well predicted by the classical DLVO theory. However, a repulsive force at

R
short range was observed probably due to the strong hydration of the surfaces. As the DLVO

SC
theory treats the intervening medium as continuous, the individual properties of molecules

involved are not taken into consideration. Therefore, it cannot describe the hydration force

NU
that is required to remove the water molecules at the interface when the distance decreases to
MA
a few molecular diameters. Butt [53] also measured the force between a silicon nitride tip,

alumina, glass, and diamond particle and glass/mica surface in presence of different salt
D

concentrations. A repulsive hydration force with 3 nm decay length between silicon nitride tip
TE

and mica was observed at a high salt concentration (>3 M).


P

Another kind of non-DLVO force is the hydrophobic force. A vast amount of hydrophobic
CE

force results can be found in the literature. Pashley and Israelachvili found that the
AC

hydrophobic force is much stronger than the van der Waals attraction and a single exponential

function gave the best fit to the experimental data [35, 36]. Rabinovich and Yoon [151] found

that the hydrophobic force between a hydrophobic glass sphere and a silica plate, could be

described with both exponential and power laws.

Up to date, the origin of the hydrophobic force remains unclear. Both sample preparation

methods and experimental techniques have great impact on the result [37]. Nanobubble

bridging is one of the most representative mechanisms [152]. Hampton and Nguyen [153, 154]

used a capillary mathematical model to fit the force curves between a 1-octanol esterified

silica sphere and a silica plane, which supports nanobubble bridge hypothesis. Another
25
ACCEPTED MANUSCRIPT

possible mechanism for the long-range hydrophobic force is the spontaneous formation of

cavities. Cavitation due to metastability of the intervening liquid during approach or

separation was proposed by Christensen and Claesson [155]. However, detailed discussion

T
IP
about these origin mechanisms is beyond the scope of this review. Readers who are interested

R
in this topic can refer to the excellent reviews by Christenson and Claesson [37] and Meyer et

SC
al. [38].

In many cases it is problematic to quantitatively describe inter-particle forces because of

NU
irregular particle structure and surface heterogeneity of practical minerals. Still, the force
MA
curves can be qualitatively compared with varying the concentration of salt or pH and may

provide guidance for flotation engineering. Xing et al. [148, 149] studied the effect of calcium
D

ions on coal flotation in the presence of kaolinite clay. The interaction force between coal and
TE

kaolinite was measured using AFM (Fig. 11). No jump-in was observed in de-ionized water
P

and the repulsive force corresponded to that in the case of no kaolinite coating. A sudden
CE

jump-in was found when Ca2+ was added to a concentration of 3 mM. The repulsive
AC

electrostatic force was suppressed by excessive Ca2+ addition, and the attractive force began

to dominate the kaolinite-particle interaction. Therefore, the presence of Ca2+ in practical

flotation aggravates slime coating and deteriorates flotation selectivity.

The effect of clay types, i.e., kaolinite and montmorillonite on fine coal flotation was further

studied [156]. Only a short jump-into-contact was observed in de-ionized water for

coal-montmorillonite. This low weak attraction correlates with a lower flotation recovery and

worse selectivity for the coal-montmorillonite system compared with that of coal-kaolinite

system where a repulsive force dominated.

Shrimali et al. [157] prepared two kinds of hematite colloidal probes of around 15 µm in
26
ACCEPTED MANUSCRIPT

diameter. In one case, the hematite was glued in a way that the (001) crystal face was exposed,

while in the other case the hematite particle was oriented so that the (100) crystal surface was

exposed. AFM force measurements between the (001) hematite crystal surface and the above

T
IP
two colloidal probe were carried out in 0.001 M KCl solution with varying pH values. Strong

R
long-range attractive hydrophobic forces at pH 5.5 and pH 2.7 were observed, confirming that

SC
hematite is hydrophobic at these pH values. At alkaline pH values, only repulsive forces were

found indicating that surface hydroxylation takes place at alkaline pH making the hematite

surface hydrophilic.
NU
MA
D
P TE
CE

Fig.11 Force curves between coal substrate and a ~50-µm-diameter kaolinite particle in
AC

presence of different calcium concentrations: (a) approach curves; (b) retraction curves. Both

attractive force and adhesion force increased with increasing calcium concentration. Reprinted

with permission from [148], Copyright 2016, American Chemical Society.

Often it is more difficult to measure the inter-particle force in the presence of polymer due to

experimental difficulties in controlling the polymer dose and the often irreversible adsorption

[158, 159]. Nevertheless, Abraham et al. [160] measured the force between silica surfaces in

the presence of copolymers of acrylamide and negatively charged acrylic acid with three

different negative charge densities (i.e, 15%, 40%, and 70% acrylic acid fraction in

27
ACCEPTED MANUSCRIPT

copolymers). More specifically, at low charge density the force was repulsive in the range of

experimental polymer concentrations (from 0.1 ppm to 50 ppm). Increasing the charge density

to 40 % or 70%, an attractive adhesion force was observed at a low polymer concentration

T
IP
due to polymer bridging. A purely repulsive force was observed again when the concentration

R
increased to higher levels.

SC
Zhou et al. [161] also studied the effect of charge density (10, 40, and 100%) of cationic

polymers on the force between silica surfaces using AFM. The 10% charged polymer

NU
produced steric repulsion upon approach and long-range adhesion force at the optimum
MA
flocculation concentration. In contrast, the 40 and 100% charged polymers produced

attraction during the approach, due to the attractive electrostatic force, and strong adhesion
D

during the retraction stage at optimum polymer dosages. It was also found that the polymer
TE

dose that produced the optimum flocculation and the maximum compressive yield stress
P

typically corresponded to the polymer concentration that produced the maximum adhesion
CE

force in AFM force measurement for each polymer.


AC

Ofori et al. [162] examined the coal tailings flocculation performance using different types

and concentrations of flocculants. Averaged jump-out distances were statistically determined

from thousands of AFM force curves. The optimal flocculant concentration would be the

point corresponding to the maximum jump-out distance in AFM force curves. However,

flocculation and sedimentation tests showed optimum flocculant concentrations were much

higher than those corresponding to the optimum adhesive strength found in the AFM

measurements. This was probably due to the large surface area of particles in the concentrated

suspension used in the settling tests compared with just a few particles used in the atomic

force measurement. In concentrated suspension, so much polymer adsorbs to the particle


28
ACCEPTED MANUSCRIPT

surface that polymer is depleted. This would also lead to less polymer at the particle surfaces.

It should be noted that this effect is only important at low polymer concentrations.

Recently, a multi-particle colloidal probe technique (MPCPT) based on AFM-inverted optical

T
IP
microscope has been developed to measure inter-particle forces in situ [159, 163]. MPCPT

R
works in a colloidal suspension in presence of a number of particles, which simplifies the

SC
precise dosing of polymers and makes this approach less sensitive to impurities. Note that still

one particle was attached to the end of cantilever in MPCPT. In these experiments, silane

NU
reagents are used to hydrophobize a glass substrate and a tip-less AFM cantilever. A number
MA
of polystyrene latex particles are injected into the liquid cell. The particles deposit onto glass

surface. Then, a particle is picked up by moving the cantilever on its upside. Finally, force
D

measurement is conducted between this colloid probe and another deposited particle. The
TE

main advantage of MPCPT, compared with that of traditional colloid particle probe only with
P

one or two colloidal particles, is that the adsorption of polyelectrolyte on particles can be
CE

controlled precisely, due to the larger internal surface area in multi-particle suspension
AC

system.

Flotation is also used for oil sands separation engineering. Oil sand separation has been

studied, e.g. by Xu’s and his team from Edmonton [164]. In oil sand processing, the first step

is to liberate oil from the sand. The sand typically consists of carbonate rocks and silica. Air

bubbles are introduced to capture the heavy oil (bitumen) leaving sand particles in the pulp.

Clays account for 15-30 percent in raw oil sands [165]. Important sub-processes are:

bitumen-sand liberation, bitumen-bubble interaction, and bitumen-fine particle interaction

(clay coating). Here, we treat bitumen as the targeted particle similar to mineral flotation;

therefore, we also briefly discuss the application of AFM in bitumen-sand liberation and clay
29
ACCEPTED MANUSCRIPT

coating in this section.

Liu et al. [166] coated a thin bitumen film on a flat silica wafer with a spin coater and then

measured the force between bitumen and silica particle in 1 mM KCl solution. They found

T
IP
that both pH and calcium concentration play an important role in bitumen-silica interaction

R
(Fig. 12). Not only did the repulsive barrier increase, but also the adhesion force decreased

SC
with increasing pH or decreasing calcium concentration.

Temperature is an important parameter in oil sand processing. Increasing temperature reduces

NU
bitumen viscosity and thus decreases the adhesion force [167]. Alkaline solution with high
MA
temperature and low ion concentrations promotes bitumen liberation. Zhang et al. [168]

further suggested a synergistic effect between surfactants (dodecyltrimethylammonium


D

chloride and sodium dodecylbenzene sulfonate) and divalent cation ions (Ca2+ or Mg2+) for
TE

bitumen–silica interaction. More specifically, in alkaline solution, sodium dodecylbenzene


P

sulfonate does not adsorb on negatively charged silica and bitumen surfaces due to the strong
CE

repulsive double-layer force. However, this is not the case when divalent cation ions were
AC

added. Ca2+ or Mg2+ prefers adsorbing on silica and bitumen surface in the form of

CaOH+ and MgOH+. Thus, these monohydroxyl species act as bridges between sodium

dodecylbenzene sulfonate and the solid surface.

Liu et al. [169] measured the force between bitumen and different kinds of irregular clay

particles (montmorillonite and kaolinite). When using irregular particles as a colloidal probe a

quantitative comparison of measured forces with the standard DLVO theory of colloid

stability is difficult because the effectively interacting areas and the contact area are unknown.

Consequently, the force curves have to be only qualitatively compared. In Liu’s experiments

[169] the adhesion force between bitumen and montmorillonite was always higher than that
30
ACCEPTED MANUSCRIPT

between bitumen and kaolinite, especially in in the presence of Ca2+. This is probably due to

the large capacity of adsorption of calcium for montmorillonite, depressing the repulsive force

barriers and enhancing the adhesion force.

T
R IP
SC
NU
MA
Fig. 12 The effect of pH and calcium concentration on the interaction between bitumen

substrate and a ~5-10 µm silica particle in 1 mM KCl solution: (left) pH; (right) calcium
D

concentration. The dashed and solid lines represent the experimental and theoretical fitting
TE

results, respectively. Reprinted with permission from [166], Copyright 2005, Elsevier
P
CE

A related subject which should be briefly mentioned is bacteria-mineral interaction. A deep

understanding of bacteria-mineral interaction force is critical to mineral bioleaching


AC

engineering. To measure the force between individual bacteria and a mineral surface by AFM,

the main challenge is to anchor the bacteria onto the cantilever or flat substrate without

affecting cell activity [170-172].

Lower et al. [173] introduced a new technique named biologically-active-force probe (BAFP)

to measure the force between bacteria and different mineral surfaces (muscovite, goethite, and

graphite). They used poly-D-lysine to functionalize glass beads; the negatively charged glass

and cell-surfaces were bridged by positive charged poly-D-lysine in situ (Fig. 13). It was

found that ions, mineral surface charge, and hydrophobicity significantly affected the

31
ACCEPTED MANUSCRIPT

interaction. Diao et al. [174] compared adhesion forces between chalcopyrite and

acidithiobacillus thiooxidans/acidithiobacillus ferrooxidans under various pH conditions. The

average adhesion force for acidithiobacillus ferrooxidans was always stronger than that of

T
IP
acidithiobacillus thiooxidans, especially at low pH.

R
SC
NU
MA

Fig. 13 (a) Schematic of BAFP, (b) Scanning laser confocal image of BAFP. The green
D

particle is the cell coated glass. Scale bar is 10 µm. Reprinted with permission from [173],
TE

Copyright 2005, Elsevier


P
CE

5 Bubble-particle interaction and thin liquid film drainage

Bubble-particle interaction is the critical step for flotation. It is well accepted that the
AC

hydrophobic force is responsible for successful bubble-particle attachment [42]. AFM is a

unique tool to directly measure bubble-particle interaction. Note that AFM force measurement

between an air bubble and a solid particle is accompanied with the deformation of gas/liquid

interface under both the hydrodynamic and surface forces. The dynamic coupling between

force, bubble deformation and film drainage makes both theoretical analysis and experimental

verification challenging [41, 175].

Ducker et al. [176] and Butt [177] first used AFM to measure the force between a bubble and

a particle in 1994. A small hydrophilic or hydrophobic silica particle was attached to the end

32
ACCEPTED MANUSCRIPT

of AFM cantilever and then approached to the bubble surface at a low speed in pure water.

The observed attractive force between a hydrophilic particle and an air bubble in Ducker’s

experiment [176] probably due to contamination. When surfactant was added, the long range

T
IP
attractive force disappeared and a monotonous repulsive force emerged. In Butt’s report [177],

R
he found that a repulsive force acted between bubble and a hydrophilic glass. In contrast, a

SC
jump-into contact was observed when hydrophobic particles approached the bubble surface

and a three-phase contact was formed. In general, particles with finite contact angle towards

NU
water form a three-phase contact and attach to the air/water interface.
MA
The determination of zero distance is an essential step to convert AFM force-displacement

curve to force-distance curve [178-181]. For the interaction between solid surfaces, the zero
D

distance is deduced from constant compliance regime in a hard contact system. For the
TE

interaction of a particle with an air/water interface zero distance is a matter of definition. It is


P

convenient to set zero distance to the equilibrium position of a particle in the air/water
CE

interface. The bubble is assumed to behave as a Hookean spring under external force [180,
AC

182-186]. To illustrate the situation we consider a small spherical particle at a planar air/water

interface as a model. For particles much smaller than the capillary length     g

capillary forces dominate; gravity, buoyancy and inertia can usually be ignored. With typical

values for the density of the liquid 𝜌  1000 kg/m3, the acceleration of gravity g = 9.81 m/s

and a surface tension of  = 0.072 N/m, the capillary length for water in a gas is 2.7 mm.

“Planar” implies that the bubble radius is much larger than the particle size. For a sphere

being pulled out of an infinitely deep and extended liquid pool (Fig. 14), this force increases

linearly with the length of the capillary bridge [187-190] and can be approximated by [191]:

33
ACCEPTED MANUSCRIPT

2
F (1)
0.809  ln  R 

Eq. (1) has been derived on theoretical grounds. It agrees with experimental results, e.g. for

T
the interaction of a microparticle with a bubble [176, 181], for a microparticle moved against

IP
the surface of water from the water side [192, 193], and for particles being drawn out of a

R
liquid-air interface [194]. When the contact line is pinned, Eq. (1) is also valid, only R has to

SC
be replaced by Rsin, where  denotes the position of the pinned contact line (Fig. 14).

NU
MA
D
P TE

Fig. 14. Schematic of a small particle at a liquid surface in equilibrium (bottom) and when a
CE

force is pulling the particle upwards (top). The contact line, characterized by the angle , is
AC

assumed to slide over the particle surface at constant contact angle .

Ishida [183] studied the effect of surface hydrophobicity on bubble-particle interaction force.

With increasing contact angle, the range of the attractive force increased whereas the wetting

film rupture thickness did not changed significantly. Preuss and Butt [195, 196] studied the

influence of dodecyltrimethylammonium bromide (DTAB) and sodium dodecyl sulfate (SDS)

on bubble-silica interaction. When DTAB was added at concentrations below the CMC a

hydrophobic attraction was observed caused by the strong adsorption of trimethylammonium

head group to the silica. The exposed dodecyl group rendered the particles hydrophobic.

34
ACCEPTED MANUSCRIPT

When the DTAB concentration exceeded 6 mM, the repulsive force dominated bubble-silica

interaction again. Above 6 mM DTAB adsorbs as a double layer to silica exposing the

positively charged head group to the water phase. The air water interface is also positively

T
IP
charged due to adsorption of DTAB, leading to a strong electrostatic repulsion. In the

R
experimental concentration range of SDS, no attractive force was observed between

SC
hydrophilic silica and bubble. For a hydrophobic particle, hydrophobic force dominated the

total force. However, the range and magnitude of the attractive force decreased as SDS

NU
concentration increased and SDS adsorbs to the hydrophobic solid surface and the air/water
MA
interface.

Nguyen et al. [180-182] did a series of experiments on bubble-particle interaction using AFM.
D

For a hydrophilic silica particle, the forces at different electrolyte concentrations could be
TE

well fitted by using classical DLVO theory. With increasing ion concentration, the
P

electrostatic force decreased due to the compression of the double electrostatic layer. The
CE

effect of approach velocity on the force between a hydrophilic glass sphere and bubble was
AC

also studied [180]. The repulsive hydrodynamic force increased monotonically as the

approach speed increased. When piezo approach speeds were lower than 0.6 µm/s, the

hydrodynamic force was negligible and instead surface forces were dominant. When the

particle became hydrophobic, a sudden jump-into contact due to the attractive hydrophobic

force always observed at a low approach speed. The attractive hydrophobic force also

increased as water contact angle increased. Nguyen et al. [197] used AFM to determine the

contact angle of a micrometer-sized polyethylene sphere according to the depth of the particle

penetration into the bubble at the zero force. They observed that the contact angle measured

by AFM changes with the speed of the AFM piezo.


35
ACCEPTED MANUSCRIPT

Ishida [183] explored the particle hydrophobicity on the force between a silica particle and an

air bubble. It was found that the hydrophobicity of the particle did not significantly change the

film rupture thickness, whereas the pH of the solution played a critical role in critical film

T
IP
rupture thickness. Assemi et al. [181] used AFM to measure the force between a silica and an

R
air bubble in deionized water and KCl solutions with a low piezo speed. By fitting the force

SC
curves using DLVO theory, the surface potential on the bubble surface was obtained. The

force curves at the edge of the bubble could fit very well to theory, even at high applied forces

due to the less bubble deformation effect.


NU
MA
However, the simple assumption that the bubble surface behaves as a linear spring will not

always be appropriate [175, 198]. Chan [175] derived a no-linear force-displacement relation
D

from the solution of the Young-Laplace equation between a bubble probe and a solid plane, as
TE

shown in Eq. (2):


P

𝐹 𝐹 1 1+𝑐𝑜𝑠𝜃 1
∆𝐷 = 4𝜋𝛾 {log (8𝜋𝛾𝑅 ) + 2(1 + 2 log (1−𝑐𝑜𝑠𝜃) − 2+𝑐𝑜𝑠𝜃) − 1} (2)
CE

The additional logarithmic term leads to a weak deviation from a linear-distance relationship
AC

when the compression D is not negligible compared to the radius of the bubble.

In recent years, significant progress has been achieved on both hydrodynamic drainage

modelling and AFM-based experimental techniques. It has become possible to measure the

force, bubble deformation, and liquid film drainage simultaneously [41].

Chan et al. [175, 199] derived a model called augmented Stokes-Reynolds-Young-Laplace

model to predict the force and the evolution of the wetting film profile before the formation of

three-phase contact line in dynamic AFM experiments. The defining equations of the

augmented Stokes-Reynolds-Young-Laplace model under a no-slip boundary condition are

[175]:
36
ACCEPTED MANUSCRIPT

𝜕ℎ 1 𝜕 𝜕𝑝
= 12𝜇𝑟 𝜕𝑟 (𝑟ℎ3 𝜕𝑟 ) (3)
𝜕𝑡

2𝛾 𝛾 𝜕 𝜕ℎ
𝑝 = 𝑅 − 𝛱 − 𝑟 𝜕𝑟 (𝑟 𝜕𝑟 ) (4)
𝑏

Eqs. (3) and (4) can be solved numerically in the region 0≤ 𝑟 ≤ 𝑟𝑚𝑎𝑥 when h<<𝑅𝑏 . 𝑟𝑚𝑎𝑥 is

T
IP
selected as the value where the local separation ℎ is so large that the contribution of

R
disjoining pressure could be neglected. The Stokes-Reynolds-Young-Laplace model has the

SC
ability to capture the essential physical characteristics in AFM force measurements and

provides a basic understanding of bubble-particle interaction force and film drainage

dynamics.
NU
MA
Shi et al. [200] measured the force between an air bubble and mica hydrophobized with

octadecyltrichlorosilane (OTS) with different hydrophobicity in 500 mM NaCl solution. AFM


D

force curves were fitted by using the Stokes-Reynolds-Young-Laplace model. A jump-into


TE

contact at 7.5 nm was experimentally observed for a partially hydrophobized mica with 45°
P

water contact angle. The model reproduced the force results assuming a hydrophobic force
CE

with a decay length of 0.8 nm. Increasing the contact angle to 85°, both the critical film
AC

rupture thickness and hydrophobic force decay length increased, illustrating that the

hydrophobic force increased with water contact angle. Force measurements between air

bubbles and sphalerite and molybdenite surfaces of different water contact angle were also

conducted by Xie et al. [135, 201]. More specifically, the effect of polymer depressant (guar

gum) on the bubble-molybdenite attachment was studied. When 1 ppm guar gum was added,

polymer aggregates on the surface became visible from AFM images, resulting in a decreased

bubble-particle attachment. Further increasing guar gum concentration to 5 ppm, the contact

angle of molybdenite surface decreased from 74° to 65°. A monotonic repulsive force was

observed. The AFM force results were consistent with the practical molybdenite flotation
37
ACCEPTED MANUSCRIPT

results. The final flotation recovery decreased from 69% to 11% and to 3% after guar gum

adsorption at 1 and 5 ppm concentration solutions.

The other effective approach complementary to AFM force measurement is optical

T
IP
interferometry. Clark et al. [202-205] first described the use of evanescent wave scattering to

R
measure the separation between a solid surface and a particle that was attached to an AFM

SC
cantilever. The evanescent scattering apparatus was essentially identical to that used in the

original total internal reflection microscopy except that the light scattered back into the

NU
incident medium was collected. However, for bubble-particle interaction, the evanescent wave
MA
scattering is not suitable because the liquid has a higher refractive index than air. In 2015, Shi

et al. [206] first combined a reflection interference contrast microscope (RICM) with an AFM
D

to measure the force and the thin water film profile between a bubble and mica surface
TE

simultaneously (Fig. 15). A bubble was attached on the end of a cantilever. The dynamic
P

profiles of air-water interface were reconstructed using RICM with nanometer resolution in
CE

normal direction. It was found that the Stokes-Reynolds-Young-Laplace model can predict
AC

both the force and film drainage with a no-slip boundary condition. A stable wetting film was

observed between an air bubble and a hydrophilic mica surface. In contrast, jump-in events

where the interaction force drastically turned from repulsive to attractive were observed for

hydrophobized mica with both 45° and 90° contact angle (Fig. 16). Again, the critical film

rupture thickness increased as contact angle increased. Due to the short range of van der

Waals force and the suppression of the electrostatic force at 500 mM NaCl, an additional

attractive hydrophobic force was considered as the reason for triggering film rupture. From

the Stokes-Reynolds-Young-Laplace model, the effective hydrophobic decay length was

inferred to be 0.8 and 1.0 nm for the 45° and 90° water contact angle, respectively. These
38
ACCEPTED MANUSCRIPT

results are explained in terms of a different water structure at the solid-water interface with

different surface contact angle.

T
R IP
SC
NU
Fig. 15 Schematic of the AFM-RICM experimental setup. Adapted with permission from
MA
[206], Copyright 2015, American Chemical Society.
D
P TE
CE
AC

Fig. 16 Temporal evolution of the force (left), film profile (middle) and disjoining pressure

(right) during approach between an air bubble and hydrophobized mica surfaces (A-C: 45°

contact angle, D-E: 90° contact angle) in 500 mM NaCl solution. Open circles denote the

experimental results and solid curves represent the theoretical calculation results. Adapted

with permission from [206], Copyright 2015, American Chemical Society.

39
ACCEPTED MANUSCRIPT

Combining AFM with RICM is a landmark in our full understanding of bubble-particle

interaction [207]. In the next few years, synchronous measurement of the force and the

spatiotemporal evolution of thin water film profile between bubble and particle in presence of

T
IP
flotation reagents will shed new light on flotation mechanism [41].

R
6 Other applications

SC
In recent years, several studies aimed to determine the mineral surface charge and wetting

NU
characteristics using direct AFM force measurement [208]. Historically, a colloid probe has

usually been employed and the surface charge information was obtained by fitting the force
MA
curves with the DLVO theory. Ducker et al. [52] applied AFM to measure the force between

silica surfaces at different salt concentrations and pH values. The force curves were fitted
D
TE

using DLVO theory, where the surface potential of silica was used as the fitting parameter.

The similar procedure was also adopted to determine the surface potentials of mica/silica and
P
CE

the isoelectric point of fluorite by Hartley et al. [209] and Assemi et al. [210].

Another aspect is to measure the point of zero charge (PZC) of different faces of anisotropic
AC

layered silicates such as kaolinite, talc, pyrophyllite and illite. Traditional zeta potential

measurements by electrophoresis or streaming potential assume a uniform charge density on

these layered silicate particles. Often, however, macroscopic coagulation of dispersed

particles is not necessarily observed at the PZC. The reason for coagulation lies in the fact that

the surface charge is not distributed evenly and that on one particle regions with positive and

negative charge density may exist. However, the problem can be solved using AFM since with

the AFM local charge densities can be measured [211, 212]. Zhao et al. [213] used an

ultramicrotome cutting technique to prepare an edge surface of muscovite, while a smooth

40
ACCEPTED MANUSCRIPT

basal face was easily prepared by cleavage. A silica sphere with an 8 μm diameter was

attached to the cantilever and a series of force curves were obtained. The force profiles

between the silica and basal face were well fitted with classical DLVO theory. They found

T
IP
that pH has little effect on the surface potential of muscovite basal face. However, on the edge

R
they could not get good results due to the roughness of the mica edge surface over the large

SC
contact area.

To overcome the surface roughness effect, Gupta and Miller [62] directly used a

NU
pyramid-shaped silicon nitride tip as the probe to map the surface potential of two kaolinite
MA
basal faces (silica face and alumina face). The silicon tip was modeled as a cone with a

spherical cap at its apex. Then, the DLVO forces between tip and flat surface could be
D

reasonably approximated as the sum of the conical-substrate interaction force and the
TE

cap-substrate interaction force [214]. The isoelectric point of the silica tetrahedral face was
P

lower than that of the alumina octahedral face. Yan et al. [215] further used this approach to
CE

explore the surface potentials of both basal and edge faces of talc and muscovite. The
AC

ultramicrotome cutting technique was applied to prepare the edge face. The surface potentials

of basal faces of both talc and muscovite were always more negatively charged than that of

edge faces in the range of experimental pH (5.6-10.1) (Fig. 17). This is due to the different

charging mechanisms between basal and edge faces. The permanent negative charge of basal

faces is attributed to fixed isomorphic substitution while protonation-deprotonation reactions

are responsible for the pH-dependent properties of the edge face. The anisotropic surface

potentials of scheelite crystal and molybdenite were also studied using AFM force

measurements [58, 216]. The 101 face of scheelite crystal was the most negatively charged

surface, followed by 112 and 001 faces. For molybdenite, it was found that the surface
41
ACCEPTED MANUSCRIPT

potentials of both edge and basal faces are highly pH-dependent.

A B

T
R IP
SC
NU
C MA D
D
TE

Fig. 17 Force curves between AFM tip and different faces in 1 mM KCl solution: A, talc basal
P

plane; B, talc edge plane; C, muscovite basal plane; D, muscovite edge plane. Symbols
CE

represent experimental data and the solid lines represent theoretical fits. Adapted with
AC

permission from [215], Copyright 2011, American Chemical Society.

One can use a hydrophobized probe to map the surface hydrophobicity by fitting the force

profile using extended DLVO theory with an additional hydrophobic force taken into

consideration. A high attractive force corresponds to high surface hydrophobicity (large local

water contact angle). Lu et al. [58] used a single exponential function to fit the experimental

hydrophobic force curve between an OTS-coated tip and a molybdenite surface in 10 mM

NaCl solution at various pH (Fig. 18). An attractive hydrophobic force at all pH ranges was

observed. However, this was not the case between the tip and edge surface. This illustrated

42
ACCEPTED MANUSCRIPT

that the basal surface of molybdenite shows a hydrophobic character, while the edge surface is

hydrophilic. This observation is consistent with the contact angle results. Xie et al. [217]

adopted the same approach to map the sphalerite surface hydrophobicity before and after

T
IP
CuSO4 activation and amyl xanthate conditioning. Amyl xanthate adsorption increased surface

R
hydrophobicity.

SC
NU
MA
D
TE

Fig. 18 Hydrophobic force between OTS-coated tip and (a) face and (b) edge of molybdenite

in 10 mM NaCl solution. The dash line is the theoretical fitting result. Adapted with
P
CE

permission from [58], Copyright 2015, American Chemical Society.


AC

Colloid probe AFM can also be applied to characterize hydrophobic surface heterogeneities

on a micro scale. The colloid probe-sample interaction [218] and parameters derived from

force-distance curves [219] in the context of flotation or other processes like filtration can be

obtained. Rudolph and Peuker [220] combined hydrophobic colloid probe AFM with Raman

spectroscopy to evaluate the wettability of minerals in finely inter-grown ore in situ. The type

of mineral phase, where force measurements were conduct, could be identified by Raman

microscopy. They proposed that adhesion force mapping is a more precise indicator for the

floatability of practical minerals. The combined colloid probe AFM and Raman technique will

also be a powerful tool to investigate mineral-reagent interactions in flotation systems.

43
ACCEPTED MANUSCRIPT

7. Conclusions and perspectives

AFM is an established tool in mineral flotation used both for imaging and a surface force

sensing. It can be successfully used to image the minerals, to characterize the water structure

T
IP
at mineral-water interface, to study the adsorption of flotation reagents on mineral surfaces, to

R
measure inter-particle forces directly, and characterize bubble-particle interaction. Detailed

SC
information at the nanoscale has been obtained, greatly assisting the fundamental

understanding of flotation. One challenge is to measure precise shapes of water/air interfaces

NU
in particle-bubble experiments due to the absence of hard contact point.
MA
The main drawback of AFM is the absence of a chemical imaging capability. It is necessary to

develop new AFM-based techniques that have the ability to analyze both surface morphology
D

and chemical composition in the contact area. Optical spectroscopy analysis will be the
TE

optimal choice. Recently, significant progress has been achieved on the hybrid AFM
P

techniques such as the combination of AFM and infrared spectroscopy (IR) and Raman
CE

spectroscopy, which show great potentials in flotation research [221-223]. For AFM-IR, the
AC

AFM tip itself acts as the IR detector and thus can overcome the spatial resolution limits of

conventional IR due to the existence of optical diffraction limitation (Fig. 19). IR laser is

pre-set to a wavelength corresponding to an absorbing wavelength or quickly scanned in

wavelength and focused onto the AFM tip region. By using metallic AFM tips, a local

enhancement of the electromagnetic field at the tip end can be achieved. The IR absorption of

the substrate will cause a local thermal expansion and an additional dipolar interaction

between tip and sample. By measuring the response of the cantilever to IR absorption as a

function of wavelength, IR absorption spectra of a nano-scale region are obtained. The IR

laser can be also fixed at a constant wavelength and IR absorption as a function of position
44
ACCEPTED MANUSCRIPT

across the sample to create chemical images that show the distribution of chemical species

[221]. AFM-IR is typically performed by detecting the oscillation amplitude of the cantilever

as a function of sample position or wavelength. Detailed information on AFM-IR can be seen

T
IP
in [221]. Fig. 20 shows a resonance-enhanced AFM-IR image of a self-assembled monolayer

of alkyl thiol ethoxylate (PEG) on gold surface. The IR absorption band at 1340 cm−1 was

R
SC
attributed to a CH2 peak, and the AFM-IR image directly confirmed the location of the PEG

island regions.

NU
MA
D
TE

Fig. 19 Schematic diagram of AFM-IR. IR laser is focused on a sample near the tip of an
P
CE

AFM. The AFM tip is used as a local detector of IR absorption. Adapted with permission

from [221], Copyright 2016, American Chemical Society.


AC

Fig. 20 (a) AFM topographic image of a self-assembled PEG monolayer on gold (b) AFM-IR
45
ACCEPTED MANUSCRIPT

absorption image at 1340 cm−1 peak, corresponding to the CH2 peak as shown in the (c)

AFM-IR spectrum of the monolayer. Image reprinted with permission, Copyright 2014,

Anasys Instruments

T
IP
Raman spectroscopy, complementary to IR, is an alternative tool for identifying chemical

R
species. The combination of AFM and Raman also offers the opportunity to obtain both

SC
surface morphology and chemical composition of the sample [222]. In this context, more

complementary information on mineral surface/mineral-water interface can be obtained. For

NU
example, the specific adsorption of flotation reagents on mineral surface can be identified
MA
successfully. Note that the chemical image obtained by AFM-IR is always affected by surface

characteristic such as roughness. Tip-enhanced IR spectroscopy is difficult to operate in


D

aqueous medium due to the strong IR absorption bands of water. By contrast, AFM-Raman
TE

may allow the experiments conducted in aqueous environments, while the reproducibility of
P

the Raman signal enhancement is still a problem. Nevertheless, the complementary


CE

characterization of chemical composition using optical spectroscopy for AFM topography


AC

imaging and the synchronous measurement of the force and distance involving deformable

bubble as a force sensor will become an active area over the coming years and further shed

new light on flotation mechanism.

Acknowledgments

This research was supported by the National Nature Science Foundation of China (grant no.

51774286, 51574236, 51574240), a project funded by the China Postdoctoral Science

Foundation (2015T80606, 2014M550317) for which the authors express their appreciation.

Yaowen Xing also appreciates China Scholarship Council for the financial support for their

46
ACCEPTED MANUSCRIPT

research stay at Max Planck Institute for Polymer Research. HJB also acknowledges support

by the Max Planck Center for Complex Fluid Dynamics Fluid Dynamics of Complexity.

T
Symbols

IP
F force

R
g acceleration of gravity

SC
h thickness of the intervening water film or separation distance

NU
𝑝 hydrodynamic pressure difference

𝑅 radius of particle
MA
Rb radius of bubble before deformation

𝑟 radial distance from the center of the film


D
TE

 the distance the particle moved from its equilibrium position


P
CE

𝜌 density

𝜅 capillary length
AC

∆𝐷 AFM piezo displacement

𝛾 surface tension

𝜃 water contact angle

𝜇 viscosity

𝛱 disjoining pressure

References

[1] Nguyen AV, Schulze HJ. Colloidal Science of Flotation. New York, America: Marcel

Dekker Inc; 2004.

47
ACCEPTED MANUSCRIPT

[2] Ejtemaei M, Gharabaghi M, Irannajad M. A review of zinc oxide mineral beneficiation

using flotation method. Advances in Colloid and Interface Science, 2014;206:68-78.

[3] Aghazadeh S, Mousavinezhad SK, Gharabaghi M. Chemical and colloidal aspects of

T
IP
collectorless flotation behavior of sulfide and non-sulfide minerals. Advances in Colloid and

R
Interface Science, 2015;225:203-17.

SC
[4] Wu ZJ, Wang XM, Liu HN, Zhang HF, Miller JD. Some physicochemical aspects of

water-soluble mineral flotation. Advances in Colloid and Interface Science,

2016;235:190-200.
NU
MA
[5] Xing YW, Gui XH, Liu JT, Cao YJ, Lu Y. Effects of energy input on the laboratory column

flotation of fine coal. Separation Science Technology. 2015;50:2559-67.


D

[6] Xing YW, Gui XH, Cao YJ, Wang DP, Zhang HJ. Clean low-rank-coal purification
TE

technique combining cyclonic-static microbubble flotation column with collector


P

emulsification. Journal of Cleaner Production. 2017; 153:657-72.


CE

[7] Xing YW, Gui XH, Cao YJ, Wang YW, Xu MD, Wang DY, et al. Effect of compound
AC

collector and blending frother on froth stability and flotation performance of oxidized coal.

Powder Technology. 2017;305:166-73.

[8] Rubio J, Souza ML, Smith RW. Overview of flotation as a wastewater treatment technique.

Minerals Engineering. 2002;15:139-55.

[9] Cai X, Chen J, Liu M, Ji Y, An S. Numerical studies on dynamic characteristics of

oil-water separation in loop flotation column using a population balance model. Separation

and Purification Technology. 2017;176:134-44.

[10] Rao F, Liu Q. Froth treatment in athabasca oil sands bitumen recovery process: a review.

Energy & Fuels. 2013;27:7199-207.


48
ACCEPTED MANUSCRIPT

[11] Kasongo T, Zhou Z, Xu ZH, Masliyah J. Effect of clays and calcium ions on bitumen

extraction from Athabasca oil sands using flotation. The Canadian Journal of Chemical

Engineering. 2000;78:674-81.

T
IP
[12] Zhou F, Yan C, Wang H, Zhou S, Liang H. The result of surfactants on froth flotation of

R
unburned carbon from coal fly ash. Fuel. 2017;190:182-8.

SC
[13] Altun NE, Xiao CF, Hwang JY. Separation of unburned carbon from fly ash using a

concurrent flotation column. Fuel Processing Technology. 2009;90:1464-70.

NU
[14] Bartonova L. Unburned carbon from coal combustion ash: An overview. Fuel Processing
MA
Technology. 2015;134:136-58.

[15] Finch JA, Hardie CA. An example of innovation from the waste management industry:
D

Deinking flotation cells. Minerals Engineering. 1999;12:467-75.


TE

[16] Vashisth S, Bennington CPJ, Grace JR, Kerekes RJ. Column flotation deinking:
P

State-of-the-art and opportunities. Resources, Conservation and Recycling. 2011;55:1154-77.


CE

[17] Hosseini S, Forssberg E. XPS & FTIR study of adsorption characteristics using
AC

cationic and anionic collectors on smithsonite. Journal of Minerals and Materials

Characterization and Engineering. 2006;5(1): 21-45.

[18] Beattie DA., Kempson IM, Fan LJ, Skinner WM. Synchrotron XPS studies of collector

adsorption and co-adsorption on gold and gold: silver alloy surfaces. International Journal of

Mineral Processing. 2009; 92(3): 162-68.

[19] Giesekke, EW. A review of spectroscopic techniques applied to the study of interactions

between minerals and reagents in flotation systems. International Journal of Mineral

Processing. 1983; 11(1): 19-56.

[20] Hosseinpour S, Tang F, Wang F, Livingstone RA, Schlegel SJ, Ohto T, Bonn M, Nagata Y,
49
ACCEPTED MANUSCRIPT

Backus EHG. Chemisorbed and physisorbed water at the TiO2/water interface. Journal of

Physical Chemistry Letters. 2017; 8(10): 2195-9.

[21] Schaefer J, Gonella G, Bonn M, Backus EHG. Surface-specific vibrational

T
IP
spectroscopy of the water/silica interface: screening and interference. Physical Chemistry

R
Chemical Physics. 2017; 19(25): 16875-80.

SC
[22] Xu Z, Liu J, Choung J, Zhou Z. Electrokinetic study of clay interactions with coal in

flotation. International Journal of Mineral Processing. 2003;68:183-96.

NU
[23] Liu J, Zhou Z, Xu Z, Masliyah J. Bitumen–clay interactions in aqueous media studied by
MA
zeta potential distribution measurement. Journal of Colloid and Interface Science.

2002;252:409-18.
D

[24] Wu C, Wang L, Harbottle D, Masliyah J, Xu Z. Studying bubble–particle interactions by


TE

zeta potential distribution analysis. Journal of Colloid and Interface Science.


P

2015;449:399-408.
CE

[25] Ralston J, Fornasiero D, Hayes R. Bubble–particle attachment and detachment in


AC

flotation. International Journal of Mineral Processing. 1999;56:133-64.

[26] Nguyen A, Evans G, Schulze H. Prediction of van der Waals interaction in

bubble–particle attachment in flotation. International Journal of Mineral Processing.

2001;61:155-69.

[27] Dai Z, Fornasiero D, Ralston J. Particle–bubble collision models-a review. Advances in

Colloid and Interface Science. 2000;85:231-56.

[28] Koh P, Schwarz M. CFD modelling of bubble–particle collision rates and efficiencies in

a flotation cell. Minerals Engineering. 2003;16:1055-9.

[29] Chau T T, Bruckard WJ, Koh P, Nguyen AV. A review of factors that affect contact angle
50
ACCEPTED MANUSCRIPT

and implications for flotation practice. Advances in Colloid and Interface Science. 2009;

150(2): 106-15.

[30] Butt HJ, Kappl, M. Surface and Interfacial Forces. Weinheim: WILEY-VCH; 2010.

T
IP
[31] Israelachvili JN. Intermolecular and Surface Forces. Third Edition USA: Academic press;

R
2011.

SC
[32] Israelachvili JN, Pashley RM. Molecular layering of water at surfaces and origin of

repulsive hydration forces. Nature. 1983;306:249-250.33.

NU
[33] Butt HJ. Measuring electrostatic, van der Waals, and hydration forces in electrolyte
MA
solutions with an atomic force microscope. Biophysical Journal. 1991;60:1438-44.

[34] Israelachvili J, Wennerström H. Role of hydration and water structure in biological and
D

colloidal interactions. Nature. 1996; 379(6562): 219-25.


TE

[35] Israelachvili JN, Pashley RM. The hydrophobic interaction is long-range, decaying
P

exponentially with distance. Nature. 1982;300:341-2.


CE

[36] Israelachvili JN, Pashley RM. Measurement of the hydrophobic interaction between two
AC

hydrophobic surfaces in aqueous electrolyte solutions. Journal of Colloid and Interface

Science. 1984;98:500-14.

[37] Christenson HK, Claesson PM. Direct measurements of the force between hydrophobic

surfaces in water. Advances in Colloid and Interface Science. 2001;91:391-436.

[38] Meyer EE, Rosenberg KJ, Israelachvili J. Recent progress in understanding hydrophobic

interactions. Proceedings of the National Academy of Sciences. 2006;103:15739-15746.

[39] Laskowski JK, Kitchener JA. The hydrophilic-hydrophobic transition on silica. Journal

of Colloid and Interface Science. 1969;29:670-9.

[40] Blake TD, Kitchener JA. Stability of aqueous films on hydrophobic methylated silica.
51
ACCEPTED MANUSCRIPT

Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed

Phases. 1972;68:1435-42.

[41] Xing Y, Gui X, Pan L, Pinchasik B-E, Cao Y, Liu J, Kappl K, Butt HJ. Recent

T
IP
experimental advances for understanding bubble-particle attachment in flotation. Advances in

R
Colloid and Interface Science. 2017; 246:105-32.

SC
[42] Xing Y, Gui X, Cao Y. The hydrophobic force for bubble-particle attachment in

flotation-a brief review. Physical Chemistry Chemical Physics. 2017. DOI:

10.1039/C7CP03856A.
NU
MA
[43] Binnig G, Quate CF, Gerber C. Atomic force microscope. Physical Review Letters.

1986;56: 930-933.
D

[44] Drake B, Prater CB, Weisenhorn AL, Gould SAC, Albrecht TR, Quate CF, Cannell DS,
TE

Hansma HG, Hansma PK. Imaging crystals, polymers, and processes in water with the atomic
P

force microscope. Science. 1989;243:1586-1589.


CE

[45] Butt HJ, Berger R, Bonaccurso E, Chen Y, Wang J. Impact of atomic force microscopy
AC

on interface and colloid science. Advances in Colloid and Interface Science.

2007;133:91-104.

[46] Rugar D, Hansma P. Atomic force microscopy. Physics Today. 1990; 43(10): 23-30.

[47] Meyer E. Atomic force microscopy. Progress in Surface Science. 1992; 41(1): 3-49.

[48] Jalili N, Laxminarayana K. A review of atomic force microscopy imaging systems:

application to molecular metrology and biological sciences. Mechatronics. 2004; 14(8):

907-45.

[49] Maver U, Velnar T, Gaberšček M, Planinšek O, Finšgar, M. Recent progressive use of

atomic force microscopy in biomedical applications. TrAC Trends in Analytical Chemistry.


52
ACCEPTED MANUSCRIPT

2016; 80: 96-111.

[50] Albrecht TR, Grutter P, Horne D, Rugar D. Frequency modulation detection using

high‐ Q cantilevers for enhanced force microscope sensitivity. Journal of Applied Physics.

T
IP
1991; 69(2): 668-673.

R
[51] Jarvis SP, Ishida T, Uchihashi T, Nakayama Y, Tokumoto H. Frequency modulation

SC
detection atomic force microscopy in the liquid environment. Applied Physics A. 2001; 72(1):

S129-S132.

NU
[52] Ducker WA, Senden TJ, Pashley RM. Direct measurement of colloidal forces using an
MA
atomic force microscope. Nature. 1991;353:239.

[53] Butt HJ. Measuring electrostatic, van der Waals, and hydration forces in electrolyte
D

solutions with an atomic force microscope. Biophysical Journal. 1991;60:1438-44.


TE

[54] Butt HJ, Cappella B, Kappl M. Force measurements with the atomic force microscope:
P

Technique, interpretation and applications. Surface Science Reports. 2005;59:1-152.


CE

[55] Leiro J, Torhola M, Laajalehto K. The AFM method in studies of muscovite mica and
AC

galena surfaces. Journal of Physics and Chemistry of Solids. 2017;100:40-4.

[56] Beaussart A, Parkinson L, Mierczynska-Vasilev A, Beattie DA. Adsorption of modified

dextrins on molybdenite: AFM imaging, contact angle, and flotation studies. Journal of

Colloid and Interface Science. 2012;368:608-15.

[57] Xie L, Wang J, Shi C, Huang J, Zhang H, Liu Q, Zeng HB. Probing surface interactions

of electrochemically active galena mineral surface using atomic force microscopy. The

Journal of Physical Chemistry C. 2016;120:22433-42.

[58] Lu Z, Liu Q, Xu Z, Zeng H. Probing anisotropic surface properties of molybdenite by

direct force measurements. Langmuir. 2015;31:11409-18.


53
ACCEPTED MANUSCRIPT

[59] Gupta V, Hampton MA, Nguyen AV, Miller JD. Crystal lattice imaging of the silica and

alumina faces of kaolinite using atomic force microscopy. Journal of Colloid and Interface

Science. 2010;352:75-80.

T
IP
[60] Hampton MA, Plackowski C, Nguyen AV. Physical and chemical analysis of elemental

R
sulfur formation during galena surface oxidation. Langmuir. 2011;27:4190-4201.

SC
[61] Siretanu I, van den Ende D, Mugele F., Atomic structure and surface defects at

mineral-water interfaces probed by in situ atomic force microscopy. Nanoscale. 2016; 8(15):

8220-27.
NU
MA
[62] Gupta V, Miller JD. Surface force measurements at the basal planes of ordered kaolinite

particles. Journal of Colloid and Interface Science. 2010;344:362-71.


D

[63] Hayes RA, Ralston J. The collectorless flotation and separation of sulphide minerals by
TE

Eh control. International Journal of Mineral Processing. 1988; 23(1-2): 55-84.


P

[64] Hayes RA, Price DM, Ralston J, Smith RW. Collectorless flotation of sulphide minerals.
CE

Mineral Procesing and Extractive Metallurgy Review. 1987; 2(3): 203-34.


AC

[65] Bruening F, Cohen A. Measuring surface properties and oxidation of coal macerals using

the atomic force microscope. International Journal of Coal Geology. 2005;63:195-204.

[66] Morga R. Changes of semifusinite and fusinite surface roughness during heat treatment

determined by atomic force microscopy. International Journal of Coal Geology.

2011;88:218-26.

[67] Pan JN, Zhu HT, Hou QL, Wang HC, Wang S. Macromolecular and pore structures of

Chinese tectonically deformed coal studied by atomic force microscopy. Fuel.

2015;139:94-101.

[68] Wu D, Liu GJ, Sun RY, Chen SC. Influences of magmatic intrusion on the
54
ACCEPTED MANUSCRIPT

macromolecular and pore structures of coal: Evidences from Raman spectroscopy and atomic

force microscopy. Fuel. 2014;119:191-201.

[69] Drosthansen W. Effects of vicinal water on colloidal stability and sedimentation

T
IP
processes. Journal of Colloid and Interface Science. 1977;58:251-62.

R
[70] Du H, Miller JD. A molecular dynamics simulation study of water structure and

SC
adsorption states at talc surfaces. International Journal of Mineral Processing.

2007;84:172-84.

NU
[71] Gao J, Szoszkiewicz R, Landman U, Riedo E. Structured and viscous water in
MA
subnanometer gaps. Physical Review B. 2007;75:115415.

[72] Fenter P, Lee SS. Hydration layer structure at solid-water interfaces. MRS Bulletin.
D

2014;39:1056-61.
TE

[73] Marutschke C, Walters D, Cleveland J, Hermes I, Bechstein R, Kühnle A.


P

Three-dimensional hydration layer mapping on the (10.4) surface of calcite using amplitude
CE

modulation atomic force microscopy. Nanotechnology. 2014;25:335703.


AC

[74] Miyazawa K, Watkins M, Shluger A, Fukuma T. Influence of ions on two-dimensional

and three-dimensional atomic force microscopy at fluorite-water interfaces. Nanotechnology.

2017;28:245701.

[75] Raviv U, Laurat P, Klein J. Fluidity of water confined to subnanometre films. Nature.

2001;413:51-54.

[76] Raviv U, Perkin S, Laurat P, Klein J. Fluidity of water confined down to subnanometer

films. Langmuir. 2004;20:5322-32.

[77] Björneholm O, Hansen MH, Hodgson A, Liu LM, Limmer DT, Michaelides A, et al.

Water at Interfaces. Chemical Reviews. 2016; 116(13):7698.


55
ACCEPTED MANUSCRIPT

[78] Söngen H, Marutschke C, Spijker P, Holmgren E, Hermes IM, Bechstein R, Klassen S,

Tracey T, Foster A, Kühnle A. Chemical Identification at the Solid-Liquid Interface. Langmuir.

2017; 33(1): 125-9.

T
IP
[79] Fukuma T, Reischl B, Kobayashi N, Spijker P, Canova FF, Miyazawa K, Foster A.

R
Mechanism of atomic force microscopy imaging of three-dimensional hydration structures at

SC
a solid-liquid interface. Physical Review B. 2015;92:155412.

[80] Mezger M, Reichert H, Schöder S, Okasinski J, Schröder H, Dosch H, Palms D, Ralston

NU
J, Honkimäki V. High-resolution in situ x-ray study of the hydrophobic gap at the
MA
water-octadecyl-trichlorosilane interface. Proceedings of the National Academy of Sciences.

2006;103:18401-04.
D

[81] Doshi DA, Watkins EB, Israelachvili JN, Majewski J. Reduced water density at
TE

hydrophobic surfaces: Effect of dissolved gases. Proceedings of the National Academy of


P

Sciences. 2005;102:9458-62.
CE

[82] Jensen TR, Jensen MØ, Reitzel N, Balashev K, Peters GH, Kjaer K, Bjørnholm T. Water
AC

in contact with extended hydrophobic surfaces: Direct evidence of weak dewetting. Physical

Review Letters. 2003;90:086101.

[83] Fukuma T, Kobayashi K, Matsushige K, Yamada H. True atomic resolution in liquid by

frequency-modulation atomic force microscopy. Applied Physics Letters. 2005;87:034101.

[84] Zachariah Z, Espinosa-Marzal RM, Spencer ND, Heuberger MP. Stepwise collapse of

highly overlapping electrical double layers. Physical Chemistry Chemical Physics.

2016;18:24417-27.

[85] Fukuma T, Ueda Y, Yoshioka S, Asakawa H. Atomic-scale distribution of water

molecules at the mica-water interface visualized by three-dimensional scanning force


56
ACCEPTED MANUSCRIPT

microscopy. Physical Review Letters. 2010;104:016101.

[86] Imada H, Kimura K, Onishi H. Water and 2-propanol structured on calcite (104) probed

by frequency-modulation atomic force microscopy. Langmuir. 2013;29:10744-51.

T
IP
[87] Sader JE, Jarvis SP. Accurate formulas for interaction force and energy in frequency

R
modulation force spectroscopy. Applied Physics Letters, 2004; 84(10): 1801-03.

SC
[88] Welker J, Illek E, Giessibl FJ. Analysis of force-deconvolution methods in

frequency-modulation atomic force microscopy. Beilstein Journal of Nanotechnology. 2012; 3:

238-48.
NU
MA
[89] Söngen H, Bechstein R, Kühnle A. Quantitative atomic force microscopy. Journal of

Physics-Condensed Matter, 2017; 29(27):274001.


D

[90] Watkins M, Shluger AL. Mechanism of contrast formation in atomic force microscopy in
TE

water. Physical Review Letters. 2010;105: 196101.


P

[91] Miyazawa K, Kobayashi N, Watkins M, Shluger AL, Amano K, Fukuma T. A


CE

relationship between three-dimensional surface hydration structures and force distribution


AC

measured by atomic force microscopy. Nanoscale. 2016;8:7334-42.

[92] Fukuma T, Onishi K, Kobayashi N, Matsuki A, Asakawa H. Atomic-resolution imaging

in liquid by frequency modulation atomic force microscopy using small cantilevers with

megahertz-order resonance frequencies. Nanotechnology. 2012;23: 135706.

[93] Weber SA, Kilpatrick JI, Brosnan TM, Jarvis SP, Rodriguez BJ. High viscosity

environments: an unexpected route to obtain true atomic resolution with atomic force

microscopy. Nanotechnology. 2014; 25(17):175701.

[94] Söngen H, Nalbach M, Adam H, Kühnle A. Three-dimensional atomic force microscopy

mapping at the solid-liquid interface with fast and flexible data acquisition. Review of
57
ACCEPTED MANUSCRIPT

Scientific Instruments. 2016;87: 063704.

[95] Lohse D, Zhang XH. Surface nanobubbles and nanodroplets. Reviews of Modern Physics.

2015; 87:981-1035.

T
IP
[96] Zhang XH, Chan DYC, Wang DY, Maeda N. Stability of interfacial nanobubbles.

R
Langmuir. 2013;29:1017-23.

SC
[97] Schubert H. Nanobubbles, hydrophobic effect, heterocoagulation and hydrodynamics in

flotation. International Journal of Mineral Processing. 2005;78:11-21.

NU
[98] Sobhy A, Tao D. Nanobubble column flotation of fine coal particles and associated
MA
fundamentals. International Journal of Mineral processing. 2013;124:109-16.

[99] Parker J, Claesson P, Attard P. Bubbles, cavities, and the long-ranged attraction between
D

hydrophobic surfaces. Journal of Physical Chemistry. 1994;98:8468-80.


TE

[100] Ishida N, Inoue T, Miyahara M, Higashitani K. Nano bubbles on a hydrophobic surface


P

in water observed by tapping-mode atomic force microscopy. Langmuir. 2000;16:6377-80.


CE

[101] Lou ST, Ouyang ZQ, Zhang Y, Li XJ, Hu J, Li MQ, Yang FJ. Nanobubbles on solid
AC

surface imaged by atomic force microscopy. Journal of Vacuum Science & Technology B:

Microelectronics and Nanometer Structures Processing, Measurement, and Phenomena.

2000;18:2573-5.

[102] Peng H, Hampton MA, Nguyen AV. Nanobubbles do not sit alone at the solid-liquid

interface. Langmuir. 2013;29:6123-30.

[103] Zhang XH, Maeda N, Craig VSJ. Physical properties of nanobubbles on hydrophobic

surfaces in water and aqueous solutions. Langmuir. 2006;22:5025-35.

[104] Walczyk W, Schönherr H. Dimensions and the profile of surface nanobubbles:

tip–nanobubble interactions and nanobubble deformation in atomic force microscopy.


58
ACCEPTED MANUSCRIPT

Langmuir. 2014;30:11955-65.

[105] Walczyk W, Schon PM, Schönherr H. The effect of PeakForce tapping mode AFM

imaging on the apparent shape of surface nanobubbles. Journal of Physics: Condensed Matter.

T
IP
2013;25:184005.

R
[106] Ko HC, Hsu WH, Yang CW, Fang CK, Lu YH, Hwang IS. High-resolution

SC
characterization of preferential gas adsorption at the graphene-water interface. Langmuir.

2016;32:11164-71.

NU
[107] Walczyk W, Schönherr H. Characterization of the interaction between AFM tips and
MA
surface nanobubbles. Langmuir. 2014;30:7112-26.

[108] Zhao B, Song Y, Wang S, Dai B, Zhang L, Dong Y, Lü J, Hu J. Mechanical mapping of


D

nanobubbles by PeakForce atomic force microscopy. Soft Matter. 2013;9:8837-43.


TE

[109] Schönherr H, Hain N, Walczyk W, Wesner D, Druzhinin SI. Surface nanobubbles


P

studied by atomic force microscopy techniques: Facts, fiction, and open questions. Japanese
CE

Journal of Applied Physics. 2016; 55:08NA01.


AC

[110] Miller JD, Hu YH, Veeramasuneni S, Lu YQ. In-situ detection of butane gas at a

hydrophobic silicon surface. Colloids and Surfaces A: Physicochemical and Engineering

Aspects. 1999;154:137-47.

[111] Chan CU, Ohl CD. Total-internal-reflection-fluorescence microscopy for the study of

nanobubble dynamics. Physical Review Letters. 2012;109:174501.

[112] Li M, Tonggu L, Zhan X, Mega TL, Wang L. Cryo-EM visualization of nanobubbles in

aqueous solutions. Langmuir. 2016;32:11111-5.

[113] Switkes M, Ruberti J. Rapid cryofixation/freeze fracture for the study of nanobubbles at

solid–liquid interfaces. Applied Physics Letters. 2004;84:4759-61.


59
ACCEPTED MANUSCRIPT

[114] Zhang XH. Quartz crystal microbalance study of the interfacial nanobubbles. Physical

Chemistry Chemical Physics. 2008;10:6842-8.

[115] Hain N, Wesner D, Druzhinin SI, Schönherr H. Surface nanobubbles studied by

T
IP
time-resolved fluorescence microscopy methods combined with AFM: The impact of surface

R
treatment on nanobubble nucleation. Langmuir. 2016;32:11155-63.

SC
[116] Pan G, He G, Zhang M, Zhou Q, Tyliszczak T, Tai R, et al. Nanobubbles at hydrophilic

particle-water interfaces. Langmuir. 2016;32:11133-7.

NU
[117] Palmer LA, Cookson D, Lamb RN. The relationship between nanobubbles and the
MA
hydrophobic force. Langmuir. 2011;27:144-7.

[118] Alheshibri M, Qian J, Jehannin M, Craig VS. A history of nanobubbles. Langmuir.


D

2016;32:11086-100.
TE

[119] Zhang XH, Zhang XD, Lou ST, Zhang ZX, Sun JL, Hu J. Degassing and temperature
P

effects on the formation of nanobubbles at the mica/water interface. Langmuir.


CE

2004;20(9):3813-5.
AC

[120] Zhang XH, Zhang X, Sun J, Zhang Z, Li G, Fang H, Xiao X, Zeng X, Hu J. Detection

of novel gaseous states at the highly oriented pyrolytic graphite-water interface. Langmuir.

2007;23:1778-83.

[121] van Limbeek MA, Seddon JR. Surface nanobubbles as a function of gas type. Langmuir.

2011;27:8694-9.

[122] Zhang X, Uddin MH, Yang H, Toikka G, Ducker W, Maeda N. Effects of surfactants on

the formation and the stability of interfacial nanobubbles. Langmuir. 2012;28:10471-7.

[123] Xu C, Peng S, Qiao GG, Gutowski V, Lohse D, Zhang X. Nanobubble formation on a

warmer substrate. Soft Matter. 2014;10:7857-64.


60
ACCEPTED MANUSCRIPT

[124] Seiedi O, Rahbar M, Nabipour M, Emadi MA, Ghatee MH, Ayatollahi S. Atomic force

microscopy (AFM) investigation on the surfactant wettability alteration mechanism of aged

mica mineral surfaces. Energy & Fuels. 2010;25:183-8.

T
IP
[125] Fielden ML, Claesson PM, Verrall RE. Investigating the adsorption of the gemini

R
surfactant “12-2-12” onto mica using atomic force microscopy and surface force apparatus

SC
measurements. Langmuir. 1999;15:3924-34.

[126] Dong J, Mao G. Direct study of C12E5 aggregation on mica by atomic force microscopy

NU
imaging and force measurements. Langmuir. 2000;16:6641-7.
MA
[127] Ferrari M, Ravera F, Viviani M, Liggieri L. Characterization of surfactant aggregates at

solid–liquid surfaces by atomic force microscopy. Colloids and Surfaces A: Physicochemical


D

and Engineering Aspects. 2004;249:63-7.


TE

[128] Chennakesavulu K, Raju GB, Prabhakar S, Nair CM, Murthy K. Adsorption of oleate
P

on fluorite surface as revealed by atomic force microscopy. International Journal of Mineral


CE

Processing. 2009;90:101-4.
AC

[129] Paiva P, Monte M, Simao R, Gaspar J. In situ AFM study of potassium oleate

adsorption and calcium precipitate formation on an apatite surface. Minerals Engineering.

2011;24:387-95.

[130] Manne, S. Molecular Organization of Surfactants at Solid-Liquid Interfaces. Science.

1995; 270(5241): 1480-2.

[131] Warr GG. Surfactant adsorbed layer structure at solid/solution interfaces: impact and

implications of AFM imaging studies. Current Opinion in Colloid & Interface Science. 2000;

5(1-2): 88-94.

61
ACCEPTED MANUSCRIPT

[132] Paruchuri VK, Nguyen AV, Miller JD. Zeta-potentials of self-assembled surface

micelles of ionic surfactants adsorbed at hydrophobic graphite surfaces. Colloids and Surfaces

a-Physicochemical and Engineering Aspects. 2004; 250(1-3): 519-26.

T
IP
[133] Paruchuri VK, Nalaskowski J, Shah DO, Miller JD. The effect of cosurfactants on

R
sodium dodecyl sulfate micellar structures at a graphite surface. Colloids and Surfaces

SC
a-Physicochemical and Engineering Aspects, 2006. 272(3): 157-63.

[134] Jaschke M, Butt HJ, Graub HE, Manne S. Surfactant Aggregates at a Metal Surface.

Langmuir. 1997; 13(13): 1381-84.


NU
MA
[135] Xie L, Wang J, Yuan D, Shi C, Cui X, Zhang H, Liu Q, Liu QX, Zeng H. Interaction

mechanisms between air bubble and molybdenite surface: Impact of solution salinity and
D

polymer adsorption. Langmuir. 2017;33:2353-61.


TE

[136] Mierczynska-Vasilev A, Beattie DA. Adsorption of tailored carboxymethyl cellulose


P

polymers on talc and chalcopyrite: Correlation between coverage, wettability, and flotation.
CE

Minerals Engineering. 2010;23:985-93.


AC

[137] Rief M, Oesterhelt F, Heymann B, Gaub HE. Single molecule force spectroscopy on

polysaccharides by atomic force microscopy. Science. 1997;275:1295-7.

[138] Neuman KC, Nagy A. Single-molecule force spectroscopy: optical tweezers, magnetic

tweezers and atomic force microscopy. Nature methods. 2008;5:491-505.

[139] Liu C, Shi W, Cui S, Wang Z, Zhang X. Force spectroscopy of polymers: Beyond single

chain mechanics. Current Opinion in Solid State and Materials Science. 2005;9:140-8.

[140] Zhang X, Liu C, Wang Z. Force spectroscopy of polymers: Studying on intramolecular

and intermolecular interactions in single molecular level. Polymer. 2008;49:3353-61.

[141] Sun W, Long J, Xu Z, Masliyah JH. Study of Al(OH)3-polyacrylamide-induced


62
ACCEPTED MANUSCRIPT

pelleting flocculation by single molecule force spectroscopy. Langmuir. 2008;24:14015-21.

[142] Long J, Xu Z, Masliyah JH. Adhesion of single polyelectrolyte molecules on silica,

mica, and bitumen surfaces. Langmuir. 2006;22:1652-9.

T
IP
[143] Pensini E, Yip CM, O’Carroll D, Sleep BE. Carboxymethyl cellulose binding to mineral

R
substrates: Characterization by atomic force microscopy–based Force spectroscopy and

SC
quartz-crystal microbalance with dissipation monitoring. Journal of Colloid and Interface

Science. 2013;402:58-67.

NU
[144] Fa K, Nguyen AV, Miller JD. Interaction of calcium dioleate collector colloids with
MA
calcite and fluorite surfaces as revealed by AFM force measurements and molecular dynamics

simulation. International Journal of Mineral Processing. 2006;81:166-77.


D

[145] Fa K, Jiang T, Nalaskowski J, Miller JD. Interaction forces between a calcium dioleate
TE

sphere and calcite/fluorite surfaces and their significance in flotation. Langmuir.


P

2003;19:10523-30.
CE

[146] Free ML, Miller JD. Kinetics of 18-carbon carboxylate adsorption at the fluorite surface.
AC

Langmuir. 1997;13:4377-82.

[147] Xing Y, Li C, Gui X, Cao Y. Interaction forces between paraffin/stearic acid and

fresh/oxidized coal particles measured by atomic force microscopy. Energy & Fuels.

2017;31:3305-12.

[148] Xing, Y., Gui XH, Cao YJ. Effect of calcium ion on coal flotation in the presence of

kaolinite clay. Energy & Fuels, 2016. 30(2): 1517-23.

[149] Gui XH, Xing YW, Rong GQ, Cao YJ, Liu JT. Interaction forces between coal and

kaolinite particles measured by atomic force microscopy. Powder Technology. 2016;301:

349-55.
63
ACCEPTED MANUSCRIPT

[150] Forbes E. Shear, selective and temperature responsive flocculation: A comparison of

fine particle flotation techniques. International Journal of Mineral Processing. 2011. 99(1-4):

1-10.

T
IP
[151] Rabinovich YI, Yoon RH. Use of atomic-force microscope for the measurements of

R
hydrophobic forces between silanated silica plate and glass sphere. Langmuir.

SC
1994;10:1903-09.

[152] Tyrrell JWG, Attard P. Atomic force microscope images of nanobubbles on a

NU
hydrophobic surface and corresponding force-separation data. Langmuir. 2002;18:160-7.
MA
[153] Hampton MA, Nguyen AV. Systematically altering the hydrophobic nanobubble

bridging capillary force from attractive to repulsive. Journal of Colloid and Interface Science.
D

2009;333:800-6.
TE

[154] Hampton MA, Donose BC, Nguyen AV. Effect of alcohol-water exchange and surface
P

scanning on nanobubbles and the attraction between hydrophobic surfaces. Journal of Colloid
CE

and Interface Science. 2008;325:267-74.


AC

[155] Christensen H, Claesson P. Cavitation and the interaction between macroscopic surfaces.

Science. 1988;239:390-392.

[156] Xing YW, Xu XH, Gui XH, Cao YJ, Xu MD. Effect of kaolinite and montmorillonite on

fine coal flotation. Fuel. 2017;195:284-9.

[157] Shrimali K, Jin J, Hassas BV, Wang X, Miller JD. The surface state of hematite and its

wetting characteristics. Journal of Colloid and Interface Science. 2016; 477: 16-24.

[158] Borkovec M, Papastavrou G. Interactions between solid surfaces with adsorbed

polyelectrolytes of opposite charge. Current Opinion in Colloid & Interface Science.

2008;13:429-37.
64
ACCEPTED MANUSCRIPT

[159] Borkovec M, Szilagyi I, Popa I, Finessi M, Sinha P, Maroni P, et al. Investigating forces

between charged particles in the presence of oppositely charged polyelectrolytes with the

multi-particle colloidal probe technique. Advances in Colloid and Interface Science.

T
IP
2012;179:85-98.

R
[160] Abraham T, Christendat D, Xu Z, Masliyah J, Gohy J-F, Jérôme R. Role of

SC
polyelectrolyte charge density in tuning colloidal forces. AIChE Journal. 2004;50:2613-26.

[161] Zhou Y, Gan Y, Wanless EJ, Jameson GJ, Franks GV. Interaction forces between silica

NU
surfaces in aqueous solutions of cationic polymeric flocculants: effect of polymer charge.
MA
Langmuir. 2008;24:10920-28.

[162] Ofori P, Nguyen AV, Firth B, McNally C, Hampton MA. The role of surface interaction
D

forces and mixing in enhanced dewatering of coal preparation tailings. Fuel. 2012;97:262-8.
TE

[163] Ruiz-Cabello FJM, Maroni P, Borkovec M. Direct measurements of forces between


P

different charged colloidal particles and their prediction by the theory of Derjaguin, Landau,
CE

Verwey, and Overbeek (DLVO). The Journal of Chemical Physics. 2013;138:234705.


AC

[164] Ivanova NO, Xu Z, Liu Q, Masliyah JH. Surface forces in unconventional oil

processing. Current Opinion in Colloid & Interface Science. 2017;27:63-73.

[165] Masliyah J, Czarnecki J, Xu Z. Handbook on theory and practice of bitumen recovery

from Athabasca oil sands. Theoretical Basis. 2011;1.

[166] Liu J, Xu Z, Masliyah J. Interaction forces in bitumen extraction from oil sands. Journal

of Colloid and Interface Science. 2005;287:507-20.

[167] Long J, Drelich J, Xu Z, Masliyah JH. Effect of operating temperature on water‐based

oil sands processing. The Canadian Journal of Chemical Engineering. 2007;85:726-38.

[168] Zhang Y, Ding M, Liu J, Jia W, Ren S. Studies on bitumen–silica interaction in


65
ACCEPTED MANUSCRIPT

surfactants and divalent cations solutions by atomic force microscopy. Colloids and Surfaces

A: Physicochemical and Engineering Aspects. 2015;482:241-7.

[169] Liu J, Xu Z, Masliyah J. Role of fine clays in bitumen extraction from oil sands. AIChE

T
IP
Journal. 2004;50:1917-27.

R
[170] Diao M, Taran E, Mahler S, Nguyen AV. A concise review of nanoscopic aspects of

SC
bioleaching bacteria-mineral interactions. Advances in Colloid and Interface Science.

2014;212:45-63.

NU
[171] Meyer RL, Zhou X, Tang L, Arpanaei A, Kingshott P, Besenbacher F. Immobilisation of
MA
living bacteria for AFM imaging under physiological conditions. Ultramicroscopy.

2010;110:1349-57.
D

[172] Bowen WR, Hilal N, Lovitt RW, Wright CJ. Direct measurement of the force of
TE

adhesion of a single biological cell using an atomic force microscope. Colloids and Surfaces
P

A: Physicochemical and Engineering Aspects. 1998;136:231-4.


CE

[173] Lower SK, Tadanier CJ, Hochella MF. Measuring interfacial and adhesion forces
AC

between bacteria and mineral surfaces with biological force microscopy. Geochimica et

Cosmochimica Acta. 2000;64:3133-9.

[174] Diao M, Nguyen TA, Taran E, Mahler S, Nguyen AV. Differences in adhesion of A.

thiooxidans and A. ferrooxidans on chalcopyrite as revealed by atomic force microscopy with

bacterial probes. Minerals Engineering. 2014;61:9-15.

[175] Chan DY, Klaseboer E, Manica R. Theory of non-equilibrium force measurements

involving deformable drops and bubbles. Advances in Colloid and Interface Science.

2011;165:70-90.

[176] Ducker WA, Xu Z, Israelachvili JN. Measurements of hydrophobic and DLVO forces in
66
ACCEPTED MANUSCRIPT

bubble-surface interactions in aqueous solutions. Langmuir. 1994;10:3279-89.

[177] Butt H-J. A technique for measuring the force between a colloidal particle in water and

a bubble. Journal of Colloid and Interface Science. 1994;166:109-17.

T
IP
[178] Tabor RF, Grieser F, Dagastine RR, Chan DY. Measurement and analysis of forces in

R
bubble and droplet systems using AFM. Journal of Colloid and Interface Science.

SC
2012;371:1-14.

[179] Johnson DJ, Miles NJ, Hilal N. Quantification of particle-bubble interactions using

NU
atomic force microscopy: A review. Advances in Colloid and Interface Science.
MA
2006;127:67-81.

[180] Nguyen AV, Nalaskowski J, Miller JD. A study of bubble-particle interaction using
D

atomic force microscopy. Minerals Engineering. 2003;16:1173-81.


TE

[181] Assemi S, Nguyen AV, Miller JD. Direct measurement of particle–bubble interaction
P

forces using atomic force microscopy. International Journal of Mineral Processing. 2008;
CE

89(1): 65-70.
AC

[182] Nguyen AV, Evans GM, Nalaskowski J, Miller JD. Hydrodynamic interaction between

an air bubble and a particle: atomic force microscopy measurements. Experimental Thermal

and Fluid Science. 2004;28:387-94.

[183] Ishida N. Direct measurement of hydrophobic particle-bubble interactions in aqueous

solutions by atomic force microscopy: Effect of particle hydrophobicity. Colloids and

Surfaces A: Physicochemical and Engineering Aspects. 2007;300:293-9.

[184] Fielden ML, Hayes RA, Ralston J. Surface and capillary forces affecting air bubble−

particle interactions in aqueous electrolyte. Langmuir. 1996;12:3721-27.

[185] Englert A, Krasowska M, Fornasiero D, Ralston J, Rubio J. Interaction force between an


67
ACCEPTED MANUSCRIPT

air bubble and a hydrophilic spherical particle in water, measured by the colloid probe

technique. International Journal of Mineral Processing. 2009;92:121-7.

[186] Taran E, Hampton MA, Nguyen AV, Attard P. Anomalous time effect on particle-bubble

T
IP
interactions studied by atomic force microscopy. Langmuir. 2009;25:2797-803.

R
[187] Scheludko A. Attachment of particles to a liquid surface (capillary theory of flotation).

SC
Journal of the Chemical Society Faraday Transactions. 1976; 72(4): 2815-28.

[188] James DF. The meniscus on the outside of a small circular cylinder. Journal of Fluid

Mechanics. 2006; 63(4): 657-64.


NU
MA
[189] O'Brien SBG. The Meniscus near a Small Sphere and Its Relationship to Line Pinning

of Contact Lines. Journal of Colloid & Interface Science. 1996; 183(1): 51–6.
D

[190] Chateau X, Pitois O. Quasistatic detachment of a sphere from a liquid interface. Journal
TE

of Colloid & Interface Science. 2003; 259(2): 346-53.


P

[191] Butt H, Gao N, Papadopoulos P, Steffen W, Kappl M, Berger R. Energy dissipation of


CE

moving drops on superhydrophobic and superoleophobic surfaces. Langmuir. 2017; 33(1).


AC

[192] Mcnamee CE, Kappl M, Butt HJ, Higashitani K, Graf K. Interfacial forces between a

silica particle and phosphatidylcholine monolayers at the air-water interface. Langmuir. 2010;

26(18): 14574-81.

[193] Mcnamee CE, Kappl M, Butt H, Ally J, Shigenobu H, Iwafuji Y, et al. Forces between a

monolayer at an air/water interface and a particle in solution: influence of the sign of the

surface charges and the subphase salt concentration. Soft Matter. 2011; 7(21): 10182-92.

[194] Anachkov SE, Lesov I, Zanini M, Kralchevsky PA, Denkov ND, Isa L. Particle

detachment from fluid interfaces: theory vs. experiments. Soft Matter. 2016; 12(36): 7632-43.

[195] Preuss M, Butt HJ. Direct measurement of particle-bubble interactions in aqueous


68
ACCEPTED MANUSCRIPT

electrolyte: Dependence on surfactant. Langmuir. 1998;14:3164-74.

[196] Preuss M, Butt HJ. Direct measurement of forces between particles and bubbles.

International Journal of Mineral Processing. 1999;56:99-115.

T
IP
[197] Nguyen AV, Nalaskowski J, Miller JD. The dynamic nature of contact angles as

R
measured by atomic force microscopy. Journal of Colloid and Interface Science. 2003; 262(1):

SC
303-6.

[198] Chan DY, Manor O, Connor JN, Horn RG. Soft matter: from shapes to forces on the

nanoscale. Soft Matter. 2008;4:471-4.


NU
MA
[199] Manica R, Hendrix MHW, Gupta R, Klaseboer E, Ohl CD, Chan DYC. Effects of

hydrodynamic film boundary conditions on bubble-wall impact. Soft Matter. 2013;9:9755-58.


D

[200] Shi C, Chan DYC, Liu QX, Zeng HB. Probing the hydrophobic interaction between air
TE

bubbles and partially hydrophobic surfaces using atomic force microscopy. The Journal of
P

Physical Chemistry C. 2014;118:25000-8.


CE

[201] Xie L, Shi C, Wang J, Huang J, Lu Q, Liu Q, Zeng H. Probing the interaction between
AC

air bubble and sphalerite mineral surface using atomic force microscope. Langmuir.

2015;31:2438-46.

[202] Clark SC, Walz JY, Ducker WA. Atomic force microscopy colloid-probe measurements

with explicit measurement of particle-solid separation. Langmuir. 2004; 20(18): 7616-22.

[203] McKee CT, Clark SC, Walz JY, Ducker WA. Relationship between scattered intensity

and separation for particles in an evanescent field. Langmuir. 2005;21:5783-9.

[204] McKee CT, Mosse WKJ, Ducker WA. Measurement of the absolute separation for

atomic force microscopy measurements in the presence of adsorbed polymer. Review of

scientific instruments. 2006;77: 053706.


69
ACCEPTED MANUSCRIPT

[205] McKee CT. Investigation of non-DLVO forces using an evanescent wave atomic force

microscope: Virginia Polytechnic Institute and State University; 2006.

[206] Shi C, Cui X, Xie L, Liu QX, Chan DYC, Israelachvili JN, Zeng H. Measuring forces

T
IP
and spatiotemporal evolution of thin water films between an air bubble and solid surfaces of

R
different hydrophobicity. ACS Nano. 2015;9:95-104.

SC
[207] Schönherr H. Forces and thin water film drainage in deformable asymmetric nanoscale

contacts. ACS Nano. 2015;9:12-5.

NU
[208] Yin X, Gupta V, Du H, Wang X, Miller JD. Surface charge and wetting characteristics
MA
of layered silicate minerals. Advances in Colloid and Interface Science. 2012;179:43-50.

[209] Hartley PG, Larson A, Scales PJ. Electrokinetic and direct force measurements between
D

silica and mica surfaces in dilute electrolyte solutions. Langmuir. 1997; 13(8): 2207-14.
TE

[210] Assemi S, Nalaskowski J, Miller JD, Johnson WP. Isoelectric point of fluorite by direct
P

force measurements using atomic force microscopy. Langmuir. 2006; 22(4): 1403-5.
CE

[211] Butt HJ. Measuring local surface charge densities in electrolyte solutions with a
AC

scanning force microscope. Biophysical Journal, 1992. 63(2): 578-82.

[212] And CR, Radmacher M. Mapping local electrostatic forces with the atomic force

microscope. Langmuir. 1997; 13(10): 2825-32.

[213] Zhao H, Bhattacharjee S, Chow R, Wallace D, Masliyah JH, Xu Z. Probing surface

charge potentials of clay basal planes and edges by direct force measurements. Langmuir.

2008;24:12899-910.

[214] Drelich J, Long J, Yeung A. Determining surface potential of the bitumen‐water

interface at nanoscale resolution using atomic force microscopy. The Canadian Journal of

Chemical Engineering. 2007;85:625-34.


70
ACCEPTED MANUSCRIPT

[215] Yan L, Englert AH, Masliyah JH, Xu Z. Determination of anisotropic surface

characteristics of different phyllosilicates by direct force measurements. Langmuir.

2011;27:12996-3007.

T
IP
[216] Gao Z, Hu Y, Sun W, Drelich JW. Surface-charge anisotropy of scheelite crystals.

R
Langmuir. 2016;32:6282-8.

SC
[217] Xie L, Wang J, Shi C, Cui X, Huang J, Zhang H, Liu Q, Liu QX, Zeng H. Mapping the

nanoscale heterogeneity of surface hydrophobicity on the sphalerite mineral. The Journal of

Physical Chemistry C. 2017;121:5620-8.


NU
MA
[218] Fritzsche J, Peuker UA. Particle adhesion on highly rough hydrophobic surfaces: The

distribution of interaction mechanisms. Colloids and Surfaces A: Physicochemical and


D

Engineering Aspects. 2014;459:166-71.


TE

[219] Rudolph M, Peuker UA. Hydrophobicity of minerals determined by atomic force


P

microscopy - A tool for flotation research. Chemie Ingenieur Technik. 2014;86:865-73.


CE

[220] Rudolph M, Peuker UA. Mapping hydrophobicity combining AFM and Raman
AC

spectroscopy. Minerals Engineering. 2014;66-68:181-90.

[221] Dazzi A, Prater CB. AFM-IR: Technology and applications in nanoscale infrared

spectroscopy and chemical imaging. Chemical Reviews. 2016; 117:5146-73.

[222] Fu WY, Zhang W. Hybrid AFM for nanoscale physicochemical characterization: Recent

development and emerging applications. Small. 2017; 13(11): 1603525.

[223] Eifert A, Kranz C. Hyphenating atomic force microscopy. Analytical Chemistry.

2014; 86(11): 5190-200.

71
ACCEPTED MANUSCRIPT

T
IP
R
SC
Graphical abstract NU
MA
D
P TE
CE
AC

72
ACCEPTED MANUSCRIPT

Highlights

● Advances in the application of AFM in mineral flotation are reviewed

● Mineral surface and mineral-water interface imaging are discussed

T
IP
● Reagent adsorption, inter-particle force, and bubble-particle interaction are reviewed

R
● Complementary chemical characterization and distance measurement are necessary

SC
NU
MA
D
P TE
CE
AC

73

You might also like